Abstract
This review provides a comprehensive examination of strategies aimed at advancing low-temperature electrolysis for sustainable hydrogen production. It begins by exploring the significance and challenges associated with water splitting, followed by an in-depth discussion on the fundamentals of electrochemical water splitting and crucial performance indicators, including reversible hydrogen electrode potential, specific and mass activities, overpotential, Tafel slope, stability and durability, and Faradaic and energy efficiencies. The article then extensively discusses various emerging strategies, such as decoupled water electrolysis, hybrid water electrolysis (including reagent-sacrificing, pollutant-degrading, and value-added types), tandem water electrolysis, microbial electrolysis cells (covering reactor configurations, electrode materials, microbial populations, and substrates), and the application of external stimuli like ultrasonic, magnetic, and super gravity fields. Additionally, the challenges and advancements in seawater electrolysis are reviewed, with a focus on electrocatalysts, seawater electrolyzers, and future directions. Furthermore, the article addresses current challenges in electrolysis and electrolyzer development, offering perspectives on the future of these techniques. By delving into these strategies, this review aims to contribute to the advancement of clean energy technologies and the transition towards a hydrogen-based economy.
1 Introduction
Shifting towards carbon-neutral energy sources is crucial for sustainable progress. While renewable sources like wind, solar, and tidal power show potential, their intermittent nature presents challenges for consistent energy supply. 1 Hydrogen gas emerges as a promising alternative due to its carbon-neutral properties and flexible production methods. Unlike renewables, hydrogen production offers diverse pathways and convenient storage and transport options. However, the primary method of hydrogen production involves extracting it from fossil fuels, resulting in carbon dioxide emissions. To fully harness the potential of environmentally friendly hydrogen energy, it’s essential to develop sustainable and cost-effective methods for green hydrogen production. 2 , 3
Electrochemical water splitting stands out as a promising avenue for converting electricity sourced from renewable energy into high-purity hydrogen fuel. 4 This process involves driving the electrochemical water splitting reaction, wherein the electricity generated from renewable sources is utilized to produce high-purity hydrogen. 5 , 6 This hydrogen can then be stored, transported, and utilized through hydrogen/oxygen fuel cells, which efficiently convert the chemical energy of hydrogen into electricity, with water being the sole by-product. 7 , 8 As a result, electrochemical water electrolysis is recognized as a pivotal technology for clean energy storage and conversion, facilitating the realization of the hydrogen economy. 4 This technology encompasses two main routes: low- and high-temperature water electrolysis. 8 Low-temperature electrolysis generally refers to processes operating below 200 °C. While this includes ambient temperatures, it also encompasses slightly elevated temperatures that are used to enhance efficiency. This is in contrast to high-temperature electrolysis, which typically occurs between 500 °C and 1,000 °C, such as in solid oxide electrolysis. Among the current low-temperature technologies are alkaline water electrolysis (AWE), proton exchange membrane water electrolysis (PEMWE), and anion exchange membrane water electrolysis (AEMWE). 9 Each of these methods offers unique advantages and applications in the context of hydrogen production and utilization. 8 , 10
Water splitting involves two crucial half-reactions: the hydrogen evolution reaction (HER) occurring at the cathode and the oxygen evolution reaction (OER) at the anode. Unfortunately, both the HER and the OER exhibit sluggish kinetics and significant energy barriers, necessitating voltage inputs much higher than the theoretical thermodynamic voltage for substantial H2 production rates. This results in relatively low energy efficiencies. 11 Consequently, electrocatalysts are indispensable to expedite kinetics and reduce voltage inputs. 12 , 13 , 14 Currently, precious metals like Pt and noble metal oxides such as RuO2 and IrO2 serve as the advanced electrocatalysts for HER and OER, respectively. 8 However, the high cost, limited durability, and scarcity of these noble metal catalysts severely hamper their extensive adoption in large-scale H2 production. Rapid ongoing research took place in the last decade to discover new electrocatalysts for efficient performance of both HER and OER, with reduced energy (overpotential). 9 Among these catalysts, doped materials such as transition metal phosphides, oxides, and nitrides have shown promise due to their abundance and tunable electronic properties. 15 Nanostructured catalysts, such as metal nanoparticles supported on conductive substrates, offer high surface area and active sites for catalytic reactions. 16 Porous materials, including metal-organic frameworks (MOFs) and covalent organic frameworks (COFs), provide excellent mass transport properties and catalytic activity. 17 , 18 Additionally, earth-abundant elements like iron, cobalt, and nickel have been extensively studied for their potential as low-cost alternatives to noble metals. 4 These advancements in catalyst design hold significant potential for overcoming the limitations of noble metal catalysts and enabling the widespread adoption of water electrolysis for hydrogen production. Furthermore, the integration of novel membrane materials, including proton exchange membranes (PEMs) and ion-conducting polymers, has enabled precise control of ion transport and minimized crossover effects, leading to improved electrolyzer efficiency. 16 Additionally, advancements in reactor design, such as flow-through and flow-by configurations, have optimized mass transport and electrolyte distribution, enhancing overall system performance. 19 , 20
Moreover, ongoing research efforts have also focused on exploring new strategies in electrolysis technologies to enhance durability, efficiency, and stability. 4 These include decoupled, 21 hybrid, 22 , 23 acid/alkaline asymmetric electrolyte, 24 and tandem 18 electrolysis approaches, which integrate multiple electrochemical processes to improve overall system performance and resource utilization. Additionally, microbial electrolysis cells (MECs) have garnered attention for their ability to produce hydrogen from organic matter using microorganisms as biocatalysts. 25 Furthermore, seawater electrolysis has emerged as a promising avenue for sustainable hydrogen production, leveraging abundant seawater resources for large-scale electrolysis operations. 26 , 27 , 28 , 29
This chapter serves as a comprehensive exploration of cutting-edge strategies for low-temperature electrolysis aimed at hydrogen production. We embark on a journey through the fundamental principles of electrochemical water splitting, elucidating the underlying mechanisms and performance indicators that govern the efficiency of electrolysis systems. Building upon this foundation, we delve into emerging strategies and advancements in electrolysis technology, spanning decoupled, hybrid, and tandem electrolysis approaches, as well as microbial electrolysis cells and seawater electrolysis.
2 Electrochemical water splitting
In the realm of low-temperature electrochemical water splitting, a common setup involves immersing two electrodes in an aqueous electrolyte and applying a voltage sufficient to trigger the separation of water into hydrogen (H2) and oxygen (O2), Eq. (1). To prevent the potentially explosive mixing of these product gases, a membrane separator is typically employed. These cell configurations have undergone extensive research over the years and demonstrate effective performance under both acidic and alkaline conditions, particularly when a stable power source is available. They can achieve current densities for hydrogen and oxygen production around 1 A/cm2 for acidic regime electrolysis and 0.2–0.5 A/cm2 for alkaline regime electrolysis. 13
The process of water electrolysis is the sum of two half-reactions: the HER and the OER. These half-equations exhibit slight variations depending on the pH level at which the electrolysis occurs (Figure 1a). At standard conditions (1 atm, 25 °C) and acidic electrolytes (pH 0), the HER and OER proceed as follows (all potentials are referenced against the standard hydrogen electrode, SHE): 4
in neutral electrolytes (pH 7),
in alkaline electrolytes (pH 14),

Electrochemical water electrolysis. (a) Thermodynamic potentials of HER and OER in aqueous electrolytes at various pH values under standard conditions (25 °C, 1 atm). (b) Thermodynamics of water splitting in relation to temperature at 0.1 MPa. From. 4
The water splitting reaction (Eq. (1)) is an endothermic process, described by ΔH 0 = ΔG 0 + TΔS 0 , [ΔG0: change in Gibbs free energy (237.1 kJ/mol), ΔH0: enthalpy change (285.8 kJ/mol), and TΔS0 (ΔQ) thermal energy (48.7 kJ/mol) under standard conditions (25 °C, 1 atm)]. Enthalpy, comprising both electrical and thermal components, signifies the total energy involved. Thus, under standard conditions, the water splitting reaction can be simplified as: 4
In low-temperature water electrolysis, heat can come from external sources or Joule heating through ionic and electric currents. Under standard conditions, the theoretical minimum cell voltage of electrolysis operation, i.e., the reversible cell voltage (ERE), is 1.23 V, calculated using E RE = ΔG 0 /(nF) , where n is the number of electrons transferred per reaction, i.e., n = 2 for reaction (1), and F is Faraday’s constant (96,485 C/mol). Nevertheless, ERE decreases as temperature increases. Within the temperature range of 0–1,000 °C, ERE ranges from 1.25 to 0.91 V (Figure 1b).
The thermoneutral cell voltage (E TN ) represents the minimum voltage required for electrolysis to occur in an ideal cell without heat integration. It is s 1.48 V under standard conditions, as calculated using E TN = ΔH 0 /(nF). In the case of high-temperature electrolysis, water is supplied as steam, necessitating external evaporation. This indicates that the total energy demand ( ΔH ) for the electrolysis reaction (including thermal energy, TΔS 0 ) is supplied electrically. According to Figure 1b, for liquid water below 100 °C, the E TN is approximately 1.47–1.48 V (284–286 kJ/molH2). In the temperature range of 100–1,000 °C, when steam is supplied, it reduces to 1.26–1.29 V (243–249 kJ/molH2). This implies that the minimum electrical energy consumption for steam electrolysis compared to liquid water electrolysis can be reduced by the heat of evaporation, which is 41.0 kJ/mol at ambient pressure. 30 The total energy consumption, including water evaporation, remains almost constant from 0 to 1,000 °C.
The exploration of water electrolysis technologies encompasses three primary methods: Alkaline Electrolysis Cells (AEC), Proton Exchange Membrane Electrolysis Cells (PEMEC), and Solid Oxide Electrolysis Cells (SOEC). The setup for each technology is depicted in Figure 2, with Table 1 offering a concise overview of component materials, performance indicators, and cost parameters.

Schematic representation of three electrolysis cell technologies. Reproduced with permission from. 31
Key characteristics of AEC, PEMEC; and SOEC. 31
AEC | PEMEC | SOEC | |
---|---|---|---|
Reactions | Cathode: 2H2O+ 2e– → H2 + 2OH− Anode: 2H2O → O2 + 4H+ + 4e– |
Cathode: 2H+ + 2e– → H2 Anode: 2H2O → O2 + 4H+ + 4e– |
Cathode: H2O+2e– → H2 + O2– Anode: O2− → ½O2 + 2e– |
Electrolyte | Aqueous potassium hydroxide (20–40 wt% KOH) 11 , 227 | Polymer membrane (e.g., Nafion) 13 , 227 | Yttria stabilised Zirconia (YSZ) 34 , 228 |
Cathode material | Ni, Ni–Mo alloys 11 , 227 | Pt, Pt–Pd 13 | Ni/YSZ 34 , 228 |
Anode material | Ni, Ni–Co alloys 11 , 227 | RuO2, IrO2 13 | LSM/YSZ 34 , 228 |
Current density (A/cm2) | 0.2–0.4 13 | 0.6–2.0 13 | 0.3–2.0 34 |
Cell voltage (V) | 1.8–2.4 13 | 1.8–2.2 13 | 0.7–1.5 34 |
Voltage efficiency (%HHV) | 62–82 13 | 67–82 13 | <110 227 |
Cell area (m2) | <4 227 | <0.3 227 | <0.01 227 |
Operating temp. (°C) | 60–80 13 | 50–80 13 | 650–1,000 34 , 228 |
Operating pressure (bar) | <30 227 | <200 227 | <25 227 |
Production rate (m3H2/h) | <760 227 | <40 227 | <40 227 |
Stack energy (kWhel/m3H2) | 4.2–5.9 13 | 4.2–5.5 13 | >3.2 227 |
System energy (kWhel/m3H2) | 4.5–6.6 36 | 4.2–6.6 36 | >3.7 (>47)kWh_energy |
Gas purity (%) | > 99.5 11 | 99.99 227 | 99.9 |
Lower dynamic range (%) | 10–40 13 , 227 | 0–10 13 | >30 |
System response | Seconds 227 | Milliseconds 227 | Seconds |
Cold-start time (min) | <60 36 | <20 36 | <60 |
Stack lifetime (h) | 60,000–90,000 36 | 20,000–60,000 36 | <10,000 |
Maturity | Mature | Commercial | Demonstration |
Capital cost (€ kWel−1) | 1,000–1,200 36 | 1,860–2,320 36 | >2,000 36 |
AEC, the long-standing water electrolysis technology adopted since 1920, is extensively employed in large-scale industrial environments. 31 These systems are known for their accessibility, durability, and relatively lower capital costs, attributed to the absence of noble metals and the maturity of stack components. 5 , 11 , 32 Nonetheless, challenges like low current density, operating pressure limitations, and constraints on dynamic operation adversely impact system scalability and hydrogen production costs. Ongoing endeavors aim to improve current density, operating pressure, and system design to enable dynamic operation, thereby fostering compatibility with intermittent renewable sources. 11 , 32
PEMEC systems, originating from the solid polymer electrolyte (SPE) concept in the 1960s, were introduced to mitigate limitations associated with AEC. 33 Despite being less established than AEC, PEMEC boasts advantages such as high power density, efficiency, and operational flexibility, rendering it suitable for small-scale applications. 11 , 32 However, challenges persist, including the expense of platinum catalysts and fluorinated membrane materials, complexity arising from high-pressure operation, water purity prerequisites, and shorter lifetimes compared to AEC. Ongoing advancements focus on simplifying systems, cost reduction, and enhancing stack manufacturing processes. 11 , 32
SOEC, considered the least mature technology, has had limited commercial adoption, mainly showcased in laboratory-scale trials. 31 Utilizing solid ion-conducting ceramics as electrolytes, it enables operation at elevated temperatures. Noteworthy advantages encompass high electrical efficiency, reduced material expenses, and the capacity for reverse operation or syngas production. Nevertheless, material degradation at elevated temperatures presents a significant obstacle. Present research concentrates on stabilizing existing materials, innovating new ones, and reducing operating temperatures to facilitate commercial viability. 34
The capital costs for AEC and PEMEC systems are estimated at approximately 1000 V/kWel and 2000 V/kWel, respectively, while SOEC systems, being less prevalent commercially, surpass 2000 V/kWel. 31 AEC systems are generally cited as being two to three times more costly than steam methane reformers (SMR), the dominant hydrogen production technology relying on methane gas reformation. 35
To effectively utilize intermittent power sources, Electrolyzers necessitate prompt responses from system components, operation within a narrower dynamic range without compromising gas purity, and efficient cold-start capabilities or standby operation. 31 Although PEMEC Electrolyzers appear most adept at meeting these criteria, AEC and SOEC can also be modified, with their components tailored for intermittent power provision. 36 , 37
3 Performance indicators
Various performance and metrics indicators are widely utilized to assess the electrocatalytic performance of HER and OER electrocatalysts. These include the reversible hydrogen electrode (RHE) potential, mass and specific activities, Tafel slope, overpotential, catalytic stability, and faradaic and energy efficiencies. These metrics are pivotal in designing efficient electrocatalysts and facilitating comparisons across different materials, offering valuable insights into intrinsic activity and redox reaction kinetics.
3.1 Reversible hydrogen electrode potential
To standardize the evaluation and comparison of electrocatalytic performance across different pH conditions, it’s recommended to reference all potentials for HER and OER to the Reversible Hydrogen Electrode (RHE). 4 This reference is expressed by equation (4):
ESHE represents the measured potential of the designated reference electrode with respect to SHE, pH is the actual electrolyte pH value, R is the universal ideal gas constant, T is the absolute temperature, and F is the Faraday constant. At 298 K, the term 2.303 R T/F approximately equals 0.059 V. 4 By incorporating pH dependence, the RHE ensures that the equilibrium potentials of the O2 and H2 redox reactions are fixed at 1.23 and 0 V versus RHE at 298 K, respectively. Standardizing potential values with respect to RHE facilitates direct comparison of electrocatalyst performance across electrolytes with changeable pH. This approach enables researchers to accurately assess and compare electrocatalytic efficiencies under different environmental conditions.
Two prevalent methods are employed to convert the measured potentials of working electrodes referenced to a particular reference electrode to potentials referenced to the RHE. 4 The first method involves determining the actual pH of the electrolyte and using the formula: 38
where, Emeasured represents the measured potential of the working electrode with respect to the reference electrode, ERef0 (vs. SHE) is the standard thermodynamic potential of the reference electrode.
The alternative recommended method involves experimental calibration of the reference electrode against the RHE in the actual electrolyte during the hydrogen redox reaction. 39 This calibration process entails using two platinum electrodes as both the working and counter electrodes alongside the reference electrode in the same electrolyte conditions as those used for electrocatalytic characterization. In the protocol, the electrolyte could be pre-saturated with high-purity H2 before calibration. Cyclic voltammetry is then conducted at a low scan rate in the reversible hydrogen evolution and hydrogen oxidation potential range, and the average value of the two voltage intercepts for two branch curves at zero current yields the experimentally determined conversion factor (Eoffset). Subsequently, the Emeasured (vs. reference electrode) used in the test can be converted to RHE potential using equation (4):
Reference 38] have detailed these procedures and established the calibration curves for various reference electrodes (Ag/AgCl, Hg/HgO and Hg/Hg2Cl2) in different electrolytes. 38
3.2 Specific and mass activities
The ration of current density (j) and electrocatalyst area, at a given potential, is known as the specific activity of an electrocatalyst. Typically, three types of areas are utilized to calculate (j): the geometrical electrode area (GEA), the electrochemically active surface area (ECSA), and the BET specific surface area. 40 ECSA is commonly measured through double layer capacitance (Cdl) in a non-faradaic potential window, preferably in aprotic electrolytes. 41 For electrodes with flat and smooth surfaces, current density can be normalized with GEA. However, for porous electrodes and powder electrocatalysts, it’s advisable to assess specific activity by dividing (j) with BET surface area or ECSA. 4
On the other hand, the mass activity (in A/mgcatalyst) represents the current density normalized by the mass of the active electrocatalyst at a specific potential. An ideal electrocatalyst exhibits high specific activity and mass activity. 42 , 43 In practical applications, achieving optimal mass loading involves maximizing coverage of the conductive substrate with the electrocatalyst and optimizing the thickness of the electrocatalyst layer to enhance electron and mass transfer processes. 4
3.3 Overpotential
Overpotential (η) is the additional voltage needed to drive an electrochemical reaction beyond its thermodynamic potential. It represents the surplus voltage required to overcome energy barriers and resistances within the electrolytic cell, enabling the desired reactions. These barriers include electrical resistance in wiring and electrodes, activation energies of reactions (for the OER and HER), bubble overpotential (where gas bubbles hinder reaction access), and ionic conductivity overpotential. 4
Determining η typically involves electrochemical measurements, often using techniques like linear sweep voltammetry at low sweep rates to minimize capacitance current. 14 Under standard conditions, the reversible thermodynamic potentials for the HER and OER are 0 V and 1.23 V, respectively. Overpotential at an anticipated current density without iR recompence is given by: ηHER = ERHE − 0 V and ηOER = ERHE − 1.23 V. At j = 10 mA/cm2, overpotential (η10) is indeed a crucial parameter for assessing electrodes’ electrocatalytic activity. 44
The choice of electrode material significantly influences overpotential. Electrodes themselves can act as catalysts, reducing the activation energy of electrochemical reactions. 8 Hence, electrode design often involves doping or coating with more stable and active layers. The primary effect of an electrocatalyst is to reduce the overpotential of either the HER or OER. 11 Table 2 listing various electrocatalysts helpful for reducing overpotential or stabilizing electrodes in industrial water electrolysis further illustrate this point.
Hydrogen and oxygen overpotentials of different electrode materials. 11
Electrode composition | Temp. (°C) | Electrolyte | C (M) | j (A/m2) | η (mV) | Ref |
---|---|---|---|---|---|---|
HER | ||||||
Ni–Fe–Mo–Zn | 80 | KOH | 6 | 1,350 | 83 | 229 |
Ni–S–Co | 80 | NaOH | 28 wt% | 1,500 | 70 | 230 |
Ni50 %-Zn | N/A | NaOH | 6.25 | 1,000 | 168 | 231 |
MnNi3·6Co0·75Mn0·4Al0.27 | 70 | NaOH | 30 wt% | 1,000 | 39 | 232 |
Ti2Ni | 70 | NaOH | 30 wt% | 1,000 | 16 | 233 |
Ni50 %Al | 25 | NaOH | 1 | 1,000 | 114 | 234 |
Ni75 %Mo25 % | 80 | KOH | 6 | 3,000 | 185 | 235 |
Ni80 %Fe18 % | 80 | KOH | 6 | 3,000 | 270 | 235 |
Ni73 %W25 % | 80 | KOH | 6 | 3,000 | 280 | 235 |
Ni60 %Zn40 % | 80 | KOH | 6 | 3,000 | 225 | 235 |
Ni90 %Cr10 % | 80 | KOH | 6 | 3,000 | 445 | 235 |
OER | ||||||
Ni + spinel type Co3O4 | 25 | KOH | 1 | 1,000 | 235 ± 7 | 236 |
Ni + La doped Co3O4 | 25 | KOH | 1 | 1,000 | 224 ± 8 | 236 |
MnOx modified Au | 25 | KOH | 0.5 | 100 | 300 | 237 |
Li10 % doped Co3O4 | RT a | KOH | 1 | 10 | 550 | 238 |
Ni | 90 | KOH | 50 wt% | 1,000 | 300 | 239 |
La0·5Sr0·5CoO3 | 90 | KOH | 50 wt% | 1,000 | 250 | 239 |
Ni0·2Co0·8LaO3 | 90 | KOH | 50 wt% | 1,000 | 270 | 239 |
-
a, room temperature.
3.4 Tafel slope
The Tafel equation (Eq. (12)) represents the connection between the overpotential (at each electrode) and the current density’s logarithm: 11
where a = (2.3 R T)/(αF)log(j0) and b = (−2.3 R T)/(αF), T is the absolute temperature (in K), α is the symmetry factor variable, and F is the Faraday constant. “b” is referred to as the ‘Tafel slope’, it helps to define the rate determining step by examining the sensitivity of the current response to the applied potential. 4 A lower Tafel slope indicates greater electrocatalytic kinetics. Some electrocatalysts exhibit Tafel slope values that do not conform to linear fitting of the plot of log(i) versus overpotential (η) across a wide potential range. 45 , 46 , 47 This leads to overpotential-dependent Tafel slopes. 4 , 48
Table 3 presents a comparison of kinetic parameters, including current density and Tafel slope, for HER and OER on various metal electrode materials. If hydrogen adsorption on the cathode is the rate-determining step, electrode materials with more boundaries and cavities in their surface structure will create more electrolysis centers for hydrogen adsorption. 11 Conversely, if hydrogen desorption is the rate-determining step, physical properties such as surface perforation or roughness could increase electron transfer by providing extra reaction area or preventing bubbles growing, thereby enhancing the rate of electrolysis. 11 It is important to note that both the Tafel slope and η10 are significantly influenced by the loading mass of the electrocatalyst, as demonstrated in the comprehensive study by Anantharaj Kundu. 49 This research confirmed that increasing the mass of the electrocatalyst leads to a notable decrease in both the Tafel slope and η10. In the study, five NiO/CC electrodes with incrementally increasing catalyst loadings of 0.205, 0.310, 0.410, 0.615, and 0.820 mg cm⁻2 were prepared using the conventional drop-casting method. These electrodes were then evaluated for OER performance in 1 M KOH, where the dependency of overpotentials and Tafel slope on catalyst loading was investigated. The results clearly showed that both η10 and the Tafel slope decreased as the catalyst loading increased.
Tafel slope of HER and OER on different electrode metals. 11
Metal | Electrolyte | Temp. (°C) | I0 (A/m2) | Tafel slope (mV) | Ref. |
---|---|---|---|---|---|
HER | |||||
Ni | 1 M NaOH | 20 | 1.1 × 10−2 | 121 | 240 |
Fe | 2 M NaOH | 20 | 9.1 × 10−2 | 133 | 241 |
Pb | 6 M NaOH | 25 | 4.0 × 10−2 | 121 | 242 |
Zn | 6 M NaOH | 25 | 8.5 × 10−6 | 124 | 242 |
Co | 0.5 M NaOH | 25 | 4.0 × 10−3 | 118 | 243 |
Pt | 0.1 M NaOH | 22 | 4.0 | 105 | 244 |
Au | 0.1 M NaOH | 25 | 4.0 × 10−2 | 120 | 244 |
OER | |||||
Pt | 30 % KOH | 80 | 1.2 × 10−5 | 46 | 245 |
Ir | 1 M NaOH | N/A | 1.0 × 10−7 | 40 | 246 |
Rh | 1 M NaOH | N/A | 6.0 × 10−8 | 42 | 246 |
Ni | 50 % KOH | 90 | 4.2 × 10−2 | 95 | 247 |
Co | 30 % KOH | 80 | 3.3 × 10−2 | 126 | 245 |
Fe | 30 % KOH | 80 | 1.7 × 10−1 | 191 | 245 |
3.5 Stability and durability
The stability of direct water splitting involves efficiently producing hydrogen and oxygen through electrochemical processes over time. Various factors, like the corrosive nature of electrolytes, influence stability in electrochemical cells. Electrode materials must exhibit high corrosion resistance to ensure prolonged stability, while membrane stability is crucial for effectively separating HER and OER. 44 Catalyst stability is equally important for sustained high performance, and maintaining selective ion transport across membranes is essential to prevent unintended crossovers and instability. Controlling operating conditions is imperative to mitigate adverse stability effects. Large-scale systems often face scaling issues, impacting efficiency and stability. Regular monitoring, maintenance, and ongoing research in material science, catalyst, and system engineering enhance long-term stability.
Meanwhile, evaluating the stability of electrocatalysts and two-electrode electrolyzers involves techniques like cyclic voltammetry or chronopotentiometry. Durability tests typically last from tens to thousands of hours, conducted at a current density of 10 mA/cm2. In 4 , 48 industrially relevant hydrogen production, electrolyzers must operate at higher current densities, usually ranging from 0.5 to 2 A/cm2, while demonstrating extreme stability and a lifespan of thousands of hours to years. 4 Monitoring dissolved metal cation concentrations and other ions in electrolytes, along with quantifying the percentage of weight loss of active metal centers after long-term stability tests, is advisable.
3.6 Faradaic and energy efficiencies
The Faradaic efficiency (FE) measures the actual amount of produced H2 or O2 compared to the theoretically calculated amount, assuming all the passed charge contributes solely to forming H2 or O2. The FE is calculated using the formula: 4
where I is the current (A), n = 2 for H2 and 4 for O2, is the number of electrons transferred to produce 1 mol gas, and F = 96,485.3 C/mol is Faraday’s constant. Water-gas displacement method or Gas chromatography (GC) are typically used to measure the amount of H2 or O2 produced, while an electrochemical workstation monitors the passed charge. Ideally, an electrocatalyst should exhibit 100 % faradaic efficiency, while losses occur in water electrolyzers when electrons and ions contribute in undesirable reactions at the electrodes and/or in the electrolytes systems.
Energy efficiency (also called electrical efficiency or electricity-to-hydrogen energy conversion efficiency) is another crucial factor for assessing the effectiveness of water electrolysis cells or systems. Cell-level energy efficiency is commonly evaluated by considering only the electricity energy input for the electrolysis cell, excluding other energy inputs such as heat sources and auxiliaries. 4 The energy efficiency (EE) is calculated using the formula: 4
where nH2,measuredn is in mole/s, Pdc is the power supplied by the external direct-current power supply (W), U is the voltage applied (V), FE is the faradaic efficiency, ΔHH2 is the reaction enthalpy (heat of reaction), and ETN is the thermoneutral voltage derived from ΔHH2 (ETN = ΔH0/nF, where ΔH0 = 285.8 kJ/mol under standard conditions 4 ). Numerous factors need to be considered when calculating EE. A detailed examination of these factors can be found in the review authored by Li et al. 4
4 Emerging strategies for electrochemical water splitting
To overcome a range of technical challenges in water electrolysis, such as overcoming energy barriers, integrating incompatible electrocatalysts, and managing reactant mass transfer, several strategies have been invented. These include decoupled, hybrid, and tandem water electrolysis, as well as biological electrolysis cell approaches. Additionally, seawater electrolysis stands as a separate technique due to its distinct characteristics and potential. In the next sections, we’ll explore each of these techniques, detailing their principles and accomplishments.
4.1 Decoupled water electrolysis
4.1.1 The concept
The decoupled water electrolysis concept was proposed to overcome several disadvantages of classical cell electrolyzers, where the HER and OER happened simultaneously (Figure 3a), possibly lead to H2/O2 crossover, even if an ostensibly gas impermeable membrane (or diaphragm). 21 Thus, scenario is particularly high probable at low current densities or under high gas pressure electrolysis. 50 This problem could touch the H2 purity, in addition to the risk of creating of an extremely explosive or oxidative environment due to the coexistence H2, O2 and electrocatalysts under favorable electrocatalytic conditions, which pose safety concerns, degrade the electrolyzer and shorten its operation lifespan. 4 , 51 , 52 These issues call for alternative electrolyzer designs, not only avoiding the crossover of H2/O2 gases but also allowing suppleness in controlling electrolysis products.

Comparison of traditional ‘coupled’ electrolysis and decoupled electrolysis for water splitting: (a) Conventional water electrolysis in acidic conditions. (b) Decoupled oxygen generation and mediator reduction. (c) Decoupled mediator reoxidation and hydrogen evolution. In panels (b) and (c), the circle denotes the oxidized form of the decoupling agent “medaitor”, while the diamond represents the reduced form of this decoupling agent. From. 54
Figure 3b and c presents schematic illustrations of a decoupled water electrolyzer. In this system, during the HER at the cathode, a redox mediator (dissolved in the electrolyte) undergoes oxidation at the anode instead of the OER. This oxidized redox mediator is then reduced back to its initial state at the same electrode, while the OER is coupled to another working electrode. Through the alternation of these steps, hydrogen and oxygen are generated sequentially. This method allows for the temporal and/or spatial separation of HER and OER, facilitated by redox mediators. 21 Essentially, the redox mediator serves as a reversible electron and/or proton donor/acceptor. 21 When the redox mediator can simultaneously accept or donate both protons and electrons, it can act as a pH buffer during electrolysis and is referred to as an electron coupled-proton buffer, ECPB. 53
Mediators can be categorized into three main types: soluble, dispersed, and solid-state decoupling agents. 54 An example of a dissolved mediator is the polyoxometalate Li6[P2W18O62], which has the remarkable capacity to reversibly accept up to 18 electrons in water media. Transition metal salts, such as those involving the V(III)/V(II) and Ce(III)/Ce(IV) redox couples, 55 as well as complexes like Fe(CN)63−/Fe(CN)64−, 56 are also commonly used in this category. Additionally, organic systems like hydroquinone sulfonate and anthraquinone-2,7-disulfonic acid serve as dissolved mediators. 57 , 58 Dispersed mediators, on the other hand, involve particulate systems dispersed in a solution containing a redox couple. For instance, Zhao et al. 59 utilized BiVO4 particles dispersed in an aqueous medium comprising the Fe(III)/Fe(II) redox pair. Upon solar light radiation, BiVO4 particles oxidize water to generate O2 while reducing Fe(III) to Fe(II) at the same time. Solid-state mediators, exemplified by NiOOH/Ni(OH)2, 60 are integrated with electrodes and operate without the need for solubilization. These solid mediators facilitate electron transfer reactions directly at the electrode surface. The choice of mediator phase significantly influences the physical components and operation of the system. Indeed, soluble redox mediator systems can operate under static or flow conditions, as the mediator can be pumped between chambers. Conversely, in systems with solid-state mediators, the decoupling agent is typically integrated with an electrode that can be moved between cells to interact with either the HER or OER. Detailed analysis of key studies in this area will be discussed further.
The success of redox mediator-assisted decoupled water electrolysis hinges largely on the electrochemical properties of the redox mediator. Several crucial criteria must be considered when selecting suitable candidates for redox mediators: 61
High solubility in aqueous media (of soluble mediators),
Rapid and reversible electron transfer between the redox mediator and the inert electrode, ensuring that the pH of the reaction medium remains unchanged throughout both the OER and HER steps.
Well-positioned redox potential between the onset HER and OER potentials,
Excellent durability over a large number of redox cycles,
Economical composition and synthesis.
4.1.2 Decoupling water splitting systems
Several innovative strategies for nonconventional water splitting have been summarized by You and Sun, 18 wherein H2 production from water splitting is integrated with other oxidation reactions of greater utility than O2 evolution. Additionally, Liu et al. 61 Paul and Symes, 54 Wallace and Symes 62 have provided valuable mini-reviews on decoupled water splitting agents, highlighting recent progress in the field. More recently, McHugh et al. 21 reported a comprehensive review article focusing on redox mediators for water splitting. Table 4 provides an overview of various decoupling agents along with their key performance metrics, demonstrating their effectiveness in electrochemical water splitting. Readers interested in more detailed information are encouraged to refer to the provided references for further insights.
Decoupling agent | Redox potential (V vs SHE) |
pH for redox potential | Maximum operation | Best Faradic efficiency | Ref. | |
---|---|---|---|---|---|---|
HER | OER | |||||
|
||||||
H3PMo12O40 | +0.50, +0.65 | 0.3 | 5 cycles | 100 % | 100 % | 53 , 248 , 249 |
H4[SiW12O40] | 0.0, −0.22 | 0.5 | 9 cycles | 75 ± 7 % | 100 % | 64 , 250 , 251 |
H3PW12O40 | +0.237, −0.036 | 0.42 | – | 46.6 % | – | 63 |
H4SiO4·12MoO3 | +0.509 | 0.69 | – | – | – | 63 |
H6ZnW12O40 | −0.078, –0.198 | 0.4 | 200 cycles | 95.5 % | – | 252 |
[P2W18O62]6– | + 0.3, + 0.1, 0 to −0.5 | 1 | 100 cycles | – | – | 253 |
V(III)/V(II) | −0.26 | 0 | – | 96 ± 4 % | – | 55 , 254 |
Ce(IV)/Ce(III) | + 1.48 | 0 | – | – | 78.8 % | 254 |
Fe(CN)63−/Fe(CN)64– | + 0.77 | 7.2 | – | 99.88 % | – | 56 , 255 |
FcNCl | + 0.6 | 6.5 | 20 cycles | 100 % | 100 % | 50 |
Hydroquinone sulfonate | + 0.65 | 0.7 | 20 cycles | 98 ± 7 % | 91 ± 5 % | 57 |
AQDS | + 0.214 | 0 | 100 cycles | 100 % | 100 % | 58 |
NiOOH/Ni(OH)2 | + 0.55 | 14 | 100 cycles | 100 % | 100 % | 60 , 65 , 66 |
FeOx | + 0.8 to + 1.5 | 14 | 50 cycles | >90 % | >90 % | 256 |
MnO2/MnOOH | + 0.15 | 14 | 10 cycles | 100 % | – | 257 |
Fe3O4/FeOOH | – | – | – | 100 % | – | 258 |
CoO | – | – | – | 100 % | – | 259 |
PTPAn | + 0.789 + 0.689 | 0.3 | 120 cycles | 98.7 % | 98.7 % | 67 |
PTO | + 0.46 + 0.589 | 0.3 | 300 cycles | – | – | 260 |
PANI | + 0.45 + 0.91 | 0.3 | 40 cycles | – | – | 261 |
-
j, current density; FcNCl, (ferrocenylmethyl)trimethylammonium chloride; AQDS, anthroquinone-2,7-disulfonic acid; PTPAn, polytriphenylamine; PTO, pyrene-4,5,9,10-tetraone; PANI, polyaniline.
Symes and Cronin achieved the initial practical implementation of decoupled oxygen and hydrogen evolution through electrolytic water splitting in 2013. 53 They employed the polyoxometalate phosphomolybdic acid (H3PMo12O40) to separate the OER from the HER under acidic conditions, as described by the following equations:
Therefore, during the oxidation of water to produce O2, protons, and electrons, the polyoxometalate underwent reduction, forming H5PMo12O40. Upon subsequent reoxidation, H5PMo12O40 was converted back to H3PMo12O40, releasing protons and electrons which were utilized to generate hydrogen at the cathode. Subsequently, the same researchers established a range of mediators (organic redox), such as 1,4-hydroquinone derivatives and anthraquinone-2,7-disulfonic acid. 57 , 58 Additionally, Symes’s and Cronin’s groups investigated other inorganic polyoxometalate complexes, including phosphotungstic acid (PTA) and silicotungstic acid (STA), as alternative redox mediators with more negative redox potentials than the HER onset potentials of the catalysts. 63 , 64 Unlike previous redox mediator systems, which required two electrochemically driven steps, PTA and STA mediators are initially reduced (electrochemically) and accept protons during the OER. In the subsequent non-electrochemical step, the reduced PTA or STA evolve hydrogen spontaneously when making contact with Pt, Ni2P, Ni5P4, or Mo2C catalysts without additional energy input, regenerating the oxidized PTA or STA, as the HER onset potentials (on those catalysts) are more positive than the redox potentials of the reduced from of STA and PTA. 62 , 63 This approach also accomplished decoupled HER and OER.
After Symes’s and Cronin’s works, various types of decoupled systems/setups were proposed. 21 In 2014, Amstutz et al. 55 introduced a novel vanadium- and cerium-based anolyte system for decoupled electrolysis. The adopted device configuration (Figure 4a) allowed for both conventional operation as a redox flow battery and the conversion of energy to hydrogen, offering large flexibility for renewable energy usage. They utilized a V(III) salt in sulfuric acid, in which V(III) is reduced to V(II) on a graphite felt electrode:

Schematic representation of a sample of experimental setups utilized by researchers for decoupled water electrolysis. From. 54
The position of the V(III)/V(II) redox couple (−0.26 V vs SHE) is more negative than the proton reduction couple, allowing for spontaneous hydrogen generation when the V(II) solution is exposed to hydrogen evolution catalysts. The HER was achieved over an Mo2C catalyst with a Faradaic yield of up to 96 %. At the anode, the oxidation of an acidic solution of Ce(III) to Ce(IV) took place:
The position of Ce(III)/Ce(IV) redox under these circumstances is more anodic than that for water oxidation. This means that Ce(IV) can oxidize water to generate oxygen, becoming reduced back to Ce(III) in the process. This catalytic water oxidation reaction was achieved over RuO2 nanoparticles, although the Faradaic yield for the OER was somewhat less than optimal (78 % ± 8 %) due to a side reaction involving degradation of the RuO2 catalyst by Ce(IV). The authors suggested transitioning to an all-vanadium system to mitigate this issue, where the reduction of V(V) to V(IV) would mediate the OER.
Ho et al. 55 introduced a similar combined chemical/electrochemical system for hydrogen production (Figure 4b). A photoelectrochemical cell was employed to reduce V(III) to V(II) while performing OER at the anode. The V(II) solution was directed to a flask containing a Mo2C catalyst for on-demand H2 evolution. A bipolar membrane maintained a significant pH differential between the anode and cathode compartments, crucial for the system’s operation. The dual electrolyte system comprised an anodic compartment with a Ni mesh oxygen evolution electrode in KOH concentrated solution and a cathodic compartment with a Pt foil electrode in sulfuric acid. At the Ni anode, OH− was oxidized, producing electrons, water, and oxygen. These electrons reduced V(III) ions to V(II), and the charged V(II) solution was transferred to a detached reactor for hydrogen production. The study investigated hydrogen production efficiency at various depths of charge of the vanadium solution, showing Faradaic efficiencies ranging from 83 % to 87 %. Additionally, the system demonstrated hydrogen generation under elevated pressures, with efficiencies approaching 60 % at pressures of up to ten atmospheres. Furthermore, the authors showcased the system’s direct powering by renewable energy through coupling it to a photovoltaic module. Hydrogen was generated with a Faradaic efficiency of 80 %, under solar irradiation, achieving a solar-to-hydrogen energy conversion efficiency of 5.80 %.
Goodwin and Walsh 56 presented a novel method to spatially decouple the OER/HER using the Fe(CN)63−/Fe(CN)64− redox couple and a closed bipolar electrode. In their setup (Figure 4b), they utilized carbon cloths connected by a wire as the closed bipolar electrode, bridging between two separate two-compartment cells. Each cell contained a solution of K3Fe(CN)6/K4Fe(CN)6 in 0.1 m KOH, with one end of the closed bipolar electrode submerged in each solution. The device operated by facilitating water oxidation at a Pt electrode coupled with the reduction of Fe(CN)63− at one end of the closed bipolar electrode in one cell. In the other cell, oxidation of Fe(CN)64− occurred at the opposite end of the closed bipolar electrode, accompanied by hydrogen evolution at the second platinum electrode.
Li et al. 50 investigated the use of ferrocene derivatives, specifically (ferrocenylmethyl)trimethylammonium chloride (FcNCl), as a decoupling agent to separate the HER and OER in a two-compartment H-cell setup (Figure 4d). One compartment contained a carbon counter electrode in a solution of FcNCl, while the other compartment housed electrodes for the HER and OER. The system enabled temporally separated HER and OER to take place, demonstrating stability over 20 redox cycles without significant pH variation. Furthermore, the study highlighted the potential for sunlight-driven hydrogen production, achieving currents of approximately 20 mA/cm2 for the HER using a photovoltaic module under solar irradiation.
Zhao et al. 59 proposed a more direct decoupled solar-to-hydrogen system, where sunlight activates BiVO4 particles dispersed in an aqueous solution containing the Fe(III)/Fe(II) redox couple. Under solar irradiation, BiVO4 particles oxidize water to produce oxygen while simultaneously reducing Fe(III) to Fe(II). The Fe(II) ions can then be re-oxidized to Fe(III) in an electrolytic cell, generating hydrogen at the cathode with 100 % Faradaic efficiency. Notably, this process operates at cell potentials significantly lower than conventional methods. Although conceptually similar to previous solar-to-hydrogen systems, Zhao et al.’s work achieved remarkable solar-to-hydrogen efficiencies of around 2 %, a substantial improvement over earlier studies.
Chen et al. 60 proposed a novel approach to decoupled systems where the electrode served as the redox mediator. In their single-compartment setup, a nickel hydroxide (Ni(OH)₂) electrode underwent oxidation at the cathode, facilitating the HER. The electrode could then be regenerated through reduction coordinated with the OER at the anode. Similarly, Landman et al. 65 , 66 proposed a similar system but implemented it in two separate chambers using photoelectrochemistry. Ma et al. 67 also proposed a comparable system utilizing polytriphenylamine (PTPAn) as the electrode material. They demonstrated that this system achieved near quantitative Faradaic efficiencies for H₂ and O₂ production, with a photovoltaic-driven setup exhibiting an excellent overall solar-to-hydrogen conversion efficiency of 5.4 %. Additionally, Dotan et al. 68 presented a decoupled system involving Ni(OH)₂, where the oxidation of Ni(OH)₂ was coordinated with H₂ production, and the electrode was regenerated through introducing a heated solution to the electrochemical cell.
These examples showcase the potential of decoupled water splitting systems, although they currently lack the capability to generate electricity in one of their decoupled cycles, hindering their function as a dual-use H₂ generation/electricity storage system. An example of such a system described by Bienvenu 69 involved the reduction of Zn2+ ions to Zn0 at the cathode during OER at the anode. Upon polarity switching, the Zn0 electrode underwent spontaneous oxidation, resulting in the generation of both electricity and hydrogen at the cathode.
In conclusion, decoupled water electrolysis systems represent a promising solution to mitigate the challenges associated with traditional electrolyzers, offering efficient hydrogen production while avoiding gas crossover and ensuring safety. These systems, facilitated by various redox mediators and innovative configurations, demonstrate significant potential for advancing clean hydrogen generation technologies. However, further research is needed to address limitations and enhance the scalability and dual-use capabilities of decoupled systems for broader applications in renewable energy and beyond.
4.2 Hybrid water electrolysis
In parallel with the concept of decoupled water electrolysis, researchers have been exploring hybrid water electrolysis as a means to mitigate the limitations of the OER, such as high oxidation overpotential, thereby reducing the overall energy input required for hydrogen generation. This approach involves replacing the OER with alternative, thermodynamically more favorable oxidation reactions using various substrates like alcohols, 70 , 71 amines, 72 aldehydes, 73 hydrazine, 74 , 75 urea, 74 , 76 and more. 77 Fan et al., 78 Kahlstorf et al., 77 Ran et al., 16 Chen et al. 22 and Du et al. 23 have extensively reviewed the progress in hybrid water electrolysis. Notably, this system holds promise for waste treatment or valorization, provided the waste stream meets certain criteria: 79
Firstly, the substance must be soluble for efficient operation.
Secondly, its theoretical oxidation potentials should be lower than the OER to minimize energy consumption.
Thirdly, waste valorization should either generate economic benefits through electrochemical oxidation or, if it is not suitable for valorization, it should be transformed into less harmful byproducts.
Lastly, a constant waste output is crucial for research value and practical availability.
Several wastes, including hydrogen sulfide, cellulose wastes, crude glycerol, and plastics, meet these criteria. 79 Hydrogen sulfide and glycerol dissolve readily, while cellulose and certain plastics can be hydrolyzed to yield water-soluble byproducts such as ethylene glycol and (EG) 5-hydroxymethylfurfural (HMF). Figure 5 illustrates alternative oxidation reactions for waste valorization, along with their oxidation potentials, indicating potential energy savings. Significantly, the oxidation potential of the sulfide reaction lies below the HER potential of 0 V, suggesting significant potential for energy saving 80 and, sometimes, the reaction can proceed without the need for electricity consumption. 81

Schematic illustration of alternative oxidation reactions for hybrid water electrolysis systems with theoretical oxidation potentials axis. From. 79
Due et al. 23 classified waste valorization into three main categories: (i) Reagent-sacrificing type, which enables H2 production under low voltages while sacrificial reactants (such as ammonia and hydrazine) are oxidized to non-valuable products; (ii) Pollutant-degradation type, where the sacrificial reactants undergoing oxidation are pollutants (e.g., organic dyes, urea …); and (iii) value-added type, which involves upgrading organic reactants (e.g., glucaric acid, tetrahydroisoquinolines, biomass-derived chemicals) to valuable products. However, practical outcomes depend on various factors, including reaction products, conditions, and catalysts. Consequently, careful product selection and cost-effective catalyst development are vital for future research in this area. A comprehensive analysis of these factors is provided in the review by Wang et al. 79
4.2.1 Reagent-sacrificing type
Hydrazine and ammonia electrooxidation are typical examples of this category. 23 This system involves the anodic conversion of hydrazine and ammonia, both commercially valuable substrates, into non valuable products like N2 (Eqs. (19) and (20)).
Crucially, the thermodynamic potential of these reactions, i.e., −0.38 V for reaction (19), is markedly lower than that of the OER (1.23 V), resulting in a significant decrease in the overall cell voltage needed for water splitting. 23 In many cases, hydrazine oxidation reaction (HzOR) enabled the attainment of a current density of up to 500 mA/cm2 at a cell voltage of 1 V without generating O2. Additionally 75 , 82 , 83 , 10 mA/cm2 could be achieved at a microvoltage level in the presence of hydrazine. 23 Bifunctional electrocatalysts have been developed to facilitate the HER coupled with hydrazine oxidation, further enhancing the efficiency of hydrogen generation. the addition of 0.02 M hydrazine to a 1 M KOH solution reduced the required voltage for reaching 15 mA/cm2 from 1.97 V to 0.63 V, utilizing the bifunctional Ni@Pd–Ni alloy nanosheets as electrodes. 84 Sun et al. 85 reported that CoS2/TiM nanoarrays and Ni2P/NF catalysts enable efficient hydrogen evolution at relatively low cell voltages, with the former achieving hydrogen evolution with only a cell voltage of 0.81 V to achieve 100 mA/cm2 (pure water splitting required 1.89 V to achieve this current density). Furthermore, Ni2P/NF catalysts enable H2 production with 1.0 V at 500 mA/cm2, exhibiting remarkable durability and highly Faradaic efficiency for H2 generation. 82 Moreover, Tang et al. 86 introduced a porous nitrogen-doped carbon material encapsulating cobalt/LaCoOx hybrid nanoparticles (Co/LaCoOx@N–C) as a bifunctional catalyst for both HER and HzOR. This catalyst demonstrated remarkable electrochemical activity in HzOR, achieving 69.2 mA/cm2 at only 0.3 V. Moreover, the development of direct hydrazine fuel cells (DHFCs) offers a promising avenue for sustainable hydrogen generation. In these systems, hydrazine serves as the feedstock, leveraging its high energy density and utilization rate. 87 , 88
Durable bifunctional electrocatalysts like P,W co-doped Co3N nanowire on NF (PW–Co3N NWA/NF) demonstrate remarkable electrochemical performance in hydrazine splitting, facilitating efficient hydrogen production with minimal energy consumption. 87 PW-Co3N NWA/NF achieves hydrogen evolution with an exceptionally low cell voltage of 28 mV to reach 10 mA/cm2, outperforming overall water splitting. 87 The oxidation of hydrazine can sustain a current density of 200 mA/cm2 over an overpotential of 607 mV, underscoring the enhanced energy efficiency of hydrazine oxidation coupled with HER for H2 production. 87 Furthermore, the integration of direct hydrazine fuel cells with overall hydrazine-splitting units (using Fe-doped CoS2 nanosheets as electrocatalysts for both the HER and HzOR) in a self-powered system demonstrates the feasibility of efficient hydrogen evolution without external power input, achieving an H2 evolution rate of 9.95 mmol/h, a 98 % Faradaic efficiency, and a 20-h stability (electrode:3D nanoporous NixCo1-xSe nanorod array). 83
Similar to hydrazine oxidation, the oxidation of ammonia (Eq. (20)) offers a thermodynamically more favorable alternative to OER. Ammonia, being more ubiquitous and easily accessible compared to hydrazine, holds greater promise as a replacement for OER. 23 While Pt has traditionally been recognized as the most effective catalyst for ammonia oxidation, 23 , 89 its high cost and scarcity have spurred the exploration of non-noble metal-based alternatives. Mei et al. 90 introduced a catalyst consisting of N-doped NiZnCu layered double hydroxides (LDH) with reduced graphene oxide on nickel foam (N-LDH/rGO). In a two-electrode system utilizing the prepared NiZnCu LDH as both anode and cathode, 50 mA/cm2 was achieved at 0.769 V, surpassing the noble-metal system where Pt/C served as the cathode and IrO2 served as the anode by 0.170 V. Furthermore, this two-electrode system maintained the current density for at least 18 h. Further studies reported in the literature highlight the advantageous use of ammonia in advancing HER using various synthesized electrocatalysts for the cathode material, such as Ni(OH)2–Cu2O, N–NiZnCu LDH, Ni50Cu50/C, among others. 91 , 92 , 93
Overall, despite the added expense associated with sacrificial reagents, their utilization ensures stable water electrolysis at high current densities without O2 production, rendering them an optimal choice for hybrid water electrolysis systems.
4.2.2 Pollutant-degrading type
Sacrificial reagents tend to increase the overall system cost compared to the value of the starting substrate (such as hydrazine or ammonia), despite the potential reduction in cell voltage. To address this challenge, these substrates could be substituted with environmental pollutants, offering opportunities for oxidative degradation (remediation) while also advancing HER. When integrated with water electrolysis, this process of anode oxidation, resembling waste treatment, is referred to as the pollutant-degrading type. 23
A typical example is Urea (CO(NH2)2), commonly found in industrial wastewater and domestic sewage, although organic dyes are also viable alternatives. Combining the urea oxidation reaction (UOR) with water electrolysis can achieve ecological preservation and clean energy production by efficiently generating H2 and purifying urea-rich wastewater or urine (Gnana Kumar et al., 2020). The UOR (Eq. (21)) coupled with HER (Eq. (6)) in alkaline electrolysis results in the overall reaction depicted in Eq. (22): 94
Reaction (21) has undergone extensive examination across various cathode and anode materials. The overall cell potential required to attain 10 mA/cm2 (V10) fluctuates depending on the specific anode/cathode pairing and the concentration of urea. This variability is demonstrated below:
V10 = 1.41 V for MnO2/CoPx couple using 1.0 M KOH with 0.5 M CO(NH2)2, 95
V10 = 1.38 V for Ni3Se4/CoPx couple using 1.0 M KOH with 0.1 M CO(NH2)2, 96
V10 = 1.305 V for N–NiZnCu LDH/N–NiZnCu LDH couple using 1.0 M KOH with 0.3 M CO(NH2)2, 90
V10 = 1.4 V for CoP/C/CoP/C couple using 1.0 M KOH with 0.1 M CO(NH2)2, 97
V10 = 0.89 V for N–CoP/N–CoP couple using 0.5 M H2SO4, 0.96 M FeSO4/0.74 M Fe3(SO4)2. 80
In all of these systems, successful degradation of urea has been accomplished. Research into the mechanism and kinetics of UOR on Ni-based electrocatalysts in an alkaline medium indicates that urea oxidation takes place subsequent to the formation of NiOOH under electrooxidation conditions. 22 NiOOH then interacts with urea molecules, replenishing active sites for subsequent adsorption/desorption of intermediates. 22
In the other hand, Zhang et al. 98 explored the combination of HER with the decomposition of methylene blue, a textile dye. They utilized FeNiP nano-sheets coated with a carbon shell as the cathode catalyst. The oxidation was observed at approximately 1.2 V across different concentrations of methylene blue. Due to the excellent protection and conductivity supplied by the carbon shell, the catalyst exhibited an overpotential of 104 mV to achieve a current density of 10 mA/cm2, which remained stable even after 1,000 cycles of CV scans. Additionally, under a constant current density of 1 mA/cm2, entire degradation of methylene blue could be achieved in 53 min, where the corresponding potential remained steady for over 2,500 s.
In conclusion, considering the ongoing escalation of environmental pollution, the coupling of HER with pollutant degradation undeniably holds promising prospects for the future. However, since real-world pollutants are much more complex, further research efforts should be directed towards realizing real hybrid electrolysis under practical scenarios.
4.2.3 Value-added type
Table 5 provides an overview of recent electrooxidation of organic (EOO) reactions considered value-added. Briefly, small alcohols undergo anodic partial oxidation, yielding aldehydes/ketones in a 2e− process or carboxylic acids in a 4e− process, while aldehydes can be oxidized to carboxylic acids. Oxidative coupling of two alcohols enables ester production, exemplified by ethyl acetate from ethanol. Ring-opening reactions, like adipic acid production from cyclohexanone/cyclohexanol, are also feasible. 99 Larger biomass-derived compounds, such as 5-hydroxymethylfurfural (HMF) or glycerol, yield various products based on reaction conditions and catalysts. Glycol oxidation, for instance, yields multiple commercially relevant products, including formic acid, lactic acid, and carbon dioxide. 100 , 101 Alkenes enable the synthesis of epoxides, diols, and higher oxidized species, while amines can be oxidized to nitriles via dehydrogenation or coupled to form diazo compounds. Some detailed analyses oof the results were discussed in the subsequent sections.
Sample of electrooxidation of organics (EOO) producing value-added chemicals within hybrid water electrolysis systems. 262
Reaction type | Anode catalyst | Organic substrate | concentrations | Product | EOER (VRHE) at 10 mA/cm2 | VEOO (VRHE) at 10 mA/cm2 | Ref. |
---|---|---|---|---|---|---|---|
Co(OH)2@HOS | Methanol | 3.0 M in 1.0 M KOH | Formate | 1.571 | 1.385 | 263 | |
Co3O4 NSs | Ethanol | 1.0 M in 1.0 M KOH | Ethyl acetate | 1.50 | 1.445 | 264 | |
Rh nanosheets | 2-Propanol | 1.0 M in 1.0 M KOH | Acetone | 1.61 | 0.233 | 265 | |
NC@CuCo2N x | Benzyl alcohol | 15 mM in 1.0 M KOH | Benzaldehyde | 1.46 | 1.25 | 266 | |
Mo–Ni alloy NP | Benzyl alcohol | 10 mM in 1.0 M KOH | Benzoic acid | 1.49 | 1.345 | 267 | |
PdAg | Ethylene glycol | 1 M in 0.5 M KOH | Glycolic acid | 1.55 | 0.57 | 268 | |
MoO x /Pt | Glycerol | 0.1 M in 1.0 M KOH | Glycerate | – | – | 269 | |
Co–Ni alloy | Glucose | 0.1 M in 1.0 M KOH | Gluconic acid/gluconolactone | – | – | 270 | |
Ni3S2 | Furfuryl alcohol | 10 mM in 1.0 M KOH | 2-Furoic acid | ca. 1.50 (onset) | ca. 1.35 (onset) | 271 | |
Aldehyde oxidation | Ni2P/Ni | FUR | 30 mM in 1.0 M KOH) | 2-Furoic acid | 1.55 (onset) | 1.34 (onset | 272 |
PbO2 | FUR | 10 mM in pH 1.0 H2SO4 | Maleic acid | 1.85 (onset) | 1.60 (onset) | 273 | |
MoO2–FeP@C | HMF | 10 mM in 1.0 M KOH | FDCA | 1.474 | 1.359 | 274 | |
(FeCrCoNiCu)3O4NSs | HMF | 50 mM in 1.0 M KOH | FDCA | 1.50 (onset) | 1.35 (onset) | 275 | |
Amine oxidation | NiSe | Benzylamine | 1 mmol in 1.0 M KOH | Benzyl nitrile | 1.48 (onset) | 1.34 (onset) | 72 |
VP-Ni(OH)2 NSs | Propylamine | 10 mmol 1.0 M KOH | Propionitrile | 1.53 | 1.36 | 276 | |
Nitroalkane oxidation | NiSe nanorods | α-nitrotoluene | 0.4 mmol in 1.0 M KOH | E-nitroethene | – | – | 277 |
S-DHD of THIQ | Ni2P 1,2,3,4- | Tetrahydroisoquinoline | 0.5 mmol in 1.0 M KOH | 3,4-Dihydroisoquinoline | ca. 1.48 (onset) | ca. 1.10 (onset) | 278 |
Sulfide oxidation | CoFe-LDH | Diphenyl sulfide | 0.25 M in MeCN/H2O | Diphenyl sulfoxide | 1.90 V | 1.39 V | 279 |
-
S-DHD of THIQ, semi-dehydrogenation of Dihydroisoquinolines; FUR, furfural; HMF, 5-hydroxymethylfurfural; FDCA, 2,5-Furandicarboxylic acid.
Hydrogen sulfide (H2S), a notorious environmental contaminant, can be recovered as elemental sulfur, which has diverse applications like pesticide manufacturing and lithium-sulfur batteries. 79 Integrating the sulfide oxidation reaction (SOR) with the HER in hybrid water electrolysis systems offers a more favorable theoretical voltage (Figure 5) and faster reaction kinetics than traditional water electrolysis, leading to reduced energy consumption. 102
Besides elemental sulfur, SOR can yield sulfites, thiosulfates, and sulfates. However, sulfur’s insolubility water leads to deposition on the anode, forming a high-resistivity passivating layer (∼1015 Ω/m). 79 While highly alkaline electrolytes may mitigate this, a more effective approach involves preventing sulfur deposition through interface engineering. Zhang et al. 103 introduced a “self-cleaning” concept, developing a hydrophobic NiS2 catalyst that resists sulfur species deposition, enabling prolonged desulfurization (100 h at 20 mA/cm2). Zhang et al. 104 proposed “chainmail catalysts”, constructing a graphene-encapsulated CoNi alloy catalyst (CoNi@NGs) preventing direct metal-sulfur contact. This catalyst catalyzed sulfide oxidation via graphene shell-induced electron transfer, exhibiting high activity over 500 h (30 mA/cm2). Various electrocatalysts were also developed by various researchers, showing variable stabilities (e.g., Ni–Co–C: 64 h at 100 mA/cm2, 102 CoS2@C/MXene: 180 h at 30 mA/cm2, 105 WS2: 14 h at 100 mA/cm2 106 ).
Crude glycerol, a by-product constituting 10 wt% of the biodiesel production process, poses a significant oversupply issue, with global annual production surpassing 5 billion liters. 79 Anodic oxidation within hybrid water electrolysis systems not only converts glycerol into valued chemicals but also concurrently generates hydrogen at the cathode, offering notable carbon-neutral advantages. 107 Glycerol’s susceptibility to nucleophilic attack, owing to its three adjacent hydroxyl groups, results in a spectrum of oxidation products ranging from C1 to C3, including formic acid, lactic acid, glycolic acid, and dihydroxyacetone. Oxidizing glycerol to formic acid is favored due to its potential to minimize byproduct formation. The theoretical oxidation potential for glycerol to formic acid (Eq. 24, 0.69 V) is significantly lower than that of the OER (Figure 5). However, efficient catalysts are crucial for this conversion, with nickel compounds garnering considerable attention. 79
Li et al. 108 introduced a carbon fiber cloth-supported NiMoN nanoplate catalyst, requiring only 1.36 V voltage in a hybrid water electrolysis system at 10 mA/cm2 and achieving a remarkable Faradaic efficiency of 95 % for formic acid production. Zhu et al. 109 proposed a Ni3N/Co3N heterogeneous nanowire catalyst, achieving 1 A/cm2 at a voltage of 2.01 V and realizing an impressive Faradaic efficiency of 94.6 % toward formate production. Tran et al. 110 explored the performance of glycerol oxidation under various crystalline MnO2 catalysts, i.e., α-, β-, and γ-MnO2. The three phases showed selectivity exceeding 90 % for the C3 product, yet γ-MnO2 demonstrated superior stability and current density compared to the other phases.
The cellulosic waste conversion into individual components, such as glucose, furfural, oligosaccharides, and 5-hydroxymethylfurfural (HMF), through hydrolysis processes holds promise for waste valorization. 111 HMF serves as a crucial platform chemical facilitating the subsequent conversion of cellulose waste into several valuable products, including 2,5-diformylfuran (DFF), 5-formyl-2-furancarboxylic acid (FFCA), 5-hydroxymethyl-2-furancarboxylic acid (HMFCA), and 2,5-furandicarboxylic acid (FDCA). 112 The theoretical voltage necessary for the oxidation of HMF to FDCA is merely 0.30 V, as shown in Eq. (25), significantly lower than conventional water electrolysis (1.23 V), enabling resource recovery and hydrogen production with low energy consumption.
Grabowski et al. achieved a 71 % yield of FDCA through the HMF electrochemical oxidation in a 1.0 M NaOH solution using nickel oxide/hydroxide catalysts. 113 Nickel-based catalysts have demonstrated efficacy in oxidizing HMF through an electrochemical-chemical coupling mechanism. 113
In conclusion, the integration of the HER with anodic electrooxidation of organics significantly reduces the cell voltage required for water splitting, leading to improved hydrogen production. Despite these advantages and recent advancements, this coupling technology is in its nascent stage, and its transformation efficiency remains inadequate. Therefore, the development of high-performance electrodes, precise identification of active sites, and strategic construction of the electrolyzer are crucial for future industrial applications.
4.2.4 Perspectives
The integration of the HER with anodic electrooxidation of organics significantly reduces the cell voltage required for water splitting, leading to improved hydrogen production. Despite these advantages and recent advancements, this coupling technology is in its nascent stage, and its transformation efficiency remains inadequate. To advance hybrid water electrolysis further, several critical aspects must be addressed to realize commercial and large-scale applicationss: 22 , 79
Accelerating catalyst discovery: Expediting the discovery of high-performance catalysts is imperative. For this regard, utilizing DFT calculations and machine learning can streamline catalyst design and enhance our comprehension of the structure-performance relationship. This approach holds potential for developing catalysts with superior activity and selectivity.
Employing advanced characterization Techniques: Integration of advanced in situ techniques and analytical methods is essential for comprehending the dynamic behavior of catalysts during electrochemical process. Techniques such as X-ray absorption spectroscopy and Raman spectroscopy facilitate real-time monitoring of reaction intermediates and catalyst structure evolution. This deeper understanding can inform the design of high-performance catalysts for efficient hybrid water splitting.
Developing bifunctional catalysts: The development of bifunctional catalysts based on earth-abundant elements is crucial for cost-effective and practical applications of hybrid water splitting. These catalysts can simultaneously facilitate multiple oxidation reactions, including biomass and pollutant oxidation, while efficiently harvesting hydrogen with minimal energy consumption. They offer a simplified and economically viable solution for large-scale hydrogen production.
Advancing membrane-based systems: Implementation of membrane-based systems is indispensable for practical HWE, enabling the efficient separation of anode and cathode products. Selecting membranes based on the applied electrolyte ensures optimal product separation and overall system performance. These membrane-based systems are pivotal for realizing large-scale and commercial application of HWE technology.
Integrating with solar systems: Integrating electrolyzers with solar systems represents a key strategy for further reducing energy costs in hybrid water electrolysis. By leveraging solar energy, the electrolysis process becomes more sustainable and economically viable for large-scale implementation.
4.3 Tandem water electrolysis
The tandem water electrolysis approach addresses challenges in H2 storage and transportation, 18 utilizing directly the in situ generated H2 during the process to produce valuable chemical fuels like methane (CH4) and ammonia (NH3) through efficient biological catalysts. 4 This method, promising for enhancing traditional H2 utilization, was pioneered by Nichols et al. 114 Their setup (Figure 6a) involved an airtight two-compartment electrochemical cell with Methanosarcina barkeri (M. barkeri) and CO2 in the cathodic chamber. Sustainable electricity or solar input drives water splitting using inorganic catalysts to produce H2, consumed by M. barkeri with CO2 to generate CH4. GC analysis of the headspace gas products revealed a linear increase in CH4 production over 11.5 h with applied currents (1.0−7.5 mA) and a Faradaic efficiency above 74 % (Figure 6b). The hybrid system operated robustly over 7 days with a Faradaic efficiency close to 75 % (Figure 6c). Isotope labeling confirmed that the carbon source for CH4 was the original CO2. The use of 13CO2 resulted in the detection of only 13CH4 via HRMS (Figure 6d), while the use of 12CO2 led to the observation of only 12CH4. This hybrid system offers significant flexibility in terms of electrocatalysts for HER, a range of abundant and biocompatible inorganic catalysts, including NiMo and NiS, can be employed. 114 Furthermore, when combined with photoelectrodes, it enables direct conversion of solar energy into chemical compounds, illustrating the potential of this tandem water electrolysis approach. 114

Electrocatalytic reduction of carbon dioxide to methane with a hybrid platinum/M. barkeri platform. (a) Schematic representation of tandem water electrolysis. (b, c) variation of CH4 production and faradaic efficiency with applied current. (d) Analysis of gas products using high-resolution mass spectrometry. (e and f) High-resolution mass spectrometry of headspace gases after electrolysis under an atmosphere of (e) 12CO2 and (f) 13CO2 in rich media. Error bars represent SD with n = 3 independent experiments in all cases. From. 114
Liu et al. 115 showcased a tandem water electrolysis system akin to the one described earlier, but with a twist: converting the generated H2 into biomass or alcohol fuels with CO2. In their setup, they employed Co–P as the HER electrocatalyst and Ralstonia eutropha as the biocatalyst. This system exhibited a CO2 reduction energy efficiency of approximately 50 %, capable of scrubbing 180 g of CO2 per kW-h of electricity. According to the research, Integration of this hybrid device with existing photovoltaic systems could yield a CO2 reduction energy efficiency of around 10 %, surpassing that of natural photosynthetic systems.
In another study, Liu et al. 116 introduced a novel tandem water electrolysis system designed to produce NH3 from N2 and H2O using autotrophic bacterium Xanthobacter autotrophicus. By harnessing solar energy and renewable electricity to drive water splitting, the approach holds promise as a renewable synthesis platform for the nitrogen reduction reaction (NRR). The researchers demonstrated that the production of NH3 and X. autotrophicus biofertilizer effectively links atmospheric N2 to plant biomass. This hybrid inorganic–biological approach for the NRR could potentially serve as a renewable biological and chemical synthesis platform, contingent upon the biomachinery involved in water splitting coupling.
Overall, these innovative tandem water electrolysis systems represent significant strides toward sustainable energy and chemical production, offering promising avenues for addressing environmental challenges and advancing renewable synthesis platforms. However, additional research is needed to optimize these systems for practical applications and to further enhance their efficiency and reliability.
4.4 Microbial electrolysis cells
Microbial electrolysis cells (MECs) represent a bioelectrochemical technology that leverages microbial metabolism to facilitate electrolysis reactions. Unlike microbial fuel cells (MFCs), which primarily generate electricity from microbial activity, MECs are specifically designed for hydrogen production via the electrochemical reduction of protons at the cathode. A schematic diagram illustrating commonly used reactors for MECs, including single- and two-chambered configurations, is depicted in Figure 7. 117 Within these systems, microbial communities oxidize organic matter found in sources such as wastewater or other organic substrates at the anode. This oxidation process releases electrons, protons, CO2, and other byproducts. 118 While protons migrate to the cathode, electrons travel through an external circuit to reach the cathode. At the cathode, these electrons participate in the reduction of protons derived from water, resulting in the production of hydrogen gas. 119

Schematic diagrams illustrating (a) a single-chambered microbial electrolysis cell (MEC) and (b) a two-chambered MEC designed specifically for hydrogen production. From. 117
The anodic oxidation reaction of various organic substrates in an MEC, along with their respective potentials at pH 7, is provided below: 117 , 120 , 121 , 122
Using the cathodic reaction of Eq. (2), the overall MEC reactions for acetate and glucose are given in equations 32 and 33. 117 According to the stoichiometries, MEC enables a theoretical maximum hydrogen yield of 4 mol per mole of acetate. However, a higher theoretical H2 yield of 12 mol/mol could be achieved using glucose.
The voltage required for HER (-0.412 V) at the cathode is provided through both microbial electrical supply (−0.3 V) and an extra-applied voltage. The additional applied voltage typically amounts to around 0.14 V, as theoretically calculated for acetate [Ecathode – Eanode = −0.41- (−0.3) = 0.14 V]. However, in practice, this voltage often exceeds 0.2 V due to considerations such as electrode overpotentials and ohmic losses. 123 Numerous studies reported a typically applied potential range of 0.5–1.0 V (Table 6), with a minimum requirement of 0.25 V to achieve reasonable current densities and hydrogen generation rates in practical-scale MECs. 124 Nonetheless, this applied potential remains significantly lower than that required for water electrolysis (i.e., oxygen yield, >1.23 V), which remains one of the most important advantages of MECs against other traditional techniques. MECs exhibit the capability to generate hydrogen using a diverse range of organic wastewater types as substrates (Table 6), offering the advantage of simultaneous pollutant treatment. Moreover, an array of external power sources, including fuel cells and solar cells, have been harnessed to supply energy to MECs (Figure 7b). Another critical aspect of the electrohydrogenesis process is the internal resistance of the MEC reactor, which is influenced by factors such as electrolyte concentration, construction materials of the MEC, and the presence or absence of ion exchange membranes. The primary objective in hydrogen production is to minimize the internal resistance of the MEC, facilitating efficient electron transfer across the electrodes. To achieve this goal, various strategies and approaches have been explored and implemented.
Applied potential pf microbial electrolysis cell (MEC) based on using wastewater and pure substrate. 117
MEC configuration | Substrate | Applied voltage (V) | H2 production m3/m3/d | Cathodic H2 recovery (%) | Ref. |
---|---|---|---|---|---|
Single chamber | Sludge wastewater | 0.9 | 0.038 | 15.56–20.05 | 280 |
Single chamber | Dairy wastewater | 0.7 | 0.2 | – | 281 |
Single chamber | P-glycerol | 0.5 | 0.8 | – | 122 |
Dual-chambers | Acetate | 1 | 50 | 90 | 282 |
Dual-chambers | Acetate | 0.6 | 1.1 | – | 283 |
Single chamber | Milk | 0.8 | 0.086 | 91 | 168 |
Pilot-scale MEC | Domestic wastewater | 0.6 | 0.015 | 70 | 284 |
Single chamber | Glucose | 0.5 | 0.83 | 55 | 122 |
Single chamber | Proteins | 0.6 | 0.42 | 87 | 285 |
Dual-chambers | Acetate | 0.45 | 0.37 | 90 | 126 |
Dual-chambers | Acetate | 0.5 | 0.02 | 57 | 286 |
Dual-chambers | Domestic wastewater | 0.5 | 0.01 | 42 | 166 |
Single chamber | Acetate | 0.8 | 3.12 | 96 | 134 , 283 |
Single chamber | Acetate | 0.6 | 0.69 | 87 | 132 |
Dual-chambers | Glycerol | 1.2 | 0.46 | 85 | 287 |
Dual-chambers | Vegetable wastewater | 0.8 | 0.025 | 11.29 | 288 |
Pilot-scale MEC | Urban wastewater | 0.5 | 0.041 | 90 | 289 |
Dual-chambers | Industrial wastewater | 1 | 0.03 | 38.5 | 290 |
4.4.1 Reactor configurations
Diverse reactor configurations and materials have been explored under different operational conditions, with single-chambered and two-chambered setups being the most prevalent (Table 6). 25 External voltage necessary for the process can be sourced from conventional power supplies or renewable energy, with solar energy standing out as the most promising, cleanest, efficient, and cost-effective option. When the process is conducted under solar illumination, it is referred to as solar-assisted MECs. 117
Initial research on MECs focused on H-type two-chambered configurations, typically constructed from glass or various plastics and separated by a membrane (Figure 7b). 125 , 126 , 127 Commonly used membranes include proton exchange membrane (PEM), anion-exchange membrane (AEM), cation exchange membrane (CEM), and bipolar membranes. 119 In these systems, microorganisms in the anodic chamber, known as exoelectrogens, oxidize organic matter to release electrons. These electrons flow through an external circuit to the cathodic chamber, where hydrogen gas is produced through reduction reactions. The membrane facilitates proton transfer while maintaining gas separation to preserve hydrogen purity. Two-chambered MECs typically achieve around 90 % substrate degradation in the anodic chamber and hydrogen production efficiency ranging from 60 % to 75 %. For instance, an H-type MEC fed with acetate generated 1.19 mmol of hydrogen at 0.540 V, 125 while a solar-assisted dual-chambered MEC exhibited faster hydrogen production with a Pt-coated electrode at 0.7 V. 128 Besides, higher H2 levels of up to 50 m3/m3/d with acetate under certain conditions, have been recorded in dual-chambered MECs (Table 6). Nevertheless, the operation of dual-chambered MECs is prone to energy losses attributed to factors like Ohmic loss, activation loss, and concentration loss. 126 Furthermore, the membrane itself contributes to an overpotential. Rozendal et al. 129 found that concentration energy loss could reach 0.38 V due to differences in concentration and pH gradient between the two chambers. However, hydrogen production using dual-chambered MECs may achieve higher levels, up to 50 m3/m3/d with acetate under certain conditions. The calculated cathodic hydrogen recovery and energy input for different cell designs, substrates, and operating voltages are summarized in Table 6.
In contrast, single-chambered MECs lack a membrane between the anode and cathode (Figure 7a), making them compact and cost-effective. These systems offer lower internal resistance and overpotential due to the absence of a membrane. Stacked systems with anodes surrounding the cathode further minimize resistance. 130 , 131 However, a major challenge with single-chambered MECs is methane interference from methanogens, compromising hydrogen purity. 129 , 132 Studies comparing mixed culture and pure culture in single-chambered MECs observed methane co-production in mixed cultures, negatively impacting hydrogen production rates. 132 , 133 Nevertheless, some studies have demonstrated high hydrogen recovery and production rates in single-chamber systems without membranes. 134 Studies on single-chambered MECs have reported hydrogen production rates up to 3.12 m3/m3/d under different operating conditions and feed compositions (Table 6). Despite achieving higher production rates, single-chambered MECs exhibit lower purity due to methane presence. 134 , 135 Suppression of methanogens in MEC reactors could enhance hydrogen purity, further improving bio-hydrogen quality. 135
4.4.2 Electrode materials
The effectiveness of MECs technology relies on the selection of suitable and cost-effective electrode materials, with carbon-based materials being the most common choices. 136 These electrodes could enable microbial organisms (MOs) to form biofilms, facilitating crucial electron transfer from the anode to the cathode for hydrogen production. 136 , 137 Anode materials must meet specific criteria to support biofilm formation and efficient electron transfer. Graphite-based materials like granular graphite beds, granules, graphite felt, and brushes have shown high performance and efficiency due to their porosity and surface area. 25 However, maintaining surface area while preventing clogging and biofouling poses a challenge. 137 Various electrode configurations, such as porous and plate types, offer advantages in conductivity, cost, and stability, but increasing their surface area for improved HER presents difficulties. 138 A list of adopted electrode materials used for MECs is shown in Table 7 and Figure 8. Further details are given in subsequent sections, with additional information can been retrieved in Ref. 25 , 117 , 119 , 123 , 139 , 140 .
Physio-Biological parameters involved in MEC performance. 25
Substrate | Chamber | Membrane | Cathode | Anod | Microorganism | Ref. |
---|---|---|---|---|---|---|
Dyes | Single | No | Graphite brush | TiO2 coated Ni foam | Mixed culture | 291 |
Acetate | Single | Ultrex (cation exchange membrane) | Carbon fbre brush | SiNWs photocathode | Mixed culture | 292 |
Acetate | Dual | AMI7001 (AEM) | Carbon fibers with stainless steel fram | Stainless steel mesh | Mixed culture | 293 |
Acetate | Dual | Nafion 117 | Non-wet-proofed carbon paper | Carbon paper coated with Pt –d | Mixed culture | 294 |
Wastewater activated sludge | Single | No | Carbon fiber brushes | Stainless steel mesh | Mixed culture | 295 |
Acetate | Single | No | Graphite plates | Ni mesh | Pure culture geobacter sulfurreducens | 156 |
Cellulose and acetate | Single | No | Carbon brushes | Carbon cloth | Mixed culture | 296 |
Corn stack | Single | No | Graphite felt | Pt coated carbon cloth | Mixed culture | 297 |
Domestic wastewater | Dual | Battery separator, rhinohide | Arbon felt with SS mesh | SS wool/mesh | Mixed culture | 179 |
Industrial wastewater/food processing WW | Single | No | Graphite fibre brushes | Carbon cloth with | Mixed culture | 167 |
Liquid of pressed municipal solid waste | Single | No | Graphite felt | Carbon cloth with Ti/RuO2 | Mixed culture | 298 |
Sugar industry effluents double | Dual | Yes (PEM Nafion 117) | Graphite | Ni foam, SS mesh 304 and Ni plat | Mixed culture | 299 |
Swine wastewater | Single | No | Graphite brush | Carbon cloth with 0.5 Pt mg/cm2 | Mixed culture | 300 |
Fermentation effluents | Single | No | Graphite brushes | Wet proof air cathode carbon cloth with 0.5 Pt mg/cm2 | Mixed culture | 301 |
Romanian Marine waters | Double | PEM Nafion | Glassy carbon | Graphite rod | Mixed culture | 302 |
Sodium acetate | Double | CEM ralex | Graphite felt | Graphite felt biocathodes (-0.7 V) | Mixed culture | 303 |
Sodium acetate | Single | No | Non-wet-proofed carbon cloth | Pt/C | Mixed culture | 304 |
Glucose + sodium acetate | Single | No | Graphite brush] | MoS2-GO nickel foam (NF)cathode | Mixed culture | 145 |

Various conventional and composite electrode types: a) Graphite rods, b) graphite felt, c) carbon cloth, d) carbon fiber brush, e) carbon paper, f) platinized wet-proof carbon cloth, g) stainless steel mesh, h) nickel mesh, i) carbon nanotubes, j) bioanode or biocathode, k) graphene. Adopted from. 25
4.4.2.1 Anode
An ideal anode material for MECs requires good biocompatibility, a large specific surface area, high conductivity, and corrosion resistance to support bacterial attachment, growth, and electron transfer. 119 It should also account for factors like cost, practicality, material availability, and stability. 119 Carbon-based materials are the most commonly used in MECs due to their affordability, conductivity, and abundance, 140 including carbon paper, carbon cloth, carbon felt, carbon mesh, graphite felt, reticulated vitrified carbongraphite rod, graphite fiber brush, and graphite. 25 While carbon paper, carbon cloth, carbon mesh, and carbon felt are all utilized as standard anode materials in MECs due to their high porosity, each material has its limitations. For example, carbon paper is fragile and lacks durability despite being thin and easily connected to open circuit wires, 28 whereas carbon cloth is more flexible and durable but relatively more expensive. 29 Carbon felt has practical properties but suffers from high resistance, 141 and carbon mesh is inexpensive but too thin to maintain structure. 141 Graphitic materials offer better electrical conductivity and stability than ordinary carbon-based materials but come at a higher price. 119 For instance, graphite rod is relatively inexpensive among graphitic materials but limited by low surface area, 142 while graphite fiber brush is considered a promising future anode material due to its high surface area, low resistance, and market availability. 142
Carbon nanotube (CNT) anodes represent a recent development in MEC research, known for their nanometer size, mechanical stability, lightweight, large surface area, excellent conductivity, and low material cost. 139 Higher electron transfer amplification has been reported with the use of CNTs. 139 , 143 Carbon nanotube/polyaniline nano-structure composites and CNTs with gold and palladium nanoparticles are also being established and assessed as anodic materials. 144 , 145 , 146 However, CNTs can be expensive and may present clogging issues, hindering their widespread application. 138
Nanocomposites comprising oxides of titanium (Ti), manganese (Mn), tin (Sn), iron (Fe), among others, combined with carbon-based materials, have also proven effective as anode materials. 25 These composite electrodes enhance performance, promote bacterial cell adhesion, and reduce ohmic loss. 147 TiO2-based electrodes offer biocompatibility and chemical stability, 148 while SnO2-based composite electrodes provide structural stability at a low cost. 149 Incorporating iron oxide in composite electrodes has shown to enhance extracellular electron transfer, while nickel oxide and manganese oxide are also under investigation. 148 Conducting polymers like polyaniline (PANI) along with Pt nanoparticles over graphene oxide and PANI, carbon nanotubes on graphite felt, and PEDOT-poly(3,4-ethylenedioxythiophene) over carbon cloth, graphite plate, and graphite felt have been utilized as doping materials to enhance conductivity and anodic performance. 150 , 151 PHBV-poly(3-hydroxybutyrate-co-3-hydroxyvalerate) mixed with carbon nanofibers has also been explored for improved conductivity and cell adhesion. 152
4.4.2.2 Cathode
A variety of cathode materials, including carbon-based materials and certain transition metals like stainless steel mesh and titanium, have been explored. 25 Cathode materials typically require coupling with a catalyst to enhance the rate of HER. While platinum is the preferred catalyst for HER due to its effectiveness, its high cost and limited availability restrict its widespread use. 119 Moreover, platinum is susceptible to contamination by compounds like sulfide and cyanide, negatively affecting its performance. 153 First-row transition metal compounds are emerging as promising alternatives to platinum-based catalysts for HER due to their abundance, stability, and decent catalytic activity. 4 Nickel-based materials and stainless steel are among the extensively studied catalysts due to their natural abundance, stable electrochemical properties, low cost, and high catalytic activity. 119 For instance, Ni–Co–P electrodeposited on stainless steel 316 cathode exhibited superior performance, achieving the highest coulombic efficiency and an overall energy hydrogen production rate. 154 Further studies have demonstrated the effectiveness of modified cathodes, such as Ni–W and Ni–Mo co-deposited on Ni foam, showing improved corrosion stability and electrocatalytic activity. 155 Phosphorization treatment of stainless-steel cathodes has also been shown to enhance hydrogen production current density. 119 Electroformed Ni mesh cathodes have shown promise as practical alternatives for hydrogen production in MECs, exhibiting comparable performance to Pt/CC cathodes. 156 Overall, the use of catalysts has significantly improved the hydrogen evolution performance of MEC cathodes. The increasing focus on developing Pt-free electrocatalysts for HER water electrolysis underscores the need for continued research in advancing electrocatalysts for hydrogen production in MECs.
4.4.3 Microbial populations: exoelectrogens
Extensive research on exoelectrogens utilized in MECs has been comprehensively reviewed by Kadier et al., 139 Liu et al. 140 and Parkhey and Gupta. 123 Among the exoelectrogenic bacteria for hydrogen production, Shewanella oneidensis and Shewanella putrefaciens from the Gammaproteobacteria family, and Geobacter sulfurreducens from the Deltaproteobacteria family are frequently reported. 139 Additionally, bacteria from other genetic groups such as Ochrobactrum and Rhodopseudomonas from Alphaproteobacteria, Arcobacter from Epsilonproteobacteria, Rhodoferax from Betaproteobacteria, Clostridium from Firmicutes, Propionibacterium from Actinobacteria, and Geothrix from Acidobacteria have also shown anodic respiratory capabilities. 139
These anode-respiring bacteria (ARB) employ one of three mechanisms for extracellular electron transfer: 140 , 157 (i) a direct contact, where exoelectrogens transfer electrons through their outer membrane directly to the anode-contacting region, 157 (ii) utilizing soluble electron shuttles such as phenazines (e.g., S. oneidensis MR 1), quinones (e.g., S. putrefaciens), and flavins (e.g., Shewanella), 158 , 159 , 160 and (iii) via nanowires or connecting appendages, which is commonly observed in Shewanella and Geobacter. 161 Geobacter species, in particular, have been extensively studied for electron transfer in bioelectrochemical systems, and their dominance in MEC reactors has been well demonstrated. 162 , 163 The analysis of anodic microbial communities revealed that MEC reactors with predominant populations of the Geobacteraceae family displayed higher power densities than those with varied bacterial communities. 162 , 163
While pure cultures offer a controlled system for MECs, 139 most MEC reactors utilize mixed consortia of bacteria for H2 generation. 123 Mixed cultures sourced from wastewaters and sludge present advantages over monocultures, as they are more robust to environmental changes and exhibit higher current densities, leading to enhanced coulombic efficiencies in MEC reactors. 164 , 165 , 166 , 167 However, a major drawback of mixed cultures from wastewaters is the potential for cross-contamination with methanogens, which utilize hydrogen gas to produce methane, thereby reducing overall hydrogen yields. 168 Three methods have been proposed to inhibit the growth of methanogens: 169 the first one is adjusting the environmental pH by utilizing a medium solution (pH 5.8) with phosphate buffer. The second is aerating the cathode for 15 min when methane content in the headspace exceeds 5 %. The last one is subjecting the anodes from MFCs to a 15-min boiling process before incorporating them into MECs.
4.4.4 Substrates
Substrates significantly impact hydrogen production in MECs (Figure 9), affecting yield based on their composition, concentration, and type. Various substances, such as acetic acid, lactic acid, butyric acid, glucose, glycerol, cellulose, starch, phenol, methanol, milk, and different types of wastewaters, can feed an MEC for hydrogen production. 25 , 123 , 125 , 126 , 140 , 170 , 171 These substrates are typically categorized into three groups: non-fermentable, fermentable, and various types of wastewaters. 25 Different substrates used in MECs are summarized Tables 6 and 7.

Hydrogen gas production rate of microbial electrolysis cells (MECs) depending on the substrate types. From. 117
Non-fermentable substrates, such as acetic acid, butyric acid, lactic acid, glucose, and propionic acid, are well-suited for MECs. 119 , 168 Among these, acetate stands out as the most extensively studied feed material. 139 Acetate exhibits promising results, with two-chamber MECs capable of achieving a theoretical hydrogen recovery of 4 mol H2/mol at an external applied voltage of 1 V. Moreover, single-chamber MECs fed with acetate demonstrate higher hydrogen production quantities due to reduced resistance and lower pH gradients within the system. 172 , 173
Fermentable organics, such as lignocellulosic biomass, stand out as highly viable feedstock for MECs due to their extensive study and widespread acceptance. 174 Among these, polymeric substrates like hemicellulose, cellulose, and aromatic polymers, including lignin, exhibit remarkable potential for HER. 25 Despite their cost-effectiveness and renewability, these substrates require conversion into monosaccharides or low molecular weight compounds to optimize hydrogen yield. For addressing this issue, pre-treatment methods are often employed to remove lignin from lignocellulosic biomass, thereby enhancing its digestibility within MECs. 25 , 171 A notable approach involves the combined use of dark fermentation and electro-hydrogenesis, which has shown promising results in dealing with recalcitrant lignocellulosic materials, leading to increased yields and improved rates of hydrogen production. 175 , 176 Comparative analyses have further underscored the advantages of winery wastewater over domestic wastewater as a feedstock for MECs, indicating higher outputs. 165 Effective processing of complex waste substrates involves hydrolysis and fermentation to break down complex macromolecules into simpler compounds, facilitating efficient hydrogen production. 25 , 176
Wastewater is abundant in contaminants and nutrients that require treatment. Additionally, wastewater treatment plants generate large amounts of waste activated sludge (WAS), posing a significant disposal challenge. Both wastewater and WAS can serve as valuable feedstocks for hydrogen production in microbial electrolysis cells (MECs). 170 , 177 MECs provide the dual benefit of contaminant removal and energy generation. 170 , 177 Various types of wastewater, including those rich in cellulose and glucose, as well as domestic and industrial wastewater have been considered in MECs. 165 , 170 , 177 , 178 Therefore, the integration of MECs into wastewater treatment infrastructure represents a step towards a circular economy, where waste products are transformed into valuable resources. By utilizing either WAS or wastewater itself as feedstocks for hydrogen production, MECs contribute to resource recovery and energy independence goals. 177
4.4.5 Challenges and future aspects
In summary, biological hydrogen production in MECs is gaining prominence due to its sustainable nature and low voltage requirements. However, there’s a need for improved performance and cost-effectiveness, driving the exploration of new building materials and optimized designs for MECs. 25 To enhance efficiency, reducing overpotential and associated losses is crucial. Strategies such as electrode optimization, membrane enhancements, and the use of high surface area electrodes show promise. 119 , 139 Studies have demonstrated significant overpotential reduction by introducing CO2 to the cathodic chamber and employing novel catalysts like Ni2P nanoparticles coated electrodes. 1 Additionally, integrating photocatalysts and exploring bio-cathodes offer avenues for further improvement. 117 Membrane development is also essential, with research focusing on cost-efficient alternatives to Nafion membranes and strategies to combat biofouling.
On a larger scale, pilot studies have highlighted challenges related to substrate utilization and membrane durability. 178 , 179 While phosphate buffer solution remains a popular choice for catholytes, further research into electrolytes is warranted. Understanding the role of exoelectrogens like Shewanella and Geobacter in biofilm formation and electron conduction is critical for optimizing MEC performance. 119 , 137 Techno-economic assessments and life cycle analyses are essential for evaluating MEC viability. 180
Future efforts should focus on scaling up MECs, utilizing real wastewater, and integrating MECs with other processes like anaerobic digestion for enhanced performance. While a succent review is provided here, a comprehensive overview of MEC technology and its challenges, further research is needed in electrolytes, efficient design, hydrogen storage, and materials to realize successful pilot-scale implementation. Additional reviews addressing these domains would be valuable for advancing MEC technology.
4.5 Application of external stimuli
4.5.1 Ultrasonic field
Ultrasonic irradiation stands as a widely embraced technique among researchers aiming to facilitate the gas bubbles detachment from electrodes. This method capitalizes on the phenomenon of acoustic cavitation, where tiny microbubbles form, expand, and abruptly collapse, inducing micro-agitation and intense microturbulence, known as acoustic streaming, in the vicinity of the electrode surface. 181 Essentially, ultrasonic vibrations cause the medium to oscillate back and forth, leading to a steady streaming away from the piezo location, as illustrated in Figure 10a. This oscillatory flow triggers an acoustic streaming effect, as depicted in Figure 11a, characterized by notable micro-agitation near the electrode surface and consequential mass transfer effects, fostering intense mixing. Moreover, the interaction between sound waves and gas bubbles gives rise to bubble streamers in an acoustic field, as shown in Figure 10b. 181 These oscillating gas bubble streamers, driven by primary and secondary Bjerknes forces, propel towards pressure antinodes at high speeds. This microstreaming effect engenders a plethora of physical forces, pivotal in various applications, including expediting the detachment of gas bubbles from the electrode. Among the significant forces generated during acoustic cavitation are microjets (Figure 10c) and shockwaves (Figure 10d). 181 Asymmetric collapse of cavitation bubbles, particularly near a boundary, forms high-speed liquid jets capable of pitting metal surfaces, whereas symmetrical collapse leads to the generation of high-intensity shock waves. Collectively, these ultrasound-induced actions in the solution effectively cleanse the electrode surface and augment the mass transfer of electrolytes from the solution to the electrode, thereby reducing overpotential through decreased solution and gas resistance. A 2019 review by Islam et al. 182 investigated the potential of power ultrasound to overcome limitations in electrochemical water splitting for hydrogen production. Their findings suggest ultrasound offers several advantages. Firstly, it can clean and activate electrode surfaces, leading to enhanced performance. Secondly, ultrasound promotes increased ions movement within the electrolyte and near the electrode surface, boosting mass transfer and facilitating the reaction. Thirdly, it effectively removes gas bubbles from both the electrolyte and the electrode surface, improving overall efficiency. Islam et al. also observed an increase in electrolytic efficiency (up to 15–20 %) attributed to a combination of factors like higher ion concentration and bubble detachment at the electrode surface, as shown in Figure 11. Photographic visualization of hydrogen bubbles on Pt wire electrode (Figure 12) clearly shown the cleaning effect of sonication.

Physical effects of acoustic cavitation, (a) acoustic streaming (oscillator is located at the bottom and liquid surface is at the top), (b) microstreamers, (c) microjet and (d) shockwaves generated by ultrasound and acoustic cavitation. Form. 181

Ultrasonication effect on (a) cell voltage (b) efficiency (ε) and (c) specific energy for hydrogen production (USA: Ultrasound-assisted). From. 182

Hydrogen production on Pt wire before and during sonication (US). From. 182
In the context of water sono-electrolysis, several studies have been done. Li et al. 183 investigated the effect of ultrasonic fields on water electrolysis in alkaline solutions. They observed a significant decrease in cell voltage with the application of an ultrasonic field, particularly at higher current densities and lower electrolyte concentrations. At a constant current density, the reduction in cell voltage under an ultrasonic field decreased with increasing NaOH concentration. The efficiency of H2 production increased by 5–18 % at higher current densities under an ultrasonic field, with energy savings of up to 10–25 %. Hung et al. 184 studied the combined effect of an ultrasonic field and PVDF-g-G1 BM synthesis membrane for H2 production. They found that H2 production efficiency was higher in electrolytic cells with an ultrasonic field compared to those without. Energy saving in H2 generation using PVDF-g-G1 BM membrane was 15.0–20.0 % and 8.0–12.0 % for cells operated with and without an US field, respectively. Lin and Hourng 185 investigated the impact of a 133 kHz-ultrasonic field with power ranging from 225 to 900 W, in conjunction with varied input voltages, electrolyte concentrations, and electrode gaps. They found that applying a 2.0 V potential along with the US field improved both concentration impedances and activity impedances, while also accelerating the release of hydrogen bubbles. Notably, ultrasonic power, electrolyte concentration, and electrode gap emerged as pivotal factors influencing water electrolysis performance. Under normal temperature conditions, with a 2 mm electrode gap, 4 V potential, and 40 wt% electrolyte concentration, the difference in current density between electrolysis with and without 225 W sonic power was 240 mA/cm2. Accounting for the power required for the ultrasonic wave, this resulted in a power saving of 3.5 kW and an economic power efficiency of 15 %. Additionally, they observed a 20 % increase in electrolysis efficiency with a 675 W ultrasonic power, in 10 wt% KOH, with a 5 mm electrode gap, and a specific input power.
Merabet and Kerboua 186 conducted a thorough investigation into alkaline water electrolysis, examining the impact of continuous and pulsed ultrasound under various operational parameters. They found that integrating continuous ultrasound under optimal conditions of KOH electrolyte and nickel foam electrodes led to a 7.78 % increase in cell current, corresponding to 64 % decrease in bubbles resistance. Interestingly, the effect of sonication was negligible at 45 °C (the optimal temperature), while room temperature showed better operation performance of the sono-electrolyzer in compared to silent conditions. Continuous mode sonication consistently outperformed pulsed mode, both kinetically and energetically. In a subsequent study, Merabet and Kerboua 187 introduced a membraneless sono-electrolyzer powered by photovoltaics (PV), utilizing alkaline electrolysis and indirect pulsed and continuous 60 kHz sonication. The results showed that indirect continuous sonication achieved the lowest electrode coverage percentage at 37 %, compared to 82 % without sonication and 58.3 % with pulsed ultrasonication. Furthermore, the bubble resistance decreased significantly from 569.81 mΩ without sonication to 132.54 mΩ with continuous sonication and 273.42 mΩ with pulsed sonication. Indirect continuous sonication evidenced the most effective mode for bubbles subtraction, resulting in a 76 % reduction in bubble resistance compared to no sonication and a 52 % reduction compared to pulsed sonication.
However, it should be noted that prolonged use of sonication must be approached with caution, as it can cause electrode degradation due to the asymmetric collapse of bubbles near solid surfaces, resulting in micro-jets with speeds of up to 100 m/s. 181 While low-frequency sonication (20–100 kHz) is beneficial for physical processes like cleaning and degassing, high-frequency sonication (above 100 kHz) may lead to electrolyte degradation, especially with organic electrolytes or ion liquids. The chemical effects of sonication in aqueous solutions are commonly used for the degradation of organic and inorganic contaminants. 188 , 189 Therefore, in water electrolysis for hydrogen production, frequency selection should not exceed 100 kHz. It is advisable to use low frequency with higher power to avoid any chemical effects of ultrasound. Additionally, while the power required to generate the ultrasonic field was found to be lower than the power saved from the enhanced efficiency, 183 there’s a suggestion that employing an ultrasonic field in water electrolysis might not be economically viable for hydrogen production due to the cost associated with the system required to generate the ultrasonic field. 185 This includes expenses related to transducers and utilities for measuring the acoustic pressure (e.g., hydrophones). This economic challenge could become more significant as operations scale up to industrial levels.
4.5.2 Magnetic field
Studies examining the impact of magnetic fields on electrolysis efficiency showed predominantly positive outcomes. 190 , 191 , 192 , 193 , 194 The enhancement effect of magnetic fields on electrochemical reactions primarily stems from magnetohydrodynamic convection induced by the Lorenz force, 195 , 196 which improved bubble detachment. Iida et al. 197 applied a strong magnetic field of 5 T to water electrolysis, resulting in significant reductions in cell voltage, particularly notable in alkaline solutions and at high current densities. Lin et al. 198 observed that this effect becomes more pronounced with decreased electrode spacing. Additionally, materials exhibiting ferromagnetism, like nickel, outperform paramagnetic (e.g., platinum) and diamagnetic (e.g., graphite) materials when subjected to a magnetic field. 198 For instance, using a nickel electrode with a 2 mm inter-electrode distance and a 4 V cell voltage led to a 14.60 % increase in current density. 198 Iida et al. 197 demonstrated improved water electrolysis efficiency through a magnetic field’s ability to reduce electrode overpotential [1817]. This effect was observed across various electrolyte conditions, including acidic media (0.05 M H2SO4) and alkaline solutions (4.46 M and 0.36 M KOH). Notably, the reduction in OER overpotential was greater compared to the hydrogen evolution reaction HER overpotential under the magnetic field. This difference was attributed to the varying sizes of gas bubbles produced during each process. The researchers proposed that magnetohydrodynamic (MHD) convection plays a crucial role. MHD convection influences bubble detachment from the electrode surface, leading to a significant decrease in the void fraction and the area covered by gas bubbles. This ultimately enhances the overall electrolysis efficiency. Matsushima et al., 199 through measuring supersaturation in the electrode-electrolyte interface and the mass transfer coefficient of dissolved H2 via the current interrupter method, concluded that a magnetic field promotes the mass transfer of dissolved H2, releasing supersaturation near the electrode. In another study, Matsushima et al. 200 found that the average bubble layer thickness (δ) on both cathode and anode was reduced by a magnetic field. Wang et al. 201 expanded on the explanation of magnetic field effects, suggesting that the rotational force applied to H2O molecules due to the interacting charge and induced charge of the Lorentz force from the induced magnetic field contributed to increased efficiency. They also proposed that the induced magnetic field diminished friction caused by hydrogen bonding. Consequently, ohmic voltage drops and reaction overpotential are noticeably reduced in the presence of a magnetic field, resulting in lower cell voltage for water electrolysis. Furthermore, it was found that the inclusion of a magnetic field allowed for a reduction in electrolyte concentration, 192 , 193 , 194 , 198 which would be valuable in lowering the overall electrolyte corrosive aspect, thus reducing operating costs associated with electrode corrosion.
4.5.3 Super gravity field
The slow detachment of bubbles in the electrolytic cell, especially at high cell voltages, is a significant issue in water electrolysis. Wang et al. 196 provide an insightful analysis of how a high-gravity field influences bubble detachment. Bubble movement within the cell is primarily controlled by interphase buoyancy, Δρg, where Δρ is the phase density difference and g is the applied acceleration. 196 Buoyancy acts as the driving force for bubble disengagement, causing bubbles to move in the direction of buoyancy during electrolysis. Applying an external force in alignment with the buoyancy direction can significantly enhance bubble disengagement. Consequently, in the context of water electrolysis, this approach can diminish bubble coverage on the electrode surface and reduce bubble dispersion in the electrolyte, particularly in the presence of a super gravity field. As a result, both reaction overpotential and ohmic voltage drop are mitigated, leading to a decrease in cell voltage and energy consumption.
Figure 13a depicts a captured photograph of the electrode surface during hydrogen bubble evolution under normal gravity conditions. 202 The image reveals numerous bubbles, each approximately 200 μm in diameter, adhering to the electrode surface. In contrast, Figure 13b illustrates the same process under a super gravity field, where only a few bubbles are observed, and their movement is barely discernible. To gain further insight into this phenomenon, let’s consider the scenario inverted and operate under microgravity conditions (with gravitational forces weaker than those on Earth). In such conditions, there is a higher intensity of larger bubble coverage on the electrodes, as demonstrated in Figure 13c and 13d. This scenario presents several challenges, including bubble adherence to the electrode and membrane, the accumulation of high gas volume fractions in the electrolyte, and the need to increase electrode spacing to facilitate bubble removal. 203 These factors contribute to increased overpotential and process costs.

The electrolytic formation of hydrogen and oxygen, (a,b) photographs of cathode surface during hydrogen bubble evolution process at 0.05 A/cm2 under (a) normal gravity condition (G = 1) and (b) super gravity field (G = 1 01) in 0.5 M H2SO4 aqueous solution. (c,d) gas (H2, O2) bubbles evolution in alkaline electrolyte under (c) terrestrial gravity and (d) microgravity conditions, at 8 s after starting water electrolysis using KOH solution. (e,f) Bubbles layer formed in the vicinity of electrodes’ surfaces in the absence (e) and under super gravity field (f). [(a, b): From Wang et al. 202 (c,d): From Matsushima et al. 305 (e,f): From Wang et al. 196
Therefore, the enhanced buoyancy force and reduced bubble volume under the super gravity field facilitate the rapid separation of bubbles from the electrolytic cell. 204 Consequently, the bubble layer thicknesses (δHG and δOG) near both the cathode and anode under the super gravity field (Figure 13f) are expected to be smaller than those (δH and δO) under normal gravity conditions (Figure 13e). The reduction in bubble layer thickness plays a crucial role in reducing electrolyte resistance and the ohmic voltage drop, primarily due to the diminished dispersion of bubbles within the electrolyte when exposed to the super gravity field. Importantly, the impact of the super gravity field on reducing cell voltage is more pronounced with higher current densities for water electrolysis, as indicated by the relationship between bubble layer thickness (δ) and its effect on cell voltage reduction. 204
Cheng et al. 205 were the first who demonstrate the effectiveness of a super gravity field in intensifying the water electrolysis process. They achieved a reduction in cell voltage of approximately 0.70 V at 3.0 kA/m2 under 190 g (a relative acceleration) compared to normal gravity environments. Similar findings were replicated by Wang et al. 202 , 204 The relationship between the cell voltage of water electrolysis under a super gravity field (UG) and the gravity coefficient (G) was expressed using the general formula: 204
where, U1 represents the cell voltage under normal gravity conditions (G = 1), D signifies the rate of change of cell voltage with the logarithm of G. The maximum energy saving percentage reaches approximately 17 %. 204 Cheng et al. 205 estimated that the potential energy savings could reach up to 30 kW in a 100 kA electrolytic cell, whereas it was only 2 kW for the super gravity field. The reduction in cell voltage was primarily attributed to a decrease in ohmic voltage drop compared to the reduction in reaction overpotential, particularly at higher current densities and gravity accelerations. 204
Lao et al. 203 also reported the increase of electro-hydrogen yield in a centrifugal acceleration field. At normal conditions (7.7 M KOH solution, 348 K), significant reductions in cell voltage were observed at an acceleration of approximately 16 g (∼500 rpm of rotational speed). The rotary electrolyzer achieved cell voltages about 0.25–0.5 V lower than equivalent static cells under similar conditions, depending on the current density. These voltages were comparable to those of fully developed pressurized cells in industrial settings, even without an effective electrode catalytic coating. Furthermore, at a higher acceleration of 41 g, the rotary cell sustained a current density of up to 13.5 kA/m2 without experiencing gas bubble blinding of the membranes and electrodes. This study demonstrates the potential for intensification compared to typical commercial systems with current densities of about 5 kA/m2.
4.6 Seawater electrolysis
Given the vast abundance of seawater, comprising over 96 % of the total water reservoir, and its economically viable potential, the conversion of seawater into hydrogen using renewable electricity represents a promising pathway toward achieving energy sustainability. 28 Recent advancements in research have significantly improved our understanding of seawater electrolysis mechanisms, leading to the formulation of design principles such as alkaline design criteria and the development of chloride blocking layers to enhance catalyst performance. 26 Many comprehensive reviews underscored both the significant advancements and the substantial challenges faced in seawater electrocatalysis. 26 , 27 , 28 , 29 In the subsequent sections, we delve into the essential aspects of seawater electrolysis, including challenges and recent advancements in electrode materials. Strategies for developing highly active and selective catalysts for seawater electrolysis, even in the presence of contaminants such as chloride, metal ions, and bio-organisms, are outlined, along with considerations for electrolyzer design. Additionally, various opportunities for advancing this promising technique are discussed. However, for in-depth analysis, readers are encouraged to refer to the aforementioned article reviews, i.e. 26 , 27 , 28 , 29
4.6.1 Challenges of seawater electrolysis
The complexity of seawater poses a significant challenge for seawater electrolysis, primarily due to the diverse range of components present, including microorganisms, sediment, and various ion species. Sediment and microorganisms, which primarily impact the cleanliness of seawater electrolyzers, can be addressed through filtration pretreatment. However, seawater ionic species (i.e., Cl−, Br−, SO42−, HCO3−, CO32−, F−, Mg2+, Na+, Ca2+, Cu2+, K+, Sr2+, and Cd2+) cannot be easily filtered out, these ions can lead to veritable issues at both the cathode and anode during electrolysis. Therefore, the influence of these species must be carefully considered in seawater electrolysis processes.
At the cathode, the significant challenge for HER arises from the dramatic rise in local pH near the cathode surface as the electrolysis current increases. This pH increase can lead to the formation of precipitates such as Ca(OH)2 and Mg(OH)2 from seawater cations, consequently blocking HER reactive sites. 44 Currently, two potential solutions are being explored: 28 (i) the addition of pH buffers to stabilize pH fluctuations in the seawater electrolysis system, and (ii) the design of suitable seawater electrolyzers to facilitate the separation of sediment from the cathode surface. Furthermore, certain metal ions dissolved in seawater, such as Cu2+, Cd2+, and Pb2+, have a tendency to deposit on the cathode surface at specific potentials. These deposits not only compete with HER but also significantly reduce electrode durability. 27 Therefore, catalyst designs should aim to suppress these undesirable electrochemical processes to achieve high selectivity for HER in seawater electrolysis.
At the anode, the OER is notably influenced by chemically active anions found in seawater. Specifically, chloride (Cl−) chemistry is expected to compete with the anodic OER due to its standard redox potential range of 0.9–1.6 V. 26 Pourbaix diagram incorporating both the OER and chloride chemistry is shown in Figure 14. This 29 , 206 diagram reveals that the chlorine evolution process (ClER) occurs at low pH (<3) (Eq. (35)), while hypochlorite production occurs at high pH (>7.5) (Eq. 336), representing the primary competing reactions for OER at the anode. Consequently, Cl− ions in seawater compete with the OER reaction, and the resulting Cl2 (oxidant) can lead to severe corrosion of the seawater electrolyzer.

The pourbaix diagram for the OER and the chloride chemistry in aqueous saline electrolyte. 0.5 M NaCl dissolved in this aqueous solution and no other chlorine sources, including the H2O/O2 and the Cl-/cl2/HOCl/ClO- redox couples. This figure depicts the potential-pH region where OER and chloride oxidation reactions may occur thermodynamically. The green line represents the thermodynamic equilibrium between H2O and O2. When the electrode potential is more positive than the green line, the OER process becomes thermodynamically possible. The red line shows the competing acidic oxidation of chloride to free gaseous Cl2. The black and purple lines mark the beginning of the oxidation of chloride to HOCl or OCl–. Form. 206
Although thermodynamics favor the OER, both ClER and hypochlorite production (Eqs. (35) and (36)) are two-electron reactions that are kinetically more favorable compared to the four-electron OER. Interestingly, when the electrolyte becomes alkaline, the product of Cl− oxidation transitions from HOCl to OCl−, and the thermodynamic overpotential maintains a consistent gap of 480 mV with OER (Figure 14a). Therefore, utilizing a high-performance OER catalyst under alkaline conditions and limiting the overpotential below 480 mV theoretically allows for the complete prevention of Cl− oxidation. 27 This principle, known as the ‘alkaline design criteria’ for seawater electrolysis, has been elucidated by Tong et al. 206 and Dionigi et al. 29
4.6.2 Electrocatalysts for seawater splitting
Seawater electrolysis presents a big challenge due to the presence of cationic metals and abundant corrosive chloride anions, necessitating robust and efficient electrocatalysts, particularly for the anode. 44 Figure 15 compares the performance of various tested catalysts under diverse testing conditions, including acidic, neutral, basic, simulated seawater, and natural seawater electrolytes. 28 Among the tested OER catalysts, in addition to those adopted in pure water, noble metals such as Ir and Ru, as well as transition metal oxides, hydroxides, nitrides, phosphorus-doped complexes, and layered double hydroxides (LDHs), have demonstrated high activity. 28 However, when considering only current density and overpotential, the performance differences among various OER catalysts are minimal (Figure 15), as only a portion of the currents is utilized for OER, with additional energy consumed by the oxidation of Cl−. Actual seawater electrolysis is even more complex, with many catalysts exhibiting excellent performance in simulated seawater but suffering from performance losses in natural saltwater due to the presence of numerous impurity ions. 44 The ClER reaction of Cl− competes with OER, while the deposition of heavy metal ions in seawater affects the catalytic performance of the cathode. 27 Subsequently, we will analyze numerous OER and HER electrocatalysts for seawater electrolysis.

Comparison of the activity of various HER and OER catalysts tested in acidic, neutral, and basic media. Different solutions (natural seawater, simulated seawater or buffered seawater) are distinguished by different color patches. The term overpotential (η) is used to define the overpotential at a certain current density (e.g., 5, 20 and 100 mA/cm2, labeled η5, η20 and η100). If there is no special description, it means that the current density is 10 mA/cm2. From. 28
Guided by DFT computation, Hsu et al. 207 has engineered an MHCM-z-BCC (MHCM: Transition metal hexacyanometallate; z-BCC: ZIF-67 on the basic cobalt carbonate surface) electrode, demonstrating remarkable selectivity towards the OER in seawater. While PtOER//PtHER and IrO2OER//PtHER electrolyzers produced 50 and 22 µM of Cl2, respectively, after 12 h in buffered seawater (pH ∼7), the MHCM-z-BCCOER//NiMoSHER electrolyzer exhibited no detectable Cl2 even after 100 h, indicating minimal CIER. Additionally, using the NiMoN@NiFeNOER//NiMoNHER electrolyzer in simulated seawater (1 M KOH with 0.5 M NaCl), only H2 and O2 gases were detected with a molar ratio close to 2:1, and a Faradaic efficiency of approximately 97.80 %, 208 demonstrating high selectivity for OER on the anode. Moreover, CoFe LDH displayed a Faradaic efficiency of ∼94.0 % at an overpotential (η) of 560 mV. 209 Nevertheless, deviations in Faradaic efficiency were observed for both O2 and H2 evolution during OER and HER, respectively. NiCoP exhibited approximately 96.5 % Faradaic efficiency for about 60 min, 210 while Ni5P4@Ni2+ δOδ(OH)2-δ displayed nearly 93.0 % Faradaic efficiency with minimal side of CIER reactions during electrolysis (the H2 yield increased linearly). 211 Similarly, Mn–NiO–Ni showed nearly 100 % Faradaic efficiency within 1 h, but deviations occurred after 1 h, with approximately 70.0 % Faradaic efficiency observed after 7 h, suggesting the presence of side reactions consuming about 30 % of the total charge. 212 Furthermore, NiRuIr–Graphene exhibited nearly 100 % Faradaic efficiency for 2 h during HER in seawater, 213 indicating minimal side reactions.
In more detail, Kuang et al. 214 demonstrates the effectiveness of dual-layered anode fabrication and subsequent activation in combating chloride corrosion for long-term stability in seawater electrolysis. For instance, their research shows that the NiFe-NiSx-Ni#OER//Ni–NiO–Cr2O3HER electrolyzer, equipped with an activated NiFe-NiSx-Ni anode, maintains negligible decay even at high current densities. Specifically, the activated NiFe-NiSx-Ni anode exhibits a low overpotential of ∼300 mV at 400 mA/cm2 in 1 M KOH with 0.5 M NaCl, significantly lower than the typical of ∼480 mV (threshold η). This reduced overpotential indicates its superior resistance to hypochlorite formation, a critical factor for ensuring selective oxygen evolution. Moreover, while the activated NiFe-NiSx-Ni anode remains stable for over 500 h, its counterpart, the activated NiSx-Ni anode, deteriorates rapidly in harsh electrolytes, such as 1 M KOH with 2 M NaCl. These findings underscore the importance of anode activation in achieving high corrosion resistance and OER activity, paving the way for efficient seawater electrolysis.
Yu et al. research 208 highlights the benefits of employing a metal nitride anode with a core–shell structure, particularly with a NiFe-based shell, for enhancing OER in seawater electrolysis. This configuration enables the in situ formation of thin amorphous layers during OER, improving corrosion resistance against chloride ions. Additionally, the core–shell design allows for electronic structure tuning, increased exposure of active sites, enhanced charge transfer, and improved gas evolution, collectively boosting OER efficiency in seawater. Their study demonstrates that the NiMoN@NiFeNOER//NiMoNHER electrolyzer maintains negligible decay even after 100 h of operation at 100 mA cm2 in. Furthermore, studies by Ding et al. 215 demonstrate that cellular stainless steel, characterized by high conductivity and tensile strength, facilitates efficient OER performance by exposing ample active sites and supporting effective charge transfer. Their research shows negligible decay over 110 h at 40 mA/cm2. Similarly, the nitrogen-rich, nanostructured Mo5N6, atomically thin, as observed by Jin et al., 216 enhances HER due to its modified electronic structure and increased active site exposure. Moreover, Lu et al. 212 find that manganese-doped nickel oxide/nickel heterostructures and nanoarray-structured NiCoP further optimize HER efficiency by facilitating charge transfer and gas evolution processes. Their study reports a significantly high activity with η10 of −170 mV in seawater, with reasonable stability maintained for 14 h at η10 of −140 mV. These findings underscore the importance of material design and structure in achieving high-performance electrocatalysis for seawater electrolysis.
Lv et al. 210 discovered that nanoarray structured NiCoP demonstrates high activity with η10 of approximately −287 mV in seawater, indicating its effectiveness in the HER. Moreover, it exhibits reasonable stability with η of about −290 mV for 20 h, suggesting its reliability for prolonged operation. Similarly, Huang et al. 211 reveals that the nanostructured Ni5P4 with a thin amorphous Ni5P4@Ni2+ δOδ(OH)2-δ layer displays significantly high activity, with η10 of approximately −144 mV in seawater. Furthermore, it exhibits negligible decay even at various current densities over 40 h, indicating its remarkable stability. Sarno et al. 213 investigated the performance of the nanostructured NiRuIr alloy anchored on graphene, revealing its substantial stability with approximately 90 % retention over 200 h in seawater. Additionally, Zhou et al. 217 on nanostructured T@Td-ReS2 with Re vacancies highlights enhanced HER performance in seawater. Interestingly, they observed that while the highest conductivity of the electrolyte is observed at approximately 45 °C, the highest hydrogen production rate occurs at about 35 °C. This discrepancy was attributed to the synergy between conductivity and dielectric constant, ultimately leading to the highest hydrogen production rate.
Zheng et al. 218 found that the PtMo alloy exhibits substantial activity with η10 of about −254.6 mV. Moreover, it maintains excellent stability, with 91.13 % retention at η of −800 mV for 173 h. Zhang et al. 219 compared the HER of nickel (Ni) alloys, including NiCo and NiCu, in seawater. Fabricated on pretreated Ti substrates via the electrodeposition method, NiCu displayed a nanostructure with a feather-like morphology, while NiCo exhibited a loose structure. The nanostructure of these alloys exposed abundant active sites, contributing to their improved HER performance. Specifically, in filtered seawater, NiCo exhibited an overpotential (η) of approximately −1,000 mV at a current density of 111 mA/cm2, with negligible decay observed over ∼10 h. Similarly, NiCu showed an overpotential of approximately −1,000 mV at a current density of 88 mA/cm2, with reasonable stability maintained over ∼10 h. Niu et al. 220 conducted a comparative study on the HER activity of RuCoMox alloys with RuCo alloy in filtered seawater. These alloys were deposited on pretreated titanium foil using cyclic voltammetry deposition. The RuCo alloy features nanoparticles morphology with an average size of approximately 50 nm, while the RuCoMo7 alloy exhibits a compact structure with cracks. In filtered seawater, RuCo demonstrates higher HER activity against RuCoMo, RuCoMo2, RuCoMo3, RuCoMo5, and RuCoMo7. Specifically, RuCo exhibits an overpotential η10 of approximately −387 mV, with 70 % retention at an overpotential of −1,200 mV over 12 h. Meanwhile, RuCoMox shows an η10 of approximately −550 mV, with negligible decay at an overpotential of −1,200 mV over 12 h. Li et al. 221 compared the HER performance of various PtRu-based alloys with Pt in seawater. The alloys, including PtRuMo, PtRuCo, PtRuNi, PtRuFe, and PtRuCr, were deposited on Ti mesh using cyclic voltammetry. These alloys exhibited a face-centered-cubic structure similar to Pt, with changes in crystal constants indicating alloying effects. Among the alloys, PtRuMo showed a nanostructure and the lowest charge transfer resistance, indicating enhanced charge transfer. The HER activity followed this order: PtRuMo > PtRuNi > PtRuCo > PtRuFe > PtRuCr > PtRu > Pt. PtRuMo exhibited the highest activity in seawater, with a low η10 of −196 mV and excellent stability, retaining 97.9 % efficiency for 172 h. PtRuNi, PtRuCo, PtRuFe, and PtRuCr also showed high activity with reasonable stability in seawater.
The bamboo-like multi-walled carbon nanotubes incorporating carbon-coated cobalt nanoparticles and nitrogen dopants produces a synergistic effect that enhances the performance of the HER. Research by Gao et al. 222 illustrates the remarkable activity and stability of Co/N–C for HER in seawater with a pH around 7. Co/N–C exhibits significantly higher activity in pH-buffered seawater compared to natural seawater, suggesting that neutralizing the seawater pH enhances HER activity. In pH-buffered seawater, Co/N–C achieves an η10 of around −250 mV, maintaining reasonable stability for 7 h with η10 of approximately −270 mV.
4.6.3 Seawater electrolyzers
Currently, two primary processes exist for producing green hydrogen through seawater electrolysis, with power supplied by renewable sources such as wind, solar, or tidal energy: one-step and two-step processes. 28 The one-step technique involves direct electrolysis of seawater to produce hydrogen, utilizing electrolyzers designed to handle seawater as a feedstock. In contrast, the two-step process first involves desalinating seawater to produce ultra-pure water, which is then electrolyzed to generate hydrogen. Advocates of the two-step approach argue that direct seawater splitting research lacks cost-effectiveness. 223 They contend that, considering the efficiency and costs associated with reverse osmosis technology, the expenses for initial seawater purification are significantly lower than those for seawater electrolysis. 223 , 224 Therefore, they propose focusing on the two-step method as a more economically viable option for green hydrogen production from seawater. 224
Seawater electrolyzers, on the other hand, are categorized according to their electrolyte type, with three notable technologies: 28 Direct Seawater Electrolyzer (DWE), Proton Exchange Membrane Water Electrolyzer (PEMWE), and Anion Exchange Membrane Water Electrolyzer (AEMWE). The characteristics of each type are summarized in Table 8.
Characteristics of the seawater electrolyzers. 28
Electrolyzer | DWE | PEMWE | AEMWE |
---|---|---|---|
Operating temp. | 20 °C | 50–80 °C | 40–60 °C |
Operating pressure | Standard | <70 bar | <35 bar |
Electrolyte | Seawater | Perfluorosulfonic acid membranes | Divinylbenzene polymer support with KOH or NaHCO3 |
Anode catalyst | / | IrO2 | Ni or NiFeCo alloys |
Cathode catalyst | / | Pt nanoparticles on carbon black | High surface area Ni |
Porous transport layer anode | / | Pt coated sintered porous Ti | Nickel foam |
Porous transport layer cathode | / | Sintered porous Ti or carbon cloth | Nickel foam or carbon cloth |
Bipolar plate anode | / | Pt-coated Ti | Ni-coated stainless steel |
Bipolar plate cathode | / | Au-coated Ti | Ni-coated stainless steel |
Briefly, DWE utilizes seawater as its electrolyte and requires specific electrode materials for corrosion resistance and energy efficiency. It involves a voltage of approximately 4.0 V to achieve 10 mA/cm2. The total investment cost exceeds 6,000$/kW, with estimated annual operation and maintenance (O&M) costs of 4.0–5.0 % of the investment cost. 28 PEMWE, on the other hand, relies on high-purity water and a solid polymer electrolyte (e.g., Nafion membrane). It is sensitive to impurities, with magnesium and calcium ions potentially impacting performance via precipitates deposition on the cathode. Investment costs range from 600 to 1,300 €/kW, with estimated annual O&M costs of 3.0–5.0 % of the investment cost. 28 Finally, AEMWE operates in an alkaline environment, allowing for the use of non-precious metal catalysts and less expensive materials. However, the technology is still under development, particularly regarding anionic membranes. 28 Anion exchange membranes are crucial to prevent unwanted migration of anions, such as chloride ions, during seawater electrolysis.
4.6.4 Future directions for seawater electrolysis
In summary, despite significant recent efforts in seawater electrolysis, 26 , 27 , 28 , 29 there remains ample room for improvement in achieving high-performance hydrogen production. Seawater electrolysis presents unique challenges due to its complex composition compared to freshwater electrolysis. The catalytic activity and stability of reported catalysts often fall short of practical application requirements due to insufficiently low overpotentials. Moreover, most studies utilize artificial saline water rather than real seawater. To enhance seawater electrolysis performance, several key directions for future research are proposed: 26 , 44
Conducting combined experimental and theoretical analyses to elucidate reaction pathways and active sites of catalysts: Multimetallic compounds and heterostructured catalysts show promise, but their complexity makes understanding reaction pathways and active sites challenging. Systematic research combining experimental data with theoretical analyses can guide the design of materials with desired structures and properties.
Employing in situ characterization methods to identify true active sites: Many electrocatalysts undergo surface changes during seawater electrolysis, potentially altering their active sites. In situ characterization techniques, such as Raman, XAS, and Fourier transform infrared spectroscopy, are essential for tracking changes in intermediates during catalytic reactions and guiding the design of high-efficiency catalysts.
Developing electrocatalysts with high activity and stability in seawater: Selectivity to hydrogen and oxygen evolution reactions in the presence of multiple cations and chloride anions is crucial. Modulating the electronic structure of active sites through alloying, heteroelement doping, vacancy engineering, and interface engineering can improve catalytic performance. Integration of different active materials into hybrid catalysts is also promising.
Designing advanced reactors specific to seawater electrolysis: Beyond catalysts, reactor design plays a critical role. Asymmetric reactor designs, with alkaline water in the anode chamber and seawater in the cathode chamber, offer advantages in Cl− diffusion and anode catalyst protection, crucial for seawater electrolysis.
5 Current challenges in electrolysis and electrolyzer development
Recent progress in electrolysis necessitates ongoing investigation to improve efficiency, scalability, and cost-effectiveness in hydrogen production. Table 9 outlines key research endeavors and goals in water electrolysis. Challenges include integrating electrolysis with renewable energy sources and energy grids while maintaining stability. 225 The absence of standardized methodologies in existing data, frequently sourced from commercial entities, underscores the necessity for universal testing protocols.
Key research areas, challenges, potential solutions, and objectives for water electrolysis development. 225
Research Area | Problem | Solution | Objective |
---|---|---|---|
Electrode materials | Identifying electrode materials that are both durable and efficient, capable of withstanding the harsh conditions of electrolysis, is paramount. Additionally, these materials must be cost-effective to ensure the economic viability of the electrolysis process. | Platinum, while effective as a catalyst, is prohibitively expensive, underscoring the necessity for alternative materials. | The advancement of electrodes, catalyst development, and material innovations is essential for enhancing electrolysis efficiency and reducing costs. |
Electrolyte stability | Electrolyte degradation is a critical factor influencing both the efficiency and lifespan of electrolysis cells | Research is underway to develop stable and conductive electrolytes, particularly for high-temperature and high-pressure electrolysis applications. | Investigating coatings, porous, nanostructured of bipolar plates and analyzing degradation mechanisms are essential areas of exploration |
Membrane technology | Enhancing proton exchange membranes in PEM electrolysis and ceramic membranes in SO electrolysis is crucial for advancing the technology. | Improving the durability, selectivity, and cost-effectiveness of membranes can significantly enhance the efficiency of electrolyzers. | Improving selectivity, durability, and ion conductivity |
Efficiency enhancement | The inefficiency of hydrogen production through electrolysis, which falls far short of theoretical limits, poses a significant barrier to its development, especially considering that hydrogen is primarily used as a carrier. | Achieving an efficiency increase of over 60 % would significantly enhance hydrogen’s potential for decarbonization. | Discovering configurations, materials, and operational parameters that optimize hydrogen production while minimizing energy consumption is imperative. |
Thermal management | Precise thermal management is vital to prevent overheating, maintain uniform temperature distribution within the electrolyzer cells, and mitigate thermal stress on the materials. | Enhancing thermal management systems is a crucial area of research and innovation, particularly in the realm of solid oxide electrolysis. | Precise thermal management will not only improve efficiency but also enhance durability and overall performance of these devices. |
Scale-up and cost reduction | Scaling-up electrolysis technology for industrial and commercial applications while simultaneously lowering production costs poses a significant challenge. | Economies of scale and advancements in manufacturing processes | Attaining a global electrolysis capacity capable of contributing to decarbonization, while maintaining a competitive production cost, is the ultimate goal. |
Electrolyzer lifespan | The operational lifespan of current electrolyzers typically ranges from approximately 10,000 to 40,000 h, varying based on the specific type and usage patterns. | Extending the lifespan of electrolyzers aims to reduce the frequency of replacements, thereby mitigating associated costs and environmental impacts. | Researching degradation mechanisms, investigating durable materials, and improving system design are crucial steps in addressing this challenge. |
Hydrogen purity | Guaranteeing the purity of the generated hydrogen is paramount, particularly for its utilization in fuel cell applications. | Continuously improving purification methods to effectively eliminate impurities such as oxygen and moisture remains an ongoing challenge. | Attain the requisite hydrogen purity is essential for diverse end applications in a sustainable manner. |
Dynamic operation | Electrolyzers need to dynamically adjust to fluctuating energy inputs to uphold stability and efficiency. | The fluctuating conditions pose a significant challenge for electrolyzers, especially when integrating with renewable energy sources (RESs), impeding smooth operation. | Efficiently adapting electrolysis systems to manage fluctuating renewable energy inputs is essential for their optimal performance. |
Standardization | The absence of defined standards, particularly concerning performance measurement and other parameters, poses a significant challenge. | Creating international standards and regulations for electrolysis technologies, safety protocols, and quality control is imperative. | Developing a comprehensive system of standards and regulations that facilitates the widespread adoption of water electrolysis technology is crucial. |
Lifecycle analysis | Conducting comprehensive lifecycle assessments is essential to gain a thorough understanding of the environmental footprint of electrolysis technologies. | Addressing the high lifecycle cost of electrolyzers, stemming from issues such as instability, shortened lifespan, and high initial costs, necessitates the development of innovative and cost-effective solutions. | Assessing the entire lifecycle, encompassing raw material extraction through end-of-life disposal or recycling, is crucial for understanding the environmental impact and sustainability of electrolysis technologies. |
The advancement of hydrogen production via water electrolysis necessitates a robust regulatory framework to address challenges in standardization, data reliability, and performance measurement methodologies. ISO 22734:2019(E) serves as a certification document for alkaline (ALK), proton-exchange membrane (PEM), and anion-exchange membrane (AEM) electrolyzers across various applications. 225 The ongoing work of ISO/TC197/WG34 focuses on developing standards for electrolyzer test protocols and safety requirements, emphasizing the need for standardized methods to ensure accurate performance evaluation. 225 As we delve into integrating electrolysis systems with energy grids and renewable sources, scalability becomes a critical issue. This multifaceted inquiry aims to enhance electrolysis capacities while ensuring harmonious integration within broader energy infrastructures, laying pathways for sustainable hydrogen production. Addressing specific challenges of AWE, PEM, and SO electrolysis technologies, such as cross-contamination, costly catalysts, and effective thermal management, can lead to advancements in efficiency and cost reduction. 225 Evaluating the environmental impact of these technologies, particularly in terms of CO2-equivalent emissions per kg of produced hydrogen, is crucial. 225 Gerloff’s comparative lifecycle assessment indicates that SO technology may have the lowest environmental impact, while ALK and PEM electrolysis technologies show similar performance. 226 However, this assessment remains complex and requires further investigation.
6 Conclusions
In this chapter, we have delved into the realm of cutting-edge strategies for low-temperature electrolysis aimed at efficient hydrogen production. Beginning with an introduction to electrochemical water splitting, we explored the fundamental principles underlying this process and the significance of hydrogen as a clean and sustainable energy carrier. Performance indicators were discussed to provide a framework for evaluating the efficiency and effectiveness of electrolysis systems. Moving forward, we delved into emerging strategies for electrochemical water splitting, highlighting innovative approaches such as decoupled, hybrid, and tandem electrolysis systems, as well as microbial electrolysis cells and seawater electrolysis. These strategies offer unique advantages in terms of efficiency, scalability, and environmental impact, paving the way for advancements in hydrogen production technology.
Decoupled water electrolysis presents a promising avenue for improving the overall efficiency of electrolysis systems by separately optimizing the oxygen and hydrogen evolution reactions. Hybrid approaches leverage the synergistic effects of multiple electrolysis techniques, while tandem electrolysis systems enable sequential hydrogen and oxygen production in a single device. Microbial electrolysis cells offer the potential for renewable and sustainable hydrogen production through microbial-driven electrochemical processes. Seawater electrolysis addresses the challenge of freshwater scarcity by utilizing abundant seawater resources for hydrogen production.
The exploration of these cutting-edge strategies underscores the dynamic nature of electrolysis research and development. Despite significant progress, challenges remain, including the need for cost reduction, scalability, and improved catalyst performance. However, with continued innovation and collaboration across interdisciplinary fields, the prospects for low-temperature electrolysis for hydrogen production are promising.
However, while significant strides have been made in the field of low-temperature electrolysis for hydrogen production, several challenges persist. Cost remains a key barrier to widespread adoption, necessitating the development of cost-effective materials and manufacturing processes. Scalability and system integration also pose challenges, particularly for large-scale deployment in industrial settings. Furthermore, catalyst stability and efficiency continue to be areas of active research, with ongoing efforts focused on developing durable and high-performance catalyst materials. Additionally, the integration of renewable energy sources, such as solar and wind power, into electrolysis systems presents opportunities for sustainable hydrogen production.
Looking ahead, collaboration between academia, industry, and government institutions will be essential for overcoming these challenges and realizing the full potential of low-temperature electrolysis for hydrogen production. By addressing these obstacles and harnessing the innovative strategies discussed in this chapter, we can accelerate the transition to a clean and sustainable hydrogen economy, driving towards a greener and more resilient energy future.
-
Research ethics: Not applicable.
-
Informed consent: Not applicable.
-
Author contributions: All of the parts of this research were conducted by S. Merouani, A. Dehane, O. Hamdaoui.
-
Use of Large Language Models, AI and Machine Learning Tools: None declared.
-
Conflict of interest: The authors state no conflict of interest.
-
Research funding: None declared.
-
Data availability: Not applicable.
References
1. Rousseau, R.; Etcheverry, L.; Roubaud, E.; Basséguy, R.; Délia, M.L.; Bergel, A. Microbial Electrolysis Cell (MEC): Strengths, Weaknesses and Research Needs from Electrochemical Engineering Standpoint. Appl. Energy 2020, 257, 113938; https://doi.org/10.1016/j.apenergy.2019.113938.Suche in Google Scholar
2. Nikolaidis, P.; Poullikkas, A. A Comparative Overview of Hydrogen Production Processes. Renew. Sustain. Energy Rev. 2017, 67, 597–611. https://doi.org/10.1016/j.rser.2016.09.044.Suche in Google Scholar
3. Hosseini, S. E.; Wahid, M. A. Hydrogen Production from Renewable and Sustainable Energy Resources: Promising Green Energy Carrier for Clean Development. Renew. Sustain. Energy Rev. 2016, 57, 850–866. https://doi.org/10.1016/j.rser.2015.12.112.Suche in Google Scholar
4. Li, W.; Tian, H.; Ma, L.; Wang, Y.; Liu, X.; Gao, X. Low-Temperature Water Electrolysis: Fundamentals, Progress, and New Strategies. Mater. Adv. 2022, 3, 5598–5644; https://doi.org/10.1039/d2ma00185c.Suche in Google Scholar
5. Arsad, S. R.; Arsad, A. Z.; Ker, P. J.; Hannan, M. A.; Tang, S. G. H. H.; Goh, S. M.; Mahlia, T. M. I. I. Recent Advancement in Water Electrolysis for Hydrogen Production: A Comprehensive Bibliometric Analysis and Technology Updates. Int. J. Hydrogen Energy 2024, 60, 780–801. https://doi.org/10.1016/j.ijhydene.2024.02.184.Suche in Google Scholar
6. Chi, J.; Yu, H. Water Electrolysis Based on Renewable Energy for Hydrogen Production. Chinese J. Catal. 2018, 39, 390–394. https://doi.org/10.1016/S1872-2067(17)62949-8.Suche in Google Scholar
7. Cong, Y.; Yi, B.; Song, Y. Hydrogen Oxidation Reaction in Alkaline Media: From Mechanism to Recent Electrocatalysts. Nano Energy 2018, 44, 288–303. https://doi.org/10.1016/j.nanoen.2017.12.008.Suche in Google Scholar
8. Chatenet, M.; Pollet, B. G.; Dekel, D. R.; Dionigi, F.; Deseure, J.; Millet, P.; Braatz, R. D.; Bazant, M. Z.; Eikerling, M.; Staffell, I.; Balcombe, P.; Shao-Horn, Y.; Schäfer, H. Water Electrolysis: From Textbook Knowledge to the Latest Scientific Strategies and Industrial Developments. Chem. Soc. Rev. 2022, 51, 4583–4762; https://doi.org/10.1039/d0cs01079k.Suche in Google Scholar PubMed PubMed Central
9. Vidas, L.; Castro, R. Recent Developments on Hydrogen Production Technologies: State-Of-The-Art Review with a Focus on Green-Electrolysis. Appl. Sci. 2021, 11, 11363; https://doi.org/10.3390/app112311363.Suche in Google Scholar
10. Ju, H. K.; Badwal, S.; Giddey, S. A Comprehensive Review of Carbon and Hydrocarbon Assisted Water Electrolysis for Hydrogen Production. Appl. Energy 2018, 231, 502–533; https://doi.org/10.1016/j.apenergy.2018.09.125.Suche in Google Scholar
11. Zeng, K.; Zhang, D. Recent Progress in Alkaline Water Electrolysis for Hydrogen Production and Applications. Prog. Energy Combust. Sci. 2010, 36, 307–326; https://doi.org/10.1016/j.pecs.2009.11.002.Suche in Google Scholar
12. Xu, Y.; Zhang, B. Recent Advances in Electrochemical Hydrogen Production from Water Assisted by Alternative Oxidation Reactions. ChemElectroChem 2019, 3214–3226; https://doi.org/10.1002/celc.201900675.Suche in Google Scholar
13. Carmo, M.; Fritz, D. L.; Mergel, J.; Stolten, D. A Comprehensive Review on PEM Water Electrolysis. Int. J. Hydrogen Energy 2013, 38, 4901–4934. https://doi.org/10.1016/j.ijhydene.2013.01.151.Suche in Google Scholar
14. Liu, G.; Xu, Y.; Yang, T.; Jiang, L. Recent Advances in Electrocatalysts for Seawater Splitting. Nano Mater. Sci. 2023, 5, 101–116. https://doi.org/10.1016/j.nanoms.2020.12.003.Suche in Google Scholar
15. Dincer, I.; Acar, C. Review and Evaluation of Hydrogen Production Methods for Better Sustainability. Int. J. Hydrogen Energy 2015, 40, 11094–11111. https://doi.org/10.1016/j.ijhydene.2014.12.035.Suche in Google Scholar
16. Ren, J. T.; Chen, L.; Wang, H. Y.; Tian, W. W.; Yuan, Z. Y. Water Electrolysis for Hydrogen Production: From Hybrid Systems to Self-Powered/Catalyzed Devices. Energy Environ. Sci. 2023, 17, 49–113; https://doi.org/10.1039/d3ee02467a.Suche in Google Scholar
17. Mohanty, B.; Bhanja, P.; Jena, B. K. An Overview on Advances in Design and Development of Materials for Electrochemical Generation of Hydrogen and Oxygen. Mater. Today Energy 2022, 23, 100902. https://doi.org/10.1016/j.mtener.2021.100902.Suche in Google Scholar
18. You, B.; Sun, Y. Innovative Strategies for Electrocatalytic Water Splitting. Acc. Chem. Res. 2018, 51, 1571–1580; https://doi.org/10.1021/acs.accounts.8b00002.Suche in Google Scholar PubMed
19. Phillips, R.; Edwards, A.; Rome, B.; Jones, D. R.; Dunnill, C. W. Minimising the Ohmic Resistance of an Alkaline Electrolysis Cell through Effective Cell Design. Int. J. Hydrogen Energy 2017, 1–9; https://doi.org/10.1016/j.ijhydene.2017.07.184.Suche in Google Scholar
20. Contreras, A.; Guirado, R.; Veziroglu, T. N. Design and Simulation of the Power Control System of a Plant for the Generation of Hydrogen via Electrolysis, Using Photovoltaic Solar Energy. Int. J. Hydrogen Energy 2007, 32, 4635–4640. https://doi.org/10.1016/j.ijhydene.2007.07.006.Suche in Google Scholar
21. McHugh, P. J.; Stergiou, A. D.; Symes, M. D. Decoupled Electrochemical Water Splitting: From Fundamentals to Applications. Adv. Energy Mater. 2020, 10, 1–21; https://doi.org/10.1002/aenm.202002453.Suche in Google Scholar
22. Chen, Z.; Wei, W.; Song, L.; Ni, B. J. Hybrid Water Electrolysis: A New Sustainable Avenue for Energy-Saving Hydrogen Production. Sustain. Horizons 2022, 1, 100002; https://doi.org/10.1016/j.horiz.2021.100002.Suche in Google Scholar
23. Du, L.; Sun, Y.; You, B. Hybrid Water Electrolysis: Replacing Oxygen Evolution Reaction for Energy-Efficient Hydrogen Production and beyond. Mater. Reports Energy 2021, 1, 100004; https://doi.org/10.1016/j.matre.2020.12.001.Suche in Google Scholar
24. Xu, J.; Amorim, I.; Li, Y.; Li, J.; Yu, Z.; Zhang, B.; Araujo, A.; Zhang, N.; Liu, L. Stable Overall Water Splitting in an Asymmetric Acid/Alkaline Electrolyzer Comprising a Bipolar Membrane Sandwiched by Bifunctional Cobalt-Nickel Phosphide Nanowire Electrodes. Carbon Energy 2020, 2, 646–655; https://doi.org/10.1002/cey2.56.Suche in Google Scholar
25. Gautam, R.; Nayak, J. K.; Ress, N. V.; Steinberger-Wilckens, R.; Ghosh, U. K. Bio-Hydrogen Production Through Microbial Electrolysis Cell: Structural Components and Influencing Factors. Chem. Eng. J. 2023, 455, 140535; https://doi.org/10.1016/j.cej.2022.140535.Suche in Google Scholar
26. Jiang, S.; Suo, H.; Zhang, T.; Liao, C.; Wang, Y.; Zhao, Q.; Lai, W. Recent Advances in Seawater Electrolysis. Catalysts 2022, 12, 123; https://doi.org/10.3390/catal12020123.Suche in Google Scholar
27. Mohammed-ibrahim, J.; Moussab, H. Materials Science for Energy Technologies Recent Advances on Hydrogen Production through Seawater Electrolysis. Mater. Sci. Energy Technol. 2020, 3, 780–807; https://doi.org/10.1016/j.mset.2020.09.005.Suche in Google Scholar
28. Gao, F.; Yu, P.; Gao, M. Seawater Electrolysis Technologies for Green Hydrogen Production: Challenges and Opportunities. Curr. Opin. Chem. Eng. 2022, 36, 100827; https://doi.org/10.1016/j.coche.2022.100827.Suche in Google Scholar
29. Dionigi, F.; Reier, T.; Pawolek, Z.; Gliech, M.; Strasser, P. Design Criteria, Operating Conditions, and Nickel-Iron Hydroxide Catalyst Materials for Selective Seawater Electrolysis. ChemSusChem 2016, 9, 962–972; https://doi.org/10.1002/cssc.201501581.Suche in Google Scholar PubMed
30. Buttler, A.; Spliethoff, H. Current Status of Water Electrolysis for Energy Storage, Grid Balancing and Sector Coupling via Power-To-Gas and Power-To-Liquids: A Review. Renew. Sustain. Energy Rev. 2018, 82, 2440–2454. https://doi.org/10.1016/j.rser.2017.09.003.Suche in Google Scholar
31. Schmidt, O.; Gambhir, A.; Staffell, I.; Hawkes, A.; Nelson, J.; Few, S. Future Cost and Performance of Water Electrolysis: An Expert Elicitation Study. Int. J. Hydrogen Energy 2017, 42, 30470–30492; https://doi.org/10.1016/j.ijhydene.2017.10.045.Suche in Google Scholar
32. Lehner, M.; Tichler, R.; Steinmller, H.; Kopper, M.; Steinmüller, H.; Koppe, M. Power-to-Gas: Technology and Business Models; Springer Cham: Switzerland, 2014.Suche in Google Scholar
33. Zhang, X.; Chan, S. H.; Ho, H. K.; Tan, S.-C.; Li, M.; Li, G.; Li, J.; Feng, Z. Towards a Smart Energy Network: The Roles of Fuel/Electrolysis Cells and Technological Perspectives. Int. J. Hydrogen Energy 2015, 40, 6866–6919. https://doi.org/10.1016/j.ijhydene.2015.03.133.Suche in Google Scholar
34. Laguna-Bercero, M. A. Recent Advances in High Temperature Electrolysis Using Solid Oxide Fuel Cells: A Review. J. Power Sources 2012, 203, 4–16. https://doi.org/10.1016/j.jpowsour.2011.12.019.Suche in Google Scholar
35. Schoots, K.; Ferioli, F.; Kramer, G. J.; van der Zwaan, B. C. C. Learning Curves for Hydrogen Production Technology: An Assessment of Observed Cost Reductions. Int. J. Hydrogen Energy 2008, 33, 2630–2645. https://doi.org/10.1016/j.ijhydene.2008.03.011.Suche in Google Scholar
36. Bertuccioli, L.; Chan, A.; Hart, D.; Lehner, F.; Madden, B.; Standen, E. Development of Water Electrolysis in the European Unione; 2014;Suche in Google Scholar
37. Hart, D.; Lehner, F.; Rose, R.; Lewis, J.; Klippenstein, M. The Fuel Cell Industry Review 2015; E4Tech: London, 2015.Suche in Google Scholar
38. Niu, S.; Li, S.; Du, Y.; Han, X.; Xu, P. How to Reliably Report the Overpotential of an Electrocatalyst. ACS Energy Lett. 2020, 5, 1083–1087; https://doi.org/10.1021/acsenergylett.0c00321.Suche in Google Scholar
39. Wei, C.; Rao, R. R.; Peng, J.; Huang, B.; Stephens, I. E. L.; Risch, M.; Xu, Z. J.; Shao-Horn, Y. Recommended Practices and Benchmark Activity for Hydrogen and Oxygen Electrocatalysis in Water Splitting and Fuel Cells. Adv. Mater. 2019, 31, 1806296. https://doi.org/10.1002/adma.201806296.Suche in Google Scholar PubMed
40. Anantharaj, S.; Ede, S. R.; Karthick, K.; Sam Sankar, S.; Sangeetha, K.; Karthik, P. E.; Kundu, S. Precision and Correctness in the Evaluation of Electrocatalytic Water Splitting: Revisiting Activity Parameters with a Critical Assessment. Energy Environ. Sci. 2018, 11, 744–771; https://doi.org/10.1039/C7EE03457A.Suche in Google Scholar
41. Song, F.; Li, W.; Yang, J.; Han, G.; Liao, P.; Sun, Y. Interfacing Nickel Nitride and Nickel Boosts Both Electrocatalytic Hydrogen Evolution and Oxidation Reactions. Nat. Commun. 2018, 9; https://doi.org/10.1038/s41467-018-06728-7.Suche in Google Scholar PubMed PubMed Central
42. Stoerzinger, K. A.; Risch, M.; Han, B.; Shao-Horn, Y. Recent Insights into Manganese Oxides in Catalyzing Oxygen Reduction Kinetics. ACS Catal. 2015, 5, 6021–6031; https://doi.org/10.1021/acscatal.5b01444.Suche in Google Scholar
43. Yu, L.; Sun, S.; Li, H.; Xu, Z. J. Effects of Catalyst Mass Loading on Electrocatalytic Activity: An Example of Oxygen Evolution Reaction. Fundam. Res. 2021, 1, 448–452. https://doi.org/10.1016/j.fmre.2021.06.006.Suche in Google Scholar
44. Mishra, A.; Park, H.; El-Mellouhi, F.; Suk Han, D. Seawater Electrolysis for Hydrogen Production: Technological Advancements and Future Perspectives. Fuel 2024, 361, 130636; https://doi.org/10.1016/j.fuel.2023.130636.Suche in Google Scholar
45. Cao, B.; Veith, G. M.; Neuefeind, J. C.; Adzic, R. R.; Khalifah, P. G. Mixed Close-Packed Cobalt Molybdenum Nitrides as Non-Noble Metal Electrocatalysts for the Hydrogen Evolution Reaction. J. Am. Chem. Soc. 2013, 135, 19186–19192; https://doi.org/10.1021/ja4081056.Suche in Google Scholar PubMed
46. Santos, D. M. F.; Sequeira, C. A. C.; Macciò, D.; Saccone, A.; Figueiredo, J. L. Platinum–Rare Earth Electrodes for Hydrogen Evolution in Alkaline Water Electrolysis. Int. J. Hydrogen Energy 2013, 38, 3137–3145. https://doi.org/10.1016/j.ijhydene.2012.12.102.Suche in Google Scholar
47. Liang, Z.; Ahn, H. S.; Bard, A. J. A Study of the Mechanism of the Hydrogen Evolution Reaction on Nickel by Surface Interrogation Scanning Electrochemical Microscopy. J. Am. Chem. Soc. 2017, 139, 4854–4858; https://doi.org/10.1021/jacs.7b00279.Suche in Google Scholar PubMed
48. Li, W.; Gao, X.; Xiong, D.; Wei, F.; Song, W.-G.; Xu, J.; Liu, L. Hydrothermal Synthesis of Monolithic Co3Se4 Nanowire Electrodes for Oxygen Evolution and Overall Water Splitting with High Efficiency and Extraordinary Catalytic Stability. Adv. Energy Mater. 2017, 7, 1602579. https://doi.org/10.1002/aenm.201602579.Suche in Google Scholar
49. Anantharaj, S.; Kundu, S. Do The Evaluation Parameters Reflect Intrinsic Activity of Electrocatalysts in Electrochemical Water Splitting? ACS Energy Lett. 2019, 4, 1260–1264; https://doi.org/10.1021/acsenergylett.9b00686.Suche in Google Scholar
50. Li, W.; Jiang, N.; Hu, B.; Liu, X.; Song, F.; Han, G.; Jordan, T. J.; Hanson, T. B.; Liu, T. L.; Sun, Y. Electrolyzer Design for Flexible Decoupled Water Splitting and Organic Upgrading with Electron Reservoirs. Chem 2018, 4, 637–649; https://doi.org/10.1016/j.chempr.2017.12.019.Suche in Google Scholar
51. Ghassemzadeh, L.; Kreuer, K.-D.; Maier, J.; Müller, K. Chemical Degradation of Nafion Membranes under Mimic Fuel Cell Conditions as Investigated by Solid-State NMR Spectroscopy. J. Phys. Chem. C 2010, 114, 14635–14645; https://doi.org/10.1021/jp102533v.Suche in Google Scholar
52. Berger, A.; Segalman, R. A.; Newman, J. Material Requirements for Membrane Separators in a Water-Splitting Photoelectrochemical Cell. Energy Environ. Sci. 2014, 7, 1468–1476; https://doi.org/10.1039/C3EE43807D.Suche in Google Scholar
53. Symes, M. D.; Cronin, L. Decoupling Hydrogen and Oxygen Evolution during Electrolytic Water Splitting Using an Electron-Coupled-Proton Buffer. Nat. Chem. 2013, 5, 403–409; https://doi.org/10.1038/nchem.1621.Suche in Google Scholar PubMed
54. Paul, A.; Symes, M. D. Decoupled Electrolysis for Water Splitting. Curr. Opin. Green Sustain. Chem. 2021, 29, 100453; https://doi.org/10.1016/j.cogsc.2021.100453.Suche in Google Scholar
55. Ho, A.; Zhou, X.; Han, L.; Sullivan, I.; Karp, C.; Lewis, N. S.; Xiang, C. Decoupling H2(g) and O2(g) Production in Water Splitting by a Solar-Driven V3+/2+(Aq,H2SO4)|KOH(Aq) Cell. ACS Energy Lett. 2019, 4, 968–976; https://doi.org/10.1021/acsenergylett.9b00278.Suche in Google Scholar
56. Goodwin, S.; Walsh, D. A. Closed Bipolar Electrodes for Spatial Separation of H2 and O2 Evolution during Water Electrolysis and the Development of High-Voltage Fuel Cells. ACS Appl. Mater. Interfaces 2017, 9, 23654–23661; https://doi.org/10.1021/acsami.7b04226.Suche in Google Scholar PubMed
57. Rausch, B.; Symes, M. D.; Cronin, L. A Bio-Inspired, Small Molecule Electron-Coupled-Proton Buffer for Decoupling the Half-Reactions of Electrolytic Water Splitting. J. Am. Chem. Soc. 2013, 135, 13656–13659; https://doi.org/10.1021/ja4071893.Suche in Google Scholar PubMed
58. Kirkaldy, N.; Chisholm, G.; Chen, J.-J.; Cronin, L. A Practical, Organic-Mediated, Hybrid Electrolyser that Decouples Hydrogen Production at High Current Densities. Chem. Sci. 2018, 9, 1621–1626; https://doi.org/10.1039/C7SC05388F.Suche in Google Scholar PubMed PubMed Central
59. Zhao, Y.; Ding, C.; Zhu, J.; Qin, W.; Tao, X.; Fan, F.; Li, R.; Li, C. A Hydrogen Farm Strategy for Scalable Solar Hydrogen Production with Particulate Photocatalysts. Angew. Chemie (International Ed. 2020, 59, 9653–9658; https://doi.org/10.1002/anie.202001438.Suche in Google Scholar PubMed
60. Chen, L.; Dong, X.; Wang, Y.; Xia, Y. Separating Hydrogen and Oxygen Evolution in Alkaline Water Electrolysis Using Nickel Hydroxide. Nat. Commun. 2016, 7, 1–8; https://doi.org/10.1038/ncomms11741.Suche in Google Scholar PubMed PubMed Central
61. Liu, X.; Chi, J.; Dong, B.; Sun, Y. Recent Progress in Decoupled H2 and O2 Production from Electrolytic Water Splitting. ChemElectroChem 2019, 6, 2157–2166; https://doi.org/10.1002/celc.201801671.Suche in Google Scholar
62. Wallace, A. G.; Symes, M. D. Decoupling Strategies in Electrochemical Water Splitting and beyond. Joule 2018, 2, 1390–1395; https://doi.org/10.1016/j.joule.2018.06.011.Suche in Google Scholar
63. MacDonald, L.; McGlynn, J. C.; Irvine, N.; Alshibane, I.; Bloor, L. G.; Rausch, B.; Hargreaves, J. S. J.; Cronin, L. Using Earth Abundant Materials for the Catalytic Evolution of Hydrogen from Electron-Coupled Proton Buffers. Sustain. Energy Fuels 2017, 1, 1782–1787; https://doi.org/10.1039/C7SE00334J.Suche in Google Scholar
64. Rausch, B.; Symes, M. D.; Chisholm, G.; Cronin, L. Decoupled Catalytic Hydrogen Evolution from a Molecular Metal Oxide Redox Mediator in Water Splitting. Science (80-. ) 2014, 345, 1326–1330; https://doi.org/10.1126/science.1257443.Suche in Google Scholar PubMed
65. Landman, A.; Dotan, H.; Shter, G. E.; Wullenkord, M.; Houaijia, A.; Maljusch, A.; Grader, G. S.; Rothschild, A. Photoelectrochemical Water Splitting in Separate Oxygen and Hydrogen Cells. Nat. Mater. 2017, 16, 646–651; https://doi.org/10.1038/nmat4876.Suche in Google Scholar PubMed
66. Landman, A.; Halabi, R.; Dias, P.; Dotan, H.; Mehlmann, A.; Shter, G. E.; Halabi, M.; Naseraldeen, O.; Mendes, A.; Grader, G. S.; Rothschild, A. Decoupled Photoelectrochemical Water Splitting System for Centralized Hydrogen Production. Joule 2020, 4, 448–471; https://doi.org/10.1016/j.joule.2019.12.006.Suche in Google Scholar
67. Ma, Y.; Dong, X.; Wang, Y.; Xia, Y. Decoupling Hydrogen and Oxygen Production in Acidic Water Electrolysis Using a Polytriphenylamine-Based Battery Electrode. Angew. Chemie Int. Ed. 2018, 57, 2904–2908. https://doi.org/10.1002/anie.201800436.Suche in Google Scholar PubMed
68. Dotan, H.; Landman, A.; Sheehan, S. W.; Malviya, K. D.; Shter, G. E.; Grave, D. A.; Arzi, Z.; Yehudai, N.; Halabi, M.; Gal, N.; Hadari, N.; Cohen, C.; Rothschild, A.; Grader, G. S. Decoupled Hydrogen and Oxygen Evolution by a Two-step Electrochemical–Chemical Cycle for Efficient Overall Water Splitting. Nat. Energy 2019, 4, 786–795; https://doi.org/10.1038/s41560-019-0462-7.Suche in Google Scholar
69. Bienvenu, G. Method for Co-generation of Electric Energy and Hydrogen, US Patent No. 8,617,766, Filed July 27, 2010, and Published December 31 2013.Suche in Google Scholar
70. You, B.; Liu, X.; Liu, X.; Sun, Y. Efficient H2 Evolution Coupled with Oxidative Refining of Alcohols via A Hierarchically Porous Nickel Bifunctional Electrocatalyst. ACS Catal. 2017, 7, 4564–4570; https://doi.org/10.1021/acscatal.7b00876.Suche in Google Scholar
71. Chen, G.-F.; Luo, Y.; Ding, L.-X.; Wang, H. Low-Voltage Electrolytic Hydrogen Production Derived from Efficient Water and Ethanol Oxidation on Fluorine-Modified FeOOH Anode. ACS Catal. 2018, 8, 526–530; https://doi.org/10.1021/acscatal.7b03319.Suche in Google Scholar
72. Huang, Y.; Chong, X.; Liu, C.; Liang, Y.; Zhang, B. Boosting Hydrogen Production by Anodic Oxidation of Primary Amines over a NiSe Nanorod Electrode. Angew. Chemie Int. Ed. 2018, 57, 13163–13166. https://doi.org/10.1002/anie.201807717.Suche in Google Scholar PubMed
73. Liu, W.-J.; Dang, L.; Xu, Z.; Yu, H.-Q.; Jin, S.; Huber, G. W. Electrochemical Oxidation of 5-Hydroxymethylfurfural with NiFe Layered Double Hydroxide (LDH) Nanosheet Catalysts. ACS Catal. 2018, 8, 5533–5541; https://doi.org/10.1021/acscatal.8b01017.Suche in Google Scholar
74. Babar, P.; Lokhande, A.; Karade, V.; Lee, I. J.; Lee, D.; Pawar, S.; Kim, J. H. Trifunctional Layered Electrodeposited Nickel Iron Hydroxide Electrocatalyst with Enhanced Performance towards the Oxidation of Water, Urea and Hydrazine. J. Colloid Interface Sci. 2019, 557, 10–17. https://doi.org/10.1016/j.jcis.2019.09.012.Suche in Google Scholar PubMed
75. Zhang, J.-Y.; Wang, H.; Tian, Y.; Yan, Y.; Xue, Q.; He, T.; Liu, H.; Wang, C.; Chen, Y.; Xia, B. Y. Anodic Hydrazine Oxidation Assists Energy-Efficient Hydrogen Evolution over a Bifunctional Cobalt Perselenide Nanosheet Electrode. Angew. Chemie Int. Ed. 2018, 57, 7649–7653. https://doi.org/10.1002/anie.201803543.Suche in Google Scholar PubMed
76. Yu, Z.-Y.; Lang, C.-C.; Gao, M.-R.; Chen, Y.; Fu, Q.-Q.; Duan, Y.; Yu, S.-H. Ni–Mo–O Nanorod-Derived Composite Catalysts for Efficient Alkaline Water-To-Hydrogen Conversion via Urea Electrolysis. Energy Environ. Sci. 2018, 11, 1890–1897; https://doi.org/10.1039/C8EE00521D.Suche in Google Scholar
77. Kahlstorf, T.; Hausmann, J. N.; Sontheimer, T.; Menezes, P. W. Challenges for Hybrid Water Electrolysis to Replace the Oxygen Evolution Reaction on an Industrial Scale. Glob. Challenges 2023, 7, 1–11; https://doi.org/10.1002/gch2.202200242.Suche in Google Scholar PubMed PubMed Central
78. Fan, L.; Wang, D.; Ma, K.; Zhou, C.-A.; Yue, H. Recent Advances in Hydrogen Production from Hybrid Water Electrolysis through Alternative Oxidation Reactions. ChemCatChem 2024, 16, e202301332. https://doi.org/10.1002/cctc.202301332.Suche in Google Scholar
79. Wang, Y.; Lu, S.; Yang, M.; Zhang, Z.; Zhang, J. Recent Advances in Anode Catalysts for Waste Valorization through Hybrid Water Electrolysis: Towards Sustainability beyond Hydrogen Production. Mater. Today Sustain. 2024, 25, 100630; https://doi.org/10.1016/j.mtsust.2023.100630.Suche in Google Scholar
80. Zhou, Q.; Shen, Z.; Zhu, C.; Li, J.; Ding, Z.; Wang, P.; Pan, F.; Zhang, Z.; Ma, H.; Wang, S.; Zhang, H. Nitrogen-Doped CoP Electrocatalysts for Coupled Hydrogen Evolution and Sulfur Generation with Low Energy Consumption. Adv. Mater. 2018, 30, 1800140. https://doi.org/10.1002/adma.201800140.Suche in Google Scholar PubMed
81. Sun, F.; He, D.; Yang, K.; Qiu, J.; Wang, Z. Hydrogen Production and Water Desalination with On-Demand Electricity Output Enabled by Electrochemical Neutralization Chemistry. Angew. Chemie Int. Ed. 2022, 61, e202203929. https://doi.org/10.1002/anie.202203929.Suche in Google Scholar PubMed
82. Tang, C.; Zhang, R.; Lu, W.; Wang, Z.; Liu, D.; Hao, S.; Du, G.; Asiri, A. M.; Sun, X. Energy-Saving Electrolytic Hydrogen Generation: Ni2P Nanoarray as a High-Performance Non-Noble-Metal Electrocatalyst. Angew. Chemie Int. Ed. 2017, 56, 842–846. https://doi.org/10.1002/anie.201608899.Suche in Google Scholar PubMed
83. Liu, X.; He, J.; Zhao, S.; Liu, Y.; Zhao, Z.; Luo, J.; Hu, G.; Sun, X.; Ding, Y. Self-Powered H2 Production with Bifunctional Hydrazine as Sole Consumable. Nat. Commun. 2018, 9, 4365; https://doi.org/10.1038/s41467-018-06815-9.Suche in Google Scholar PubMed PubMed Central
84. Du, M.; Sun, H.; Li, J.; Ye, X.; Yue, F.; Yang, J.; Liu, Y.; Guo, F. Integrative Ni@Pd-Ni Alloy Nanowire Array Electrocatalysts Boost Hydrazine Oxidation Kinetics. ChemElectroChem 2019, 6, 5581–5587. https://doi.org/10.1002/celc.201901303.Suche in Google Scholar
85. Ma, X.; Wang, J.; Liu, D.; Kong, R.; Hao, S.; Du, G.; Asiri, A. M.; Sun, X. Hydrazine-Assisted Electrolytic Hydrogen Production: CoS2 Nanoarray as a Superior Bifunctional Electrocatalyst. New J. Chem. 2017, 41, 4754–4757; https://doi.org/10.1039/C7NJ00326A.Suche in Google Scholar
86. Gao, L.; Xie, J.; Liu, S.; Lou, S.; Wei, Z.; Zhu, X.; Tang, B. Crystalline Cobalt/Amorphous LaCoOx Hybrid Nanoparticles Embedded in Porous Nitrogen-Doped Carbon as Efficient Electrocatalysts for Hydrazine-Assisted Hydrogen Production. ACS Appl. Mater. Interfaces 2020, 12, 24701–24709; https://doi.org/10.1021/acsami.0c02124.Suche in Google Scholar PubMed
87. Li, Y.; Zhang, J.; Liu, Y.; Qian, Q.; Li, Z.; Zhu, Y.; Zhang, G. Partially Exposed RuP2 Surface in Hybrid Structure Endows its Bifunctionality for Hydrazine Oxidation and Hydrogen Evolution Catalysis. Sci. Adv. 2020, 6, eabb4197; https://doi.org/10.1126/sciadv.abb4197.Suche in Google Scholar PubMed PubMed Central
88. Feng, G.; Kuang, Y.; Li, P.; Han, N.; Sun, M.; Zhang, G.; Sun, X. Single Crystalline Ultrathin Nickel–Cobalt Alloy Nanosheets Array for Direct Hydrazine Fuel Cells. Adv. Sci. 2017, 4, 1600179. https://doi.org/10.1002/advs.201600179.Suche in Google Scholar PubMed PubMed Central
89. Peng, W.; Xiao, L.; Huang, B.; Zhuang, L.; Lu, J. Inhibition Effect of Surface Oxygenated Species on Ammonia Oxidation Reaction. J. Phys. Chem. 2011, 115, 23050–23056; https://doi.org/10.1021/jp2081148.Suche in Google Scholar
90. Hu, S.; Tan, Y.; Feng, C.; Wu, H.; Zhang, J.; Mei, H. Synthesis of N Doped NiZnCu-Layered Double Hydroxides with Reduced Graphene Oxide on Nickel Foam as Versatile Electrocatalysts for Hydrogen Production in Hybrid-Water Electrolysis. J. Power Sources 2020, 453, 227872. https://doi.org/10.1016/j.jpowsour.2020.227872.Suche in Google Scholar
91. Huang, J.; Cai, J.; Wang, J. Nanostructured Wire-In-Plate Electrocatalyst for High-Durability Production of Hydrogen and Nitrogen from Alkaline Ammonia Solution. ACS Appl. Energy Mater. 2020, 3, 4108–4113; https://doi.org/10.1021/acsaem.9b02545.Suche in Google Scholar
92. Xu, W.; Du, D.; Lan, R.; Humphreys, J.; Miller, D. N.; Walker, M.; Wu, Z.; Irvine, J. T. S.; Tao, S. Electrodeposited NiCu Bimetal on Carbon Paper as Stable Non-Noble Anode for Efficient Electrooxidation of Ammonia. Appl. Catal. B Environ. 2018, 237, 1101–1109. https://doi.org/10.1016/j.apcatb.2016.11.003.Suche in Google Scholar
93. Dong, B.-X.; Tian, H.; Wu, Y.-C.; Bu, F.-Y.; Liu, W.-L.; Teng, Y.-L.; Diao, G.-W. Improved Electrolysis of Liquid Ammonia for Hydrogen Generation via Ammonium Salt Electrolyte and Pt/Rh/Ir Electrocatalysts. Int. J. Hydrogen Energy 2016, 41, 14507–14518. https://doi.org/10.1016/j.ijhydene.2016.06.212.Suche in Google Scholar
94. Sun, X.; Ding, R. Recent Progress with Electrocatalysts for Urea Electrolysis in Alkaline Media for Energy-Saving Hydrogen Production. Catal. Sci. Technol. 2020, 10, 1567–1581; https://doi.org/10.1039/C9CY02618E.Suche in Google Scholar
95. Chen, S.; Duan, J.; Vasileff, A.; Qiao, S. Z. Size Fractionation of Two-Dimensional Sub-nanometer Thin Manganese Dioxide Crystals Towards Superior Urea Electrocatalytic Conversion. Angew. Chemie Int. Ed. 2016, 55, 3804–3808. https://doi.org/10.1002/anie.201600387.Suche in Google Scholar PubMed
96. Zhang, J.-Y.; Tian, X.; He, T.; Zaman, S.; Miao, M.; Yan, Y.; Qi, K.; Dong, Z.; Liu, H.; Xia, B. Y. In Situ Formation of Ni3Se4 Nanorod Arrays as Versatile Electrocatalysts for Electrochemical Oxidation Reactions in Hybrid Water Electrolysis. J. Mater. Chem. A 2018, 6, 15653–15658; https://doi.org/10.1039/C8TA06361C.Suche in Google Scholar
97. Zheng, J.; Wu, K.; Lyu, C.; Pan, X.; Zhang, X.; Zhu, Y.; Wang, A.; Lau, W.-M.; Wang, N. Electrocatalyst of Two-Dimensional CoP Nanosheets Embedded by Carbon Nanoparticles for Hydrogen Generation and Urea Oxidation in Alkaline Solution. Appl. Surf. Sci. 2020, 506, 144977. https://doi.org/10.1016/j.apsusc.2019.144977.Suche in Google Scholar
98. Zhang, B.; Lui, Y. H.; Gaur, A. P. S.; Chen, B.; Tang, X.; Qi, Z.; Hu, S. Hierarchical FeNiP@Ultrathin Carbon Nanoflakes as Alkaline Oxygen Evolution and Acidic Hydrogen Evolution Catalyst for Efficient Water Electrolysis and Organic Decomposition. ACS Appl. Mater. Interfaces 2018, 10, 8739–8748; https://doi.org/10.1021/acsami.8b00069.Suche in Google Scholar PubMed
99. Li, Z.; Li, X.; Zhou, H.; Xu, Y.; Xu, S. M.; Ren, Y.; Yan, Y.; Yang, J.; Ji, K.; Li, L.; Xu, M.; Shao, M.; Kong, X.; Sun, X.; Duan, H. Electrocatalytic Synthesis of Adipic Acid Coupled with H2 Production Enhanced by a Ligand Modification Strategy. Nat. Commun. 2022, 13, 1–12; https://doi.org/10.1038/s41467-022-32769-0.Suche in Google Scholar PubMed PubMed Central
100. Sauter, W.; Bergmann, O. L.; Schröder, U. Hydroxyacetone: A Glycerol-Based Platform for Electrocatalytic Hydrogenation and Hydrodeoxygenation Processes. ChemSusChem 2017, 10, 3105–3110. https://doi.org/10.1002/cssc.201700996.Suche in Google Scholar PubMed
101. Fan, L.; Liu, B.; Liu, X.; Senthilkumar, N.; Wang, G.; Wen, Z. Recent Progress in Electrocatalytic Glycerol Oxidation. Energy Technol. 2021, 9, 2000804. https://doi.org/10.1002/ente.202000804.Suche in Google Scholar
102. Gao, T.; Tang, X.; Li, X.; Lan, H.; Yu, S.; Wu, S.; Yue, Q.; Xiao, D. Surface Reconstructing Hierarchical Structures as Robust Sulfion Oxidation Catalysts to Produce Hydrogen with Ultralow Energy Consumption. Inorg. Chem. Front. 2023, 10, 1447–1456; https://doi.org/10.1039/D2QI02629E.Suche in Google Scholar
103. Zhang, S.; Zhou, Q.; Shen, Z.; Jin, X.; Zhang, Y.; Shi, M.; Zhou, J.; Liu, J.; Lu, Z.; Zhou, Y.-N.; Zhang, H. Sulfophobic and Vacancy Design Enables Self-Cleaning Electrodes for Efficient Desulfurization and Concurrent Hydrogen Evolution with Low Energy Consumption. Adv. Funct. Mater. 2021, 31, 2101922. https://doi.org/10.1002/adfm.202101922.Suche in Google Scholar
104. Zhang, M.; Guan, J.; Tu, Y.; Chen, S.; Wang, Y.; Wang, S.; Yu, L.; Ma, C.; Deng, D.; Bao, X. Highly Efficient H2 Production from H2S via a Robust Graphene-Encapsulated Metal Catalyst. Energy Environ. Sci. 2020, 13, 119–126; https://doi.org/10.1039/C9EE03231B.Suche in Google Scholar
105. Zhang, L.; Wang, Z.; Qiu, J. Energy-Saving Hydrogen Production by Seawater Electrolysis Coupling Sulfion Degradation. Adv. Mater. 2022, 34, 2109321. https://doi.org/10.1002/adma.202109321.Suche in Google Scholar PubMed
106. Yi, L.; Ji, Y.; Shao, P.; Chen, J.; Li, J.; Li, H.; Chen, K.; Peng, X.; Wen, Z. Scalable Synthesis of Tungsten Disulfide Nanosheets for Alkali-Acid Electrocatalytic Sulfion Recycling and H2 Generation. Angew. Chemie Int. Ed. 2021, 60, 21550–21557. https://doi.org/10.1002/anie.202108992.Suche in Google Scholar PubMed
107. Marshall, A. T.; Haverkamp, R. G. Production of Hydrogen by the Electrochemical Reforming of Glycerol–Water Solutions in a PEM Electrolysis Cell. Int. J. Hydrogen Energy 2008, 33, 4649–4654. https://doi.org/10.1016/j.ijhydene.2008.05.029.Suche in Google Scholar
108. Li, Y.; Wei, X.; Chen, L.; Shi, J.; He, M. Nickel-Molybdenum Nitride Nanoplate Electrocatalysts for Concurrent Electrolytic Hydrogen and Formate Productions. Nat. Commun. 2019, 10, 1–12; https://doi.org/10.1038/s41467-019-13375-z.Suche in Google Scholar PubMed PubMed Central
109. Zhu, Y.; Qian, Q.; Chen, Y.; He, X.; Shi, X.; Wang, W.; Li, Z.; Feng, Y.; Zhang, G.; Cheng, F. Biphasic Transition Metal Nitride Electrode Promotes Nucleophile Oxidation Reaction for Practicable Hybrid Water Electrocatalysis. Adv. Funct. Mater. 2023, 33, 2300547. https://doi.org/10.1002/adfm.202300547.Suche in Google Scholar
110. Tran, G.-S.; Vo, T.-G.; Chiang, C.-Y. Operando Revealing the Crystal Phase Transformation and Electrocatalytic Activity Correlation of MnO2 toward Glycerol Electrooxidation. ACS Appl. Mater. Interfaces 2023, 15, 22662–22671; https://doi.org/10.1021/acsami.3c00857.Suche in Google Scholar PubMed
111. Zhou, L.; Liang, R.; Ma, Z.; Wu, T.; Wu, Y. Conversion of Cellulose to HMF in Ionic Liquid Catalyzed by Bifunctional Ionic Liquids. Bioresour. Technol. 2013, 129, 450–455. https://doi.org/10.1016/j.biortech.2012.11.015.Suche in Google Scholar PubMed
112. Tong, X.; Ma, Y.; Li, Y. Biomass into Chemicals: Conversion of Sugars to Furan Derivatives by Catalytic Processes. Appl. Catal. Gen. 2010, 385, 1–13. https://doi.org/10.1016/j.apcata.2010.06.049.Suche in Google Scholar
113. Grabowski, G.; Lewkowski, J.; Skowroński, R. The Electrochemical Oxidation of 5-Hydroxymethylfurfural with the Nickel Oxide/Hydroxide Electrode. Electrochim. Acta 1991, 36, 1995. https://doi.org/10.1016/0013-4686(91)85084-K.Suche in Google Scholar
114. Nichols, E. M.; Gallagher, J. J.; Liu, C.; Su, Y.; Resasco, J.; Yu, Y.; Sun, Y.; Yang, P.; Chang, M. C. Y.; Chang, C. J. Hybrid Bioinorganic Approach to Solar-Tochemical Conversion. Proc. Natl. Acad. Sci. U. S. A. 2015, 112, 11461–11466; https://doi.org/10.1073/pnas.1508075112.Suche in Google Scholar PubMed PubMed Central
115. Liu, C.; Colón, B. C.; Ziesack, M.; Silver, P. A.; Daniel, G. Nocera Water Splitting–Biosynthetic System with CO2 Reduction Efficiencies Exceeding Photosynthesis. Science (80-. ) 2016, 352, 1210–1213; https://doi.org/10.1126/science.aaf5039.Suche in Google Scholar PubMed
116. Liu, C.; Sakimoto, K. K.; Colón, B. C.; Silver, P. A.; Nocera, D. G. Ambient Nitrogen Reduction Cycle Using a Hybrid Inorganic-Biological System. Proc. Natl. Acad. Sci. U. S. A. 2017, 114, 6450–6455; https://doi.org/10.1073/pnas.1706371114.Suche in Google Scholar PubMed PubMed Central
117. Yang, E.; Omar Mohamed, H.; Park, S. G.; Obaid, M.; Al-Qaradawi, S. Y.; Castaño, P.; Chon, K.; Chae, K. J. A Review on Self-Sustainable Microbial Electrolysis Cells for Electro-Biohydrogen Production via Coupling with Carbon-Neutral Renewable Energy Technologies. Bioresour. Technol. 2021, 320; https://doi.org/10.1016/j.biortech.2020.124363.Suche in Google Scholar PubMed
118. Zhang, Y.; Angelidaki, I. Microbial Electrolysis Cells Turning to Be Versatile Technology: Recent Advances and Future Challenges. Water Res. 2014, 56, 11–25; https://doi.org/10.1016/j.watres.2014.02.031.Suche in Google Scholar PubMed
119. Xiang, L. J.; Dai, L.; Guo, K. X.; Wen, Z. H.; Ci, S. Q.; Li, J. H. Microbial Electrolysis Cells for Hydrogen Production. Chinese J. Chem. Phys. 2020, 33, 263–284; https://doi.org/10.1063/1674-0068/cjcp2005075.Suche in Google Scholar
120. Oyiwona, G. E.; Ogbonna, J.; Anyanwu, C. U.; Ishizaki, S.; Kimura, Z.; Okabe, S. Oxidation of Glucose by Syntrophic Association between Geobacter and Hydrogenotrophic Methanogens in Microbial Fuel Cell. Biotechnol. Lett. 2017, 39, 253–259; https://doi.org/10.1007/s10529-016-2247-4.Suche in Google Scholar PubMed
121. Schröder, U. Anodic Electron Transfer Mechanisms in Microbial Fuel Cells and Their Energy Efficiency. Phys. Chem. Chem. Phys. 2007, 9, 2619–2629; https://doi.org/10.1039/B703627M.Suche in Google Scholar
122. Selembo, P. A.; Perez, J. M.; Lloyd, W. A.; Logan, B. E. High Hydrogen Production from Glycerol or Glucose by Electrohydrogenesis Using Microbial Electrolysis Cells. Int. J. Hydrogen Energy 2009, 34, 5373–5381. https://doi.org/10.1016/j.ijhydene.2009.05.002.Suche in Google Scholar
123. Parkhey, P.; Gupta, P. Improvisations in Structural Features of Microbial Electrolytic Cell and Process Parameters of Electrohydrogenesis for Efficient Biohydrogen Production: A Review. Renew. Sustain. Energy Rev. 2017, 69, 1085–1099; https://doi.org/10.1016/j.rser.2016.09.101.Suche in Google Scholar
124. Ivanov, I.; Ahn, Y.; Poirson, T.; Hickner, M. A.; Logan, B. E. Comparison of Cathode Catalyst Binders for the Hydrogen Evolution Reaction in Microbial Electrolysis Cells. Int. J. Hydrogen Energy 2017, 42, 15739–15744. https://doi.org/10.1016/j.ijhydene.2017.05.089.Suche in Google Scholar
125. Liu, W.; Wang, A.; Ren, N.; Zhao, X.; Liu, L.; Yu, Z.; Lee, D.-J. Electrochemically Assisted Biohydrogen Production from Acetate. Energy & Fuels 2008, 22, 159–163; https://doi.org/10.1021/ef700293e.Suche in Google Scholar
126. Liu, H.; Grot, S.; Logan, B. E. Electrochemically Assisted Microbial Production of Hydrogen from Acetate. Environ. Sci. Technol. 2005, 39, 4317–4320; https://doi.org/10.1021/es050244p.Suche in Google Scholar PubMed
127. Jeremiasse, A. W.; Hamelers, H. V. M.; Buisman, C. J. N. Microbial Electrolysis Cell with a Microbial Biocathode. Bioelectrochemistry 2010, 78, 39–43; https://doi.org/10.1016/j.bioelechem.2009.05.005.Suche in Google Scholar PubMed
128. Chae, K.-J.; Choi, M.-J.; Kim, K.-Y.; Ajayi, F. F.; Chang, I.-S.; Kim, I. S. A Solar-Powered Microbial Electrolysis Cell with a Platinum Catalyst-free Cathode to Produce Hydrogen. Environ. Sci. Technol. 2009, 43, 9525–9530; https://doi.org/10.1021/es9022317.Suche in Google Scholar PubMed
129. Rozendal, R. A.; Hamelers, H. V. M.; Molenkamp, R. J.; Buisman, C. J. N. Performance of Single Chamber Biocatalyzed Electrolysis with Different Types of Ion Exchange Membranes. Water Res. 2007, 41, 1984–1994; https://doi.org/10.1016/j.watres.2007.01.019.Suche in Google Scholar PubMed
130. Karthikeyan, R.; Cheng, K. Y.; Selvam, A.; Bose, A.; Wong, J. W. C. Bioelectrohydrogenesis and Inhibition of Methanogenic Activity in Microbial Electrolysis Cells - A Review. Biotechnol. Adv. 2017, 35, 758–771. https://doi.org/10.1016/j.biotechadv.2017.07.004.Suche in Google Scholar PubMed
131. Liang, D.-W.; Peng, S.-K.; Lu, S.-F.; Liu, Y.-Y.; Lan, F.; Xiang, Y. Enhancement of Hydrogen Production in a Single Chamber Microbial Electrolysis Cell through Anode Arrangement Optimization. Bioresour. Technol. 2011, 102, 10881–10885. https://doi.org/10.1016/j.biortech.2011.09.028.Suche in Google Scholar PubMed
132. Hu, H.; Fan, Y.; Liu, H. Hydrogen Production Using Single-Chamber Membrane-free Microbial Electrolysis Cells. Water Res. 2008, 42, 4172–4178. https://doi.org/10.1016/j.watres.2008.06.015.Suche in Google Scholar PubMed
133. Ruiz, Y.; Baeza, J. A.; Guisasola, A. Revealing the Proliferation of Hydrogen Scavengers in a Single-Chamber Microbial Electrolysis Cell Using Electron Balances. Int. J. Hydrogen Energy 2013, 38, 15917–15927. https://doi.org/10.1016/j.ijhydene.2013.10.034.Suche in Google Scholar
134. Call, D.; Logan, B. E. Hydrogen Production in a Single Chamber Microbial Electrolysis Cell Lacking a Membrane. Environ. Sci. Technol. 2008, 42, 3401–3406; https://doi.org/10.1021/es8001822.Suche in Google Scholar PubMed
135. Kadier, A.; Kalil, M. S.; Chandrasekhar, K.; Mohanakrishna, G.; Saratale, G. D.; Saratale, R. G.; Kumar, G.; Pugazhendhi, A.; Sivagurunathan, P. Surpassing the Current Limitations of High Purity H2 Production in Microbial Electrolysis Cell (MECs): Strategies for Inhibiting Growth of Methanogens. Bioelectrochemistry 2018, 119, 211–219. https://doi.org/10.1016/j.bioelechem.2017.09.014.Suche in Google Scholar PubMed
136. Guo, K.; Prévoteau, A.; Patil, S. A.; Rabaey, K. Engineering Electrodes for Microbial Electrocatalysis. Curr. Opin. Biotechnol. 2015, 33, 149–156. https://doi.org/10.1016/j.copbio.2015.02.014.Suche in Google Scholar PubMed
137. Escapa, A.; Mateos, R.; Martínez, E. J.; Blanes, J. Microbial Electrolysis Cells: An Emerging Technology for Wastewater Treatment and Energy Recovery. From Laboratory to Pilot Plant and beyond. Renew. Sustain. Energy Rev. 2016, 55, 942–956. https://doi.org/10.1016/j.rser.2015.11.029.Suche in Google Scholar
138. Zhou, M.; Chi, M.; Luo, J.; He, H.; Jin, T. An Overview of Electrode Materials in Microbial Fuel Cells. J. Power Sources 2011, 196, 4427–4435. https://doi.org/10.1016/j.jpowsour.2011.01.012.Suche in Google Scholar
139. Kadier, A.; Kalil, M. S.; Abdeshahian, P.; Chandrasekhar, K.; Mohamed, A.; Azman, N. F.; Logroño, W.; Simayi, Y.; Hamid, A. A. Recent Advances and Emerging Challenges in Microbial Electrolysis Cells (MECs) for Microbial Production of Hydrogen and Value-Added Chemicals. Renew. Sustain. Energy Rev. 2016, 61, 501–525; https://doi.org/10.1016/j.rser.2016.04.017.Suche in Google Scholar
140. Liu, H.; Hu, H.; Chignell, J.; Fan, Y. Microbial Electrolysis: Novel Technology for Hydrogen Production from Biomass. Biofuels 2010, 1, 129–142; https://doi.org/10.4155/bfs.09.9.Suche in Google Scholar
141. Sun, M.; Zhang, F.; Tong, Z.-H.; Sheng, G.-P.; Chen, Y.-Z.; Zhao, Y.; Chen, Y.-P.; Zhou, S.-Y.; Liu, G.; Tian, Y.-C.; Yu, H. Q. A Gold-Sputtered Carbon Paper as an Anode for Improved Electricity Generation from a Microbial Fuel Cell Inoculated with Shewanella Oneidensis MR-1. Biosens. Bioelectron. 2010, 26, 338–343. https://doi.org/10.1016/j.bios.2010.08.010.Suche in Google Scholar PubMed
142. Lv, Z.; Xie, D.; Yue, X.; Feng, C.; Wei, C. Ruthenium Oxide-Coated Carbon Felt Electrode: A Highly Active Anode for Microbial Fuel Cell Applications. J. Power Sources 2012, 210, 26–31. https://doi.org/10.1016/j.jpowsour.2012.02.109.Suche in Google Scholar
143. Qiao, Y.; Li, C. M.; Bao, S.-J.; Bao, Q.-L. Carbon Nanotube/Polyaniline Composite as Anode Material for Microbial Fuel Cells. J. Power Sources 2007, 170, 79–84. https://doi.org/10.1016/j.jpowsour.2007.03.048.Suche in Google Scholar
144. Zhang, Y.; Liu, L.; Van der Bruggen, B.; Yang, F. Nanocarbon Based Composite Electrodes and Their Application in Microbial Fuel Cells. J. Mater. Chem. A 2017, 5, 12673–12698; https://doi.org/10.1039/C7TA01511A.Suche in Google Scholar
145. Cui, H.-F.; Du, L.; Guo, P.-B.; Zhu, B.; Luong, J. H. T. Controlled Modification of Carbon Nanotubes and Polyaniline on Macroporous Graphite Felt for High-Performance Microbial Fuel Cell Anode. J. Power Sources 2015, 283, 46–53. https://doi.org/10.1016/j.jpowsour.2015.02.088.Suche in Google Scholar
146. Chen, S.; He, G.; Hu, X.; Xie, M.; Wang, S.; Zeng, D.; Hou, H.; Schröder, U. A Three-Dimensionally Ordered Macroporous Carbon Derived from a Natural Resource as Anode for Microbial Bioelectrochemical Systems. ChemSusChem 2012, 5, 1059–1063. https://doi.org/10.1002/cssc.201100783.Suche in Google Scholar PubMed
147. Alatraktchi, F. A.; Zhang, Y.; Angelidaki, I. Nanomodification of the Electrodes in Microbial Fuel Cell: Impact of Nanoparticle Density on Electricity Production and Microbial Community. Appl. Energy 2014, 116, 216–222. https://doi.org/10.1016/j.apenergy.2013.11.058.Suche in Google Scholar
148. Hindatu, Y.; Annuar, M. S. M.; Gumel, A. M. Mini-Review: Anode Modification for Improved Performance of Microbial Fuel Cell. Renew. Sustain. Energy Rev. 2017, 73, 236–248. https://doi.org/10.1016/j.rser.2017.01.138.Suche in Google Scholar
149. Wang, Q.; Huang, L.; Yu, H.; Quan, X.; Li, Y.; Fan, G.; Li, L. Assessment of Five Different Cathode Materials for Co(II) Reduction with Simultaneous Hydrogen Evolution in Microbial Electrolysis Cells. Int. J. Hydrogen Energy 2015, 40, 184–196. https://doi.org/10.1016/j.ijhydene.2014.11.014.Suche in Google Scholar
150. Gnana kumar, G.; Kirubaharan, C. J.; Udhayakumar, S.; Karthikeyan, C.; Nahm, K. S. Conductive Polymer/Graphene Supported Platinum Nanoparticles as Anode Catalysts for the Extended Power Generation of Microbial Fuel Cells. Ind. Eng. Chem. Res. 2014, 53, 16883–16893; https://doi.org/10.1021/ie502399y.Suche in Google Scholar
151. Kang, Y. L.; Ibrahim, S.; Pichiah, S. Synergetic Effect of Conductive Polymer Poly(3,4-Ethylenedioxythiophene) with Different Structural Configuration of Anode for Microbial Fuel Cell Application. Bioresour. Technol. 2015, 189, 364–369. https://doi.org/10.1016/j.biortech.2015.04.044.Suche in Google Scholar PubMed
152. Luckarift, H. R.; Sizemore, S. R.; Farrington, K. E.; Roy, J.; Lau, C.; Atanassov, P. B.; Johnson, G. R. Facile Fabrication of Scalable, Hierarchically Structured Polymer/Carbon Architectures for Bioelectrodes. ACS Appl. Mater. Interfaces 2012, 4, 2082–2087; https://doi.org/10.1021/am300048v.Suche in Google Scholar PubMed
153. Kadier, A.; Simayi, Y.; Kalil, M. S.; Abdeshahian, P.; Hamid, A. A. A Review of the Substrates Used in Microbial Electrolysis Cells (MECs) for Producing Sustainable and Clean Hydrogen Gas. Renew. Energy 2014, 71, 466–472; https://doi.org/10.1016/j.renene.2014.05.052.Suche in Google Scholar
154. Chaurasia, A. K.; Goyal, H.; Mondal, P. Hydrogen Gas Production with Ni, Ni–Co and Ni–Co–P Electrodeposits as Potential Cathode Catalyst by Microbial Electrolysis Cells. Int. J. Hydrogen Energy 2020, 45, 18250–18265. https://doi.org/10.1016/j.ijhydene.2019.07.175.Suche in Google Scholar
155. Mitov, M.; Chorbadzhiyska, E.; Nalbandian, L.; Hubenova, Y. Nickel-Based Electrodeposits as Potential Cathode Catalysts for Hydrogen Production by Microbial Electrolysis. J. Power Sources 2017, 356, 467–472. https://doi.org/10.1016/j.jpowsour.2017.02.066.Suche in Google Scholar
156. Kadier, A.; Simayi, Y.; Chandrasekhar, K.; Ismail, M.; Kalil, M. S. Hydrogen Gas Production with an Electroformed Ni Mesh Cathode Catalysts in a Single-Chamber Microbial Electrolysis Cell (MEC). Int. J. Hydrogen Energy 2015, 40, 14095–14103. https://doi.org/10.1016/j.ijhydene.2015.08.095.Suche in Google Scholar
157. Torres, C. I.; Marcus, A. K.; Lee, H. S.; Parameswaran, P.; Krajmalnik-Brown, R.; Rittmann, B. E. A Kinetic Perspective on Extracellular Electron Transfer by Anode-Respiring Bacteria. FEMS Microbiol. Rev. 2010, 34, 3–17; https://doi.org/10.1111/j.1574-6976.2009.00191.x.Suche in Google Scholar PubMed
158. Hernandez, M. E.; Kappler, A.; Newman, D. K. Phenazines and Other Redox-Active Antibiotics Promote Microbial Mineral Reduction. Appl. Environ. Microbiol. 2004, 70, 921–928; https://doi.org/10.1128/AEM.70.2.921-928.2004.Suche in Google Scholar PubMed PubMed Central
159. Newman, D. K.; Kolter, R. A Role for Excreted Quinones in Extracellular Electron Transfer. Nature 2000, 405, 94–97; https://doi.org/10.1038/35011098.Suche in Google Scholar PubMed
160. Von Canstein, H.; Ogawa, J.; Shimizu, S.; Lloyd, J. R. Secretion of Flavins by Shewanella Species and Their Role in Extracellular Electron Transfer. Appl. Environ. Microbiol. 2008, 74, 615–623; https://doi.org/10.1128/AEM.01387-07.Suche in Google Scholar PubMed PubMed Central
161. Reguera, G.; McCarthy, K. D.; Mehta, T.; Nicoll, J. S.; Tuominen, M. T.; Lovley, D. R. Extracellular Electron Transfer via Microbial Nanowires. Nature 2005, 435, 1098–1101; https://doi.org/10.1038/nature03661.Suche in Google Scholar PubMed
162. Kiely, P. D.; Rader, G.; Regan, J. M.; Logan, B. E. Long-Term Cathode Performance and the Microbial Communities that Develop in Microbial Fuel Cells Fed Different Fermentation Endproducts. Bioresour. Technol. 2011, 102, 361–366. https://doi.org/10.1016/j.biortech.2010.05.017.Suche in Google Scholar PubMed
163. Kiely, P. D.; Cusick, R.; Call, D. F.; Selembo, P. A.; Regan, J. M.; Logan, B. E. Anode Microbial Communities Produced by Changing from Microbial Fuel Cell to Microbial Electrolysis Cell Operation Using Two Different Wastewaters. Bioresour. Technol. 2011, 102, 388–394. https://doi.org/10.1016/j.biortech.2010.05.019.Suche in Google Scholar PubMed
164. Wagner, R. C.; Regan, J. M.; Oh, S.-E.; Zuo, Y.; Logan, B. E. Hydrogen and Methane Production from Swine Wastewater Using Microbial Electrolysis Cells. Water Res. 2009, 43, 1480–1488. https://doi.org/10.1016/j.watres.2008.12.037.Suche in Google Scholar PubMed
165. Cusick, R. D.; Kiely, P. D.; Logan, B. E. A Monetary Comparison of Energy Recovered from Microbial Fuel Cells and Microbial Electrolysis Cells Fed Winery or Domestic Wastewaters. Int. J. Hydrogen Energy 2010, 35, 8855–8861. https://doi.org/10.1016/j.ijhydene.2010.06.077.Suche in Google Scholar
166. Ditzig, J.; Liu, H.; Logan, B. E. Production of Hydrogen from Domestic Wastewater Using a Bioelectrochemically Assisted Microbial Reactor (BEAMR). Int. J. Hydrogen Energy 2007, 32, 2296–2304. https://doi.org/10.1016/j.ijhydene.2007.02.035.Suche in Google Scholar
167. Tenca, A.; Cusick, R. D.; Schievano, A.; Oberti, R.; Logan, B. E. Evaluation of Low Cost Cathode Materials for Treatment of Industrial and Food Processing Wastewater Using Microbial Electrolysis Cells. Int. J. Hydrogen Energy 2013, 38, 1859–1865. https://doi.org/10.1016/j.ijhydene.2012.11.103.Suche in Google Scholar
168. Montpart, N.; Rago, L.; Baeza, J. A.; Guisasola, A. Hydrogen Production in Single Chamber Microbial Electrolysis Cells with Different Complex Substrates. Water Res. 2015, 68, 601–615. https://doi.org/10.1016/j.watres.2014.10.026.Suche in Google Scholar PubMed
169. Cebecioglu, R.; Akagunduz, D.; Catal, T. Hydrogen Production in Single-Chamber Microbial Electrolysis Cells Using Ponceau S Dye. 3 Biotech 2021, 11, 27; https://doi.org/10.1007/s13205-020-02563-0.Suche in Google Scholar PubMed PubMed Central
170. Dange, P.; Pandit, S.; Jadhav, D.; Shanmugam, P.; Gupta, P. K.; Kumar, S.; Kumar, M.; Yang, Y. H.; Bhatia, S. K. Recent Developments in Microbial Electrolysis Cell-Based Biohydrogen Production Utilizing Wastewater as a Feedstock. Sustain 2021, 13, 1–37; https://doi.org/10.3390/su13168796.Suche in Google Scholar
171. Aiken, D. C.; Curtis, T. P.; Heidrich, E. S. Avenues to the Financial Viability of Microbial Electrolysis Cells [MEC] for Domestic Wastewater Treatment and Hydrogen Production. Int. J. Hydrogen Energy 2019, 44, 2426–2434; https://doi.org/10.1016/j.ijhydene.2018.12.029.Suche in Google Scholar
172. Jeremiasse, A. W.; Hamelers, H. V. M.; Croese, E.; Buisman, C. J. N. Acetate Enhances Startup of a H2-Producing Microbial Biocathode. Biotechnol. Bioeng. 2012, 109, 657–664. https://doi.org/10.1002/bit.24338.Suche in Google Scholar PubMed
173. Carmona-Martínez, A. A.; Trably, E.; Milferstedt, K.; Lacroix, R.; Etcheverry, L.; Bernet, N. Long-Term Continuous Production of H2 in a Microbial Electrolysis Cell (MEC) Treating Saline Wastewater. Water Res. 2015, 81, 149–156. https://doi.org/10.1016/j.watres.2015.05.041.Suche in Google Scholar PubMed
174. Cheng, C.-L.; Lo, Y.-C.; Lee, K.-S.; Lee, D.-J.; Lin, C.-Y.; Chang, J.-S. Biohydrogen Production from Lignocellulosic Feedstock. Bioresour. Technol. 2011, 102, 8514–8523. https://doi.org/10.1016/j.biortech.2011.04.059.Suche in Google Scholar PubMed
175. Dhar, B. R.; Elbeshbishy, E.; Hafez, H.; Lee, H.-S. Hydrogen Production from Sugar Beet Juice Using an Integrated Biohydrogen Process of Dark Fermentation and Microbial Electrolysis Cell. Bioresour. Technol. 2015, 198, 223–230. https://doi.org/10.1016/j.biortech.2015.08.048.Suche in Google Scholar PubMed
176. Ndayisenga, F.; Yu, Z.; Zheng, J.; Wang, B.; Liang, H.; Phulpoto, I. A.; Habiyakare, T.; Zhou, D. Microbial Electrohydrogenesis Cell and Dark Fermentation Integrated System Enhances Biohydrogen Production from Lignocellulosic Agricultural Wastes: Substrate Pretreatment towards Optimization. Renew. Sustain. Energy Rev. 2021, 145, 111078. https://doi.org/10.1016/j.rser.2021.111078.Suche in Google Scholar
177. Liu, W.; Huang, S.; Zhou, A.; Zhou, G.; Ren, N.; Wang, A.; Zhuang, G. Hydrogen Generation in Microbial Electrolysis Cell Feeding with Fermentation Liquid of Waste Activated Sludge. Int. J. Hydrogen Energy 2012, 37, 13859–13864. https://doi.org/10.1016/j.ijhydene.2012.04.090.Suche in Google Scholar
178. Heidrich, E. S.; Edwards, S. R.; Dolfing, J.; Cotterill, S. E.; Curtis, T. P. Performance of a Pilot Scale Microbial Electrolysis Cell Fed on Domestic Wastewater at Ambient Temperatures for a 12month Period. Bioresour. Technol. 2014, 173, 87–95. https://doi.org/10.1016/j.biortech.2014.09.083.Suche in Google Scholar PubMed
179. Cotterill, S. E.; Dolfing, J.; Curtis, T. P.; Heidrich, E. S. Community Assembly in Wastewater-Fed Pilot-Scale Microbial Electrolysis Cells. Front. Energy Res. 2018, 6, 1–12; https://doi.org/10.3389/fenrg.2018.00098.Suche in Google Scholar
180. Savla, N.; Suman; Pandit, S.; Verma, J. P.; Awasthi, A. K.; Sana, S. S.; Prasad, R. Techno-Economical Evaluation and Life Cycle Assessment of Microbial Electrochemical Systems: A Review. Curr. Res. Green Sustain. Chem. 2021, 4, 100111. https://doi.org/10.1016/j.crgsc.2021.100111.Suche in Google Scholar
181. Mohd-Yusof, N. S.; Babgi, B.; Aksu, M.; Madhavan, J.; Ashokkumar, M. Physical and Chemical Effects of Acoustic Cavitation in Selected Ultrasonic Cleaning Applications. Ultrason. Sonochem. 2016, 29, 568–576; https://doi.org/10.1016/j.ultsonch.2015.06.013.Suche in Google Scholar PubMed
182. Islam, H.; Burheim, O. S.; Pollet, B. G. Sonochemical and Sonoelectrochemical Production of Hydrogen. Ultrason. Sonochem. 2019, 51, 533–555; https://doi.org/10.1016/j.ultsonch.2018.08.024.Suche in Google Scholar PubMed
183. Li, S.De; Wang, C. C.; Chen, C. Y. Water Electrolysis in the Presence of an Ultrasonic Field. Electrochim. Acta 2009, 54, 3877–3883; https://doi.org/10.1016/j.electacta.2009.01.087.Suche in Google Scholar
184. Hung, C. Y.; Li, S.De; Wang, C. C.; Chen, C. Y. Influences of a Bipolar Membrane and an Ultrasonic Field on Alkaline Water Electrolysis. J. Memb. Sci. 2012, 389, 197–204; https://doi.org/10.1016/j.memsci.2011.10.050.Suche in Google Scholar
185. Lin, M. Y.; Hourng, L. W. Ultrasonic Wave Field Effects on Hydrogen Production by Water Electrolysis. J. Chinese Inst. Eng. Trans. Chinese Inst. Eng. A 2014, 37, 1080–1089; https://doi.org/10.1080/02533839.2014.912787.Suche in Google Scholar
186. Merabet, N. H.; Kerboua, K. Membrane Free Alkaline Sono-Electrolysis for Hydrogen Production: An Experimental Approach. Int. J. Hydrogen Energy 2024, 49, 734–753; https://doi.org/10.1016/j.ijhydene.2023.09.061.Suche in Google Scholar
187. Merabet, N. H.; Kerboua, K. Green Hydrogen from Sono-Electrolysis: A Coupled Numerical and Experimental Study of the Ultrasound Assisted Membraneless Electrolysis of Water Supplied by PV. Fuel 2024, 356, 129625; https://doi.org/10.1016/j.fuel.2023.129625.Suche in Google Scholar
188. Pétrier, C. The Use of Power Ultrasound for Water Treatment. In Power ultrasonics: applications of high-intensity ultrasound; Gallego-Juarez, JA; Graff, K., Eds.; Elsevier: The Netherlands, 2015; pp 939–963.10.1016/B978-1-78242-028-6.00031-4Suche in Google Scholar
189. Merouani, S.; Hamdaoui, O. Sonochemical Treatment of Textile Wastewater. In Water Pollution and Remediation: Photocatalysis; Inamuddin, M.P.; Asiri, A., Eds. Springer-Nature: Switzerland, 2021.10.1007/978-3-030-54723-3_5Suche in Google Scholar
190. Koza, J. A.; Mühlenhoff, S.; Zabiński, P.; Nikrityuk, P. A.; Eckert, K.; Uhlemann, M.; Gebert, A.; Weier, T.; Schultz, L.; Odenbach, S. Hydrogen Evolution under the Influence of a Magnetic Field. Electrochim. Acta 2011, 56, 2665–2675; https://doi.org/10.1016/j.electacta.2010.12.031.Suche in Google Scholar
191. Matsushima, H.; Iida, T.; Fukunaka, Y. Gas Bubble Evolution on Transparent Electrode during Water Electrolysis in a Magnetic Field. Electrochim. Acta 2013, 100, 261–264; https://doi.org/10.1016/j.electacta.2012.05.082.Suche in Google Scholar
192. Kaya, M. F.; Demir, N.; Albawabiji, M. S.; Taş, M. Investigation of Alkaline Water Electrolysis Performance for Different Cost Effective Electrodes under Magnetic Field. Int. J. Hydrogen Energy 2017, 42, 17583–17592; https://doi.org/10.1016/j.ijhydene.2017.02.039.Suche in Google Scholar
193. Lin, M. Y.; Hourng, L. W.; Wu, C. H. The Effectiveness of a Magnetic Field in Increasing Hydrogen Production by Water Electrolysis. Energy Sources, Part A Recover. Util. Environ. Eff. 2017, 39, 140–147; https://doi.org/10.1080/15567036.2016.1176090.Suche in Google Scholar
194. Lin, M. Y.; Hourng, L. W.; Hsu, J. S. The Effects of Magnetic Field on the Hydrogen Production by Multielectrode Water Electrolysis. Energy Sources, Part A Recover. Util. Environ. Eff. 2017, 39, 352–357; https://doi.org/10.1080/15567036.2016.1217289.Suche in Google Scholar
195. Burton, N. A.; Padilla, R. V.; Rose, A.; Habibullah, H. Increasing the Efficiency of Hydrogen Production from Solar Powered Water Electrolysis. Renew. Sustain. Energy Rev. 2021, 135, 110255; https://doi.org/10.1016/j.rser.2020.110255.Suche in Google Scholar
196. Wang, M.; Wang, Z.; Gong, X.; Guo, Z. The Intensification Technologies to Water Electrolysis for Hydrogen Production - A Review. Renew. Sustain. Energy Rev. 2014, 29, 573–588; https://doi.org/10.1016/j.rser.2013.08.090.Suche in Google Scholar
197. Iida, T.; Matsushima, H.; Fukunaka, Y. Water Electrolysis under a Magnetic Field. J. Electrochem. Soc. 2007, 154, E112; https://doi.org/10.1149/1.2742807.Suche in Google Scholar
198. Lin, M. Y.; Hourng, L. W.; Kuo, C. W. The Effect of Magnetic Force on Hydrogen Production Efficiency in Water Electrolysis. Int. J. Hydrogen Energy 2012, 37, 1311–1320; https://doi.org/10.1016/j.ijhydene.2011.10.024.Suche in Google Scholar
199. Matsushima, H.; Kiuchi, D.; Fukunaka, Y. Measurement of Dissolved Hydrogen Supersaturation during Water Electrolysis in a Magnetic Field. Electrochim. Acta 2009, 54, 5858–5862; https://doi.org/10.1016/j.electacta.2009.05.044.Suche in Google Scholar
200. Matsushima, H.; Iida, T.; Fukunaka, Y. Observation of Bubble Layer Formed on Hydrogen and Oxygen Gas-Evolving Electrode in a Magnetic Field. J. Solid State Electrochem. 2012, 16, 617–623; https://doi.org/10.1007/s10008-011-1392-x.Suche in Google Scholar
201. Wang, Y.; Zhang, B.; Gong, Z.; Gao, K.; Ou, Y.; Zhang, J. The Effect of a Static Magnetic Field on the Hydrogen Bonding in Water Using Frictional Experiments. J. Mol. Struct. 2013, 1052, 102–104; https://doi.org/10.1016/j.molstruc.2013.08.021.Suche in Google Scholar
202. Wang, M.; Wang, Z.; Guo, Z. Understanding of the Intensified Effect of Super Gravity on Hydrogen Evolution Reaction. Int. J. Hydrogen Energy 2009, 34, 5311–5317; https://doi.org/10.1016/j.ijhydene.2009.05.043.Suche in Google Scholar
203. Lao, L.; Ramshaw, C.; Yeung, H. Process Intensification: Water Electrolysis in a Centrifugal Acceleration Field. J. Appl. Electrochem. 2011, 41, 645–656; https://doi.org/10.1007/s10800-011-0275-2.Suche in Google Scholar
204. Wang, M.; Wang, Z.; Guo, Z. Water Electrolysis Enhanced by Super Gravity Field for Hydrogen Production. Int. J. Hydrogen Energy 2010, 35, 3198–3205; https://doi.org/10.1016/j.ijhydene.2010.01.128.Suche in Google Scholar
205. Cheng, H.; Scott, K.; Ramshaw, C. Intensification of Water Electrolysis in a Centrifugal Field. J. Electrochem. Soc. 2002, 149, D172; https://doi.org/10.1149/1.1512916.Suche in Google Scholar
206. Tong, W.; Forster, M.; Dionigi, F.; Dresp, S.; Sadeghi Erami, R.; Strasser, P.; Cowan, A. J.; Farràs, P. Electrolysis of Low-Grade and Saline Surface Water. Nat. Energy 2020, 5, 367–377; https://doi.org/10.1038/s41560-020-0550-8.Suche in Google Scholar
207. Hsu, S.-H.; Miao, J.; Zhang, L.; Gao, J.; Wang, H.; Tao, H.; Hung, S.-F.; Vasileff, A.; Qiao, S. Z.; Liu, B. An Earth-Abundant Catalyst-Based Seawater Photoelectrolysis System with 17.9% Solar-To-Hydrogen Efficiency. Adv. Mater. 2018, 30, 1707261. https://doi.org/10.1002/adma.201707261.Suche in Google Scholar PubMed
208. Yu, L.; Zhu, Q.; Song, S.; McElhenny, B.; Wang, D.; Wu, C.; Qin, Z.; Bao, J.; Yu, Y.; Chen, S.; Ren, Z. Non-Noble Metal-Nitride Based Electrocatalysts for High-Performance Alkaline Seawater Electrolysis. Nat. Commun. 2019, 10, 5106; https://doi.org/10.1038/s41467-019-13092-7.Suche in Google Scholar PubMed PubMed Central
209. Cheng, F.; Feng, X.; Chen, X.; Lin, W.; Rong, J.; Yang, W. Synergistic Action of Co-fe Layered Double Hydroxide Electrocatalyst and Multiple Ions of Sea Salt for Efficient Seawater Oxidation at Near-Neutral PH. Electrochim. Acta 2017, 251, 336–343. https://doi.org/10.1016/j.electacta.2017.08.098.Suche in Google Scholar
210. Lv, Q.; Han, J.; Tan, X.; Wang, W.; Cao, L.; Dong, B. Featherlike NiCoP Holey Nanoarrys for Efficient and Stable Seawater Splitting. ACS Appl. Energy Mater. 2019, 2, 3910–3917; https://doi.org/10.1021/acsaem.9b00599.Suche in Google Scholar
211. Huang, Y.; Hu, L.; Liu, R.; Hu, Y.; Xiong, T.; Qiu, W.; Balogun, M.-S.; Jie, T.; Pan, A.; Tong, Y. Nitrogen Treatment Generates Tunable Nanohybridization of Ni5P4 Nanosheets with Nickel Hydr(Oxy)Oxides for Efficient Hydrogen Production in Alkaline, Seawater and Acidic Media. Appl. Catal. B Environ. 2019, 251, 181–194. https://doi.org/10.1016/j.apcatb.2019.03.037.Suche in Google Scholar
212. Lu, X.; Pan, J.; Lovell, E.; Tan, T. H.; Ng, Y. H.; Amal, R. A Sea-Change: Manganese Doped Nickel/Nickel Oxide Electrocatalysts for Hydrogen Generation from Seawater. Energy Environ. Sci. 2018, 11, 1898–1910; https://doi.org/10.1039/C8EE00976G.Suche in Google Scholar
213. Sarno, M.; Ponticorvo, E.; Scarpa, D. Active and Stable Graphene Supporting Trimetallic Alloy-Based Electrocatalyst for Hydrogen Evolution by Seawater Splitting. Electrochem. Commun. 2020, 111, 106647. https://doi.org/10.1016/j.elecom.2019.106647.Suche in Google Scholar
214. Kuang, Y.; Kenney, M. J.; Meng, Y.; Hung, W.-H.; Liu, Y.; Huang, J. E.; Prasanna, R.; Li, P.; Li, Y.; Wang, L.; Lin, M. C.; McGehee, M. D.; Sun, X.; Dai, H. Solar-Driven, Highly Sustained Splitting of Seawater into Hydrogen and Oxygen Fuels. Proc. Natl. Acad. Sci. 2019, 116, 6624–6629; https://doi.org/10.1073/pnas.1900556116.Suche in Google Scholar PubMed PubMed Central
215. Huang, X.; Chang, S.; Lee, W. S. V.; Ding, J.; Xue, J. M. Three-Dimensional Printed Cellular Stainless Steel as a High-Activity Catalytic Electrode for Oxygen Evolution. J. Mater. Chem. A 2017, 5, 18176–18182; https://doi.org/10.1039/C7TA03023A.Suche in Google Scholar
216. Jin, H.; Liu, X.; Vasileff, A.; Jiao, Y.; Zhao, Y.; Zheng, Y.; Qiao, S.-Z. Single-Crystal Nitrogen-Rich Two-Dimensional Mo5N6 Nanosheets for Efficient and Stable Seawater Splitting. ACS Nano 2018, 12, 12761–12769; https://doi.org/10.1021/acsnano.8b07841.Suche in Google Scholar PubMed
217. Zhou, G.; Guo, Z.; Shan, Y.; Wu, S.; Zhang, J.; Yan, K.; Liu, L.; Chu, P. K.; Wu, X. High-Efficiency Hydrogen Evolution from Seawater Using Hetero-Structured T/Td Phase ReS2 Nanosheets with Cationic Vacancies. Nano Energy 2019, 55, 42–48. https://doi.org/10.1016/j.nanoen.2018.10.047.Suche in Google Scholar
218. Zheng, J.; Zhao, Y.; Xi, H.; Li, C. Seawater Splitting for Hydrogen Evolution by Robust Electrocatalysts from Secondary M (M = Cr, Fe, Co, Ni, Mo) Incorporated Pt. RSC Adv. 2018, 8, 9423–9429; https://doi.org/10.1039/C7RA12112A.Suche in Google Scholar
219. Zhang, Y.; Li, P.; Yang, X.; Fa, W.; Ge, S. High-Efficiency and Stable Alloyed Nickel Based Electrodes for Hydrogen Evolution by Seawater Splitting. J. Alloys Compd. 2018, 732, 248–256. https://doi.org/10.1016/j.jallcom.2017.10.194.Suche in Google Scholar
220. Niu, X.; Tang, Q.; He, B.; Yang, P. Robust and Stable Ruthenium Alloy Electrocatalysts for Hydrogen Evolution by Seawater Splitting. Electrochim. Acta 2016, 208, 180–187. https://doi.org/10.1016/j.electacta.2016.04.184.Suche in Google Scholar
221. Li, H.; Tang, Q.; He, B.; Yang, P. Robust Electrocatalysts from an Alloyed Pt–Ru–M (M = Cr, Fe, Co, Ni, Mo)-Decorated Ti Mesh for Hydrogen Evolution by Seawater Splitting. J. Mater. Chem. A 2016, 4, 6513–6520; https://doi.org/10.1039/C6TA00785F.Suche in Google Scholar
222. Gao, S.; Li, G.-D.; Liu, Y.; Chen, H.; Feng, L.-L.; Wang, Y.; Yang, M.; Wang, D.; Wang, S.; Zou, X. Electrocatalytic H2 Production from Seawater over Co, N-Codoped Nanocarbons. Nanoscale 2015, 7, 2306–2316; https://doi.org/10.1039/C4NR04924A.Suche in Google Scholar PubMed
223. Hausmann, J. N.; Schlögl, R.; Menezes, P. W.; Driess, M. Is Direct Seawater Splitting Economically Meaningful? Energy Environ. Sci. 2021, 14, 3679–3685; https://doi.org/10.1039/d0ee03659e.Suche in Google Scholar
224. Khan, M. A.; Al-Attas, T.; Roy, S.; Rahman, M. M.; Ghaffour, N.; Thangadurai, V.; Larter, S.; Hu, J.; Ajayan, P. M.; Kibria, M. G. Seawater Electrolysis for Hydrogen Production: A Solution Looking for a Problem? Energy Environ. Sci. 2021, 14, 4831–4839; https://doi.org/10.1039/D1EE00870F.Suche in Google Scholar
225. Franco, A.; Giovannini, C. Recent and Future Advances in Water Electrolysis for Green Hydrogen Generation: Critical Analysis and Perspectives. Sustainability 2023, 15, 16917; https://doi.org/10.3390/su152416917.Suche in Google Scholar
226. Gerloff, N. Comparative Life-Cycle-Assessment Analysis of Three Major Water Electrolysis Technologies while Applying Various Energy Scenarios for a Greener Hydrogen Production. J. Energy Storage 2021, 43, 102759; https://doi.org/10.1016/j.est.2021.102759.Suche in Google Scholar
227. Lehner, M.; Tichler, R.; Steinmller, H.; Kopper, M. Power-to-Gas: Technology and Business Models; Springer: USA, 2014.10.1007/978-3-319-03995-4Suche in Google Scholar
228. Kilner, A.; Skinner, S.; Irvine, J.; Edwards, P. Functional Materials for Sustainable Energy Applications; Woodhead Publishing Series in Energy: UK, 2012.10.1533/9780857096371Suche in Google Scholar
229. Crnkovic, F. C.; Machado, S. A. S.; Avaca, L. A. Electrochemical and Morphological Studies of Electrodeposited Ni–Fe–Mo–Zn Alloys Tailored for Water Electrolysis. Int. J. Hydrogen Energy 2004, 29, 249–254. https://doi.org/10.1016/S0360-3199(03)00212-X.Suche in Google Scholar
230. Han, Q.; Liu, K.; Chen, J.; Wei, X. Hydrogen Evolution Reaction on Amorphous Ni–S–Co Alloy in Alkaline Medium. Int. J. Hydrogen Energy 2003, 28, 1345–1352. https://doi.org/10.1016/S0360-3199(03)00043-0.Suche in Google Scholar
231. Sheela, G.; Pushpavanam, M.; Pushpavanam, S. Zinc–Nickel Alloy Electrodeposits for Water Electrolysis. Int. J. Hydrogen Energy 2002, 27, 627–633. https://doi.org/10.1016/S0360-3199(01)00170-7.Suche in Google Scholar
232. Hu, W. Electrocatalytic Properties of New Electrocatalysts for Hydrogen Evolution in Alkaline Water Electrolysis. Int. J. Hydrogen Energy 2000, 25, 111–118. https://doi.org/10.1016/S0360-3199(99)00024-5.Suche in Google Scholar
233. Hu, W.; Lee, J.-Y. Electrocatalytic Properties of Ti2Ni/Ni-Mo Composite Electrodes for Hydrogen Evolution Reaction. Int. J. Hydrogen Energy 1998, 23, 253–257. https://doi.org/10.1016/S0360-3199(97)00060-8.Suche in Google Scholar
234. Los, P.; Rami, A.; Lasia, A. Hydrogen Evolution Reaction on Ni-Al Electrodes. J. Appl. Electrochem. 1993, 23, 135–140; https://doi.org/10.1007/BF00246950.Suche in Google Scholar
235. Raj, I. A. Nickel-Based, Binary-Composite Electrocatalysts for the Cathodes in the Energy-Efficient Industrial Production of Hydrogen from Alkaline-Water Electrolytic Cells. J. Mater. Sci. 1993, 28, 4375–4382; https://doi.org/10.1007/BF01154945.Suche in Google Scholar
236. Singh, R. N.; Mishra, D.; Anindita; Sinha, A. S. K.; Singh, A. Novel Electrocatalysts for Generating Oxygen from Alkaline Water Electrolysis. Electrochem. Commun. 2007, 9, 1369–1373. https://doi.org/10.1016/j.elecom.2007.01.044.Suche in Google Scholar
237. El-Deab, M. S.; Awad, M. I.; Mohammad, A. M.; Ohsaka, T. Enhanced Water Electrolysis: Electrocatalytic Generation of Oxygen Gas at Manganese Oxide Nanorods Modified Electrodes. Electrochem. Commun. 2007, 9, 2082–2087. https://doi.org/10.1016/j.elecom.2007.06.011.Suche in Google Scholar
238. Hamdani, M.; Pereira, M. I. S.; Douch, J.; Ait Addi, A.; Berghoute, Y.; Mendonça, M. H. Physicochemical and Electrocatalytic Properties of Li-Co3O4 Anodes Prepared by Chemical Spray Pyrolysis for Application in Alkaline Water Electrolysis. Electrochim. Acta 2004, 49, 1555–1563. https://doi.org/10.1016/j.electacta.2003.11.016.Suche in Google Scholar
239. Wendt, H.; Hofmann, H.; Plzak, V. Materials Research and Development of Electrocatalysts for Alkaline Water Electrolysis. Mater. Chem. Phys. 1989, 22, 27–49. https://doi.org/10.1016/0254-0584(89)90030-8.Suche in Google Scholar
240. Krstajić, N.; Popović, M.; Grgur, B.; Vojnović, M.; Šepa, D. On the Kinetics of the Hydrogen Evolution Reaction on Nickel in Alkaline Solution: Part II. Effect of Temperature. J. Electroanal. Chem. 2001, 512, 27–35. https://doi.org/10.1016/S0022-0728(01)00591-5.Suche in Google Scholar
241. Gennero de Chialvo, M. R.; Chialvo, A. C. Hydrogen Evolution Reaction on a Smooth Iron Electrode in Alkaline Solution at Different Temperatures. Phys. Chem. Chem. Phys. 2001, 3, 3180–3184; https://doi.org/10.1039/b102777h.Suche in Google Scholar
242. Science, E. Hydrogen Overpotential on Pure Metals. J. Electrochem. Soc. 1971, 118, 1278–1282; https://doi.org/10.1149/1.2408305.Suche in Google Scholar
243. Correia, A. N.; Machado, S. A. S.; Avaca, L. A. Studies of the Hydrogen Evolution Reaction on Smooth Co and Electrodeposited Ni–Co Ultramicroelectrodes. Electrochem. Commun. 1999, 1, 600–604. https://doi.org/10.1016/S1388-2481(99)00122-8.Suche in Google Scholar
244. Jom, B.; BE, C.; E, Y.; RE, W. Comprehensive Treatise of Electrochemistry; Plenum Press: New York; pp 19981.Suche in Google Scholar
245. Miles, M. H.; Huang, Y. H.; Srinivasan, S. The Oxygen Electrode Reaction in Alkaline Solutions on Oxide Electrodes Prepared by the Thermal Decomposition Method. J. Electrochem. Soc. 1978, 125, 1931; https://doi.org/10.1149/1.2131330.Suche in Google Scholar
246. Damjanovic, A.; Dey, A.; Bockris, J. O. Electrode Kinetics of Oxygen Evolution and Dissolution on Rh, Ir, and Pt‐Rh Alloy Electrodes. J. Electrochem. Soc. 1966, 113, 739; https://doi.org/10.1149/1.2424104.Suche in Google Scholar
247. Miles, M. H.; Kissel, G.; Lu, P. W. T.; Srinivasan, S. Effect of Temperature on Electrode Kinetic Parameters for Hydrogen and Oxygen Evolution Reactions on Nickel Electrodes in Alkaline Solutions. J. Electrochem. Soc. 1976, 123, 332; https://doi.org/10.1149/1.2132820.Suche in Google Scholar
248. Bloor, L. G.; Solarska, R.; Bienkowski, K.; Kulesza, P. J.; Augustynski, J.; Symes, M. D.; Cronin, L. Solar-Driven Water Oxidation and Decoupled Hydrogen Production Mediated by an Electron-Coupled-Proton Buffer. J. Am. Chem. Soc. 2016, 138, 6707–6710; https://doi.org/10.1021/jacs.6b03187.Suche in Google Scholar PubMed PubMed Central
249. Li, F.; Yu, F.; Du, J.; Wang, Y.; Zhu, Y.; Li, X.; Sun, L. Water Splitting via Decoupled Photocatalytic Water Oxidation and Electrochemical Proton Reduction Mediated by Electron-Coupled-Proton Buffer. Chem. – An Asian J. 2017, 12, 2666–2669. https://doi.org/10.1002/asia.201701123.Suche in Google Scholar PubMed
250. Chisholm, G.; Cronin, L.; Symes, M. D. Decoupled Electrolysis Using a Silicotungstic Acid Electron-Coupled-Proton Buffer in a Proton Exchange Membrane Cell. Electrochim. Acta 2020, 331, 135255. https://doi.org/10.1016/j.electacta.2019.135255.Suche in Google Scholar
251. Chen, K.; Wu, C.-D. Designed Fabrication of Biomimetic Metal–Organic Frameworks for Catalytic Applications. Coord. Chem. Rev. 2019, 378, 445–465. https://doi.org/10.1016/j.ccr.2018.01.016.Suche in Google Scholar
252. Lei, J.; Yang, J.-J.; Liu, T.; Yuan, R.-M.; Deng, D.-R.; Zheng, M.-S.; Chen, J.-J.; Cronin, L.; Dong, Q.-F. Tuning Redox Active Polyoxometalates for Efficient Electron-Coupled Proton-Buffer-Mediated Water Splitting. Chem. – A Eur. J. 2019, 25, 11432–11436. https://doi.org/10.1002/chem.201903142.Suche in Google Scholar PubMed PubMed Central
253. Chen, J.-J.; Symes, M. D.; Cronin, L. Highly Reduced and Protonated Aqueous Solutions of [P2W18O62]6− for On-Demand Hydrogen Generation and Energy Storage. Nat. Chem. 2018, 10, 1042–1047; https://doi.org/10.1038/s41557-018-0109-5.Suche in Google Scholar PubMed
254. Amstutz, V.; Toghill, K. E.; Powlesland, F.; Vrubel, H.; Comninellis, C.; Hu, X.; Girault, H. H. Renewable Hydrogen Generation from a Dual-Circuit Redox Flow Battery. Energy Environ. Sci. 2014, 7, 2350–2358; https://doi.org/10.1039/C4EE00098F.Suche in Google Scholar
255. Ma, C.; Pei, S.; You, S. Closed Bipolar Electrode for Decoupled Electrochemical Water Decontamination and Hydrogen Recovery. Electrochem. Commun. 2019, 109, 106611. https://doi.org/10.1016/j.elecom.2019.106611.Suche in Google Scholar
256. Jin, Z.; Li, P.; Xiao, D. A Hydrogen-Evolving Hybrid-Electrolyte Battery with Electrochemical/Photoelectrochemical Charging from Water Oxidation. ChemSusChem 2017, 10, 483–488. https://doi.org/10.1002/cssc.201601317.Suche in Google Scholar PubMed
257. Choi, B.; Panthi, D.; Nakoji, M.; Kabutomori, T.; Tsutsumi, K.; Tsutsumi, A. A Novel Water-Splitting Electrochemical Cycle for Hydrogen Production Using an Intermediate Electrode. Chem. Eng. Sci. 2017, 157, 200–208. https://doi.org/10.1016/j.ces.2016.04.060.Suche in Google Scholar
258. Palumbo, R.; Diver, R. B.; Larson, C.; Coker, E. N.; Miller, J. E.; Guertin, J.; Schoer, J.; Meyer, M.; Siegel, N. P. Solar Thermal Decoupled Water Electrolysis Process I: Proof of Concept. Chem. Eng. Sci. 2012, 84, 372–380. https://doi.org/10.1016/j.ces.2012.08.023.Suche in Google Scholar
259. Nudehi, S.; Larson, C.; Prusinski, W.; Kotfer, D.; Otto, J.; Beyers, E.; Schoer, J.; Palumbo, R. Solar Thermal Decoupled Water Electrolysis Process II: An Extended Investigation of the Anodic Electrochemical Reaction. Chem. Eng. Sci. 2018, 181, 159–172. https://doi.org/10.1016/j.ces.2017.12.032.Suche in Google Scholar
260. Ma, Y.; Guo, Z.; Dong, X.; Wang, Y.; Xia, Y. Organic Proton-Buffer Electrode to Separate Hydrogen and Oxygen Evolution in Acid Water Electrolysis. Angew. Chemie Int. Ed. 2019, 58, 4622–4626. https://doi.org/10.1002/anie.201814625.Suche in Google Scholar PubMed
261. Wang, J.; Ji, L.; Teng, X.; Liu, Y.; Guo, L.; Chen, Z. Decoupling Half-Reactions of Electrolytic Water Splitting by Integrating a Polyaniline Electrode. J. Mater. Chem. A 2019, 7, 13149–13153; https://doi.org/10.1039/C9TA03285A.Suche in Google Scholar
262. Chen, G.; Li, X.; Feng, X. Upgrading Organic Compounds through the Coupling of Electrooxidation with Hydrogen Evolution. Angew. Chemie - Int. Ed. 2022, 61; https://doi.org/10.1002/anie.202209014.Suche in Google Scholar PubMed PubMed Central
263. Xiang, K.; Wu, D.; Deng, X.; Li, M.; Chen, S.; Hao, P.; Guo, X.; Luo, J.-L.; Fu, X.-Z. Boosting H2 Generation Coupled with Selective Oxidation of Methanol into Value-Added Chemical over Cobalt Hydroxide@Hydroxysulfide Nanosheets Electrocatalysts. Adv. Funct. Mater. 2020, 30, 1909610. https://doi.org/10.1002/adfm.201909610.Suche in Google Scholar
264. Dai, L.; Qin, Q.; Zhao, X.; Xu, C.; Hu, C.; Mo, S.; Wang, Y. O.; Lin, S.; Tang, Z.; Zheng, N. Electrochemical Partial Reforming of Ethanol into Ethyl Acetate Using Ultrathin Co3O4 Nanosheets as a Highly Selective Anode Catalyst. ACS Cent. Sci. 2016, 2, 538–544; https://doi.org/10.1021/acscentsci.6b00164.Suche in Google Scholar PubMed PubMed Central
265. Zhao, Y.; Xing, S.; Meng, X.; Zeng, J.; Yin, S.; Li, X.; Chen, Y. Ultrathin Rh Nanosheets as a Highly Efficient Bifunctional Electrocatalyst for Isopropanol-Assisted Overall Water Splitting. Nanoscale 2019, 11, 9319–9326; https://doi.org/10.1039/C9NR02153A.Suche in Google Scholar
266. Zheng, J.; Chen, X.; Zhong, X.; Li, S.; Liu, T.; Zhuang, G.; Li, X.; Deng, S.; Mei, D.; Wang, J.-G. Hierarchical Porous NC@CuCo Nitride Nanosheet Networks: Highly Efficient Bifunctional Electrocatalyst for Overall Water Splitting and Selective Electrooxidation of Benzyl Alcohol. Adv. Funct. Mater. 2017, 27, 1704169. https://doi.org/10.1002/adfm.201704169.Suche in Google Scholar
267. Cui, X.; Chen, M.; Xiong, R.; Sun, J.; Liu, X.; Geng, B. Ultrastable and Efficient H2 Production via Membrane-free Hybrid Water Electrolysis over a Bifunctional Catalyst of Hierarchical Mo–Ni Alloy Nanoparticles. J. Mater. Chem. A 2019, 7, 16501–16507; https://doi.org/10.1039/C9TA03924D.Suche in Google Scholar
268. Si, D.; Xiong, B.; Chen, L.; Shi, J. Highly Selective and Efficient Electrocatalytic Synthesis of Glycolic Acid in Coupling with Hydrogen Evolution. Chem Catal. 2021, 1, 941–955; https://doi.org/10.1016/j.checat.2021.08.001.Suche in Google Scholar
269. Yu, X.; dos Santos, E. C.; White, J.; Salazar-Alvarez, G.; Pettersson, L. G. M.; Cornell, A.; Johnsson, M. Electrocatalytic Glycerol Oxidation with Concurrent Hydrogen Evolution Utilizing an Efficient MoOx/Pt Catalyst. Small 2021, 17, 2104288. https://doi.org/10.1002/smll.202104288.Suche in Google Scholar PubMed
270. Lin, C.; Zhang, P.; Wang, S.; Zhou, Q.; Na, B.; Li, H.; Tian, J.; Zhang, Y.; Deng, C.; Meng, L.; Wu, J.; Liu, C.; Hu, J.; Zhang, L. Engineered Porous Co–ni Alloy on Carbon Cloth as an Efficient Bifunctional Electrocatalyst for Glucose Electrolysis in Alkaline Environment. J. Alloys Compd. 2020, 823, 153784. https://doi.org/10.1016/j.jallcom.2020.153784.Suche in Google Scholar
271. You, B.; Liu, X.; Jiang, N.; Sun, Y. A General Strategy for Decoupled Hydrogen Production from Water Splitting by Integrating Oxidative Biomass Valorization. J. Am. Chem. Soc. 2016, 138, 13639–13646; https://doi.org/10.1021/jacs.6b07127.Suche in Google Scholar PubMed
272. Jiang, N.; Liu, X.; Dong, J.; You, B.; Liu, X.; Sun, Y. Electrocatalysis of Furfural Oxidation Coupled with H2 Evolution via Nickel-Based Electrocatalysts in Water. ChemNanoMat 2017, 3, 491–495. https://doi.org/10.1002/cnma.201700076.Suche in Google Scholar
273. Kubota, S. R.; Choi, K.-S. Electrochemical Valorization of Furfural to Maleic Acid. ACS Sustain. Chem. Eng. 2018, 6, 9596–9600; https://doi.org/10.1021/acssuschemeng.8b02698.Suche in Google Scholar
274. Yang, G.; Jiao, Y.; Yan, H.; Xie, Y.; Wu, A.; Dong, X.; Guo, D.; Tian, C.; Fu, H. Interfacial Engineering of MoO2-FeP Heterojunction for Highly Efficient Hydrogen Evolution Coupled with Biomass Electrooxidation. Adv. Mater. 2020, 32, 2000455. https://doi.org/10.1002/adma.202000455.Suche in Google Scholar PubMed
275. Gu, K.; Wang, D.; Xie, C.; Wang, T.; Huang, G.; Liu, Y.; Zou, Y.; Tao, L.; Wang, S. Defect-Rich High-Entropy Oxide Nanosheets for Efficient 5-Hydroxymethylfurfural Electrooxidation. Angew. Chemie Int. Ed. 2021, 60, 20253–20258. https://doi.org/10.1002/anie.202107390.Suche in Google Scholar PubMed
276. Wang, W.; Wang, Y.; Yang, R.; Wen, Q.; Liu, Y.; Jiang, Z.; Li, H.; Zhai, T. Vacancy-Rich Ni(OH)2 Drives the Electrooxidation of Amino C−N Bonds to Nitrile C≡N Bonds. Angew. Chemie Int. Ed. 2020, 59, 16974–16981. https://doi.org/10.1002/anie.202005574.Suche in Google Scholar PubMed
277. Chong, X.; Liu, C.; Wang, C.; Yang, R.; Zhang, B. Integrating Hydrogen Production and Transfer Hydrogenation with Selenite Promoted Electrooxidation of α-Nitrotoluenes to E-Nitroethenes. Angew. Chemie Int. Ed. 2021, 60, 22010–22016. https://doi.org/10.1002/anie.202108666.Suche in Google Scholar PubMed
278. Huang, C.; Huang, Y.; Liu, C.; Yu, Y.; Zhang, B. Integrating Hydrogen Production with Aqueous Selective Semi-dehydrogenation of Tetrahydroisoquinolines over a Ni2P Bifunctional Electrode. Angew. Chemie Int. Ed. 2019, 58, 12014–12017. https://doi.org/10.1002/anie.201903327.Suche in Google Scholar PubMed
279. Ma, L.; Zhou, H.; Xu, M.; Hao, P.; Kong, X.; Duan, H. Integrating Hydrogen Production with Anodic Selective Oxidation of Sulfides over a CoFe Layered Double Hydroxide Electrode. Chem. Sci. 2021, 12, 938–945; https://doi.org/10.1039/D0SC05499B.Suche in Google Scholar PubMed PubMed Central
280. Hu, K.; Xu, L.; Chen, W.; Jia, S.; Wang, W.; Han, F. Degradation of Organics Extracted from Dewatered Sludge by Alkaline Pretreatment in Microbial Electrolysis Cell. Environ. Sci. Pollut. Res. 2018, 25, 8715–8724; https://doi.org/10.1007/s11356-018-1213-1.Suche in Google Scholar PubMed
281. Hasany, M.; Yaghmaei, S.; Mardanpour, M. M.; Ghasemi Naraghi, Z. Simultaneously Energy Production and Dairy Wastewater Treatment Using Bioelectrochemical Cells: In Different Environmental and Hydrodynamic Modes. Chinese J. Chem. Eng. 2017, 25, 1847–1855. https://doi.org/10.1016/j.cjche.2017.08.003.Suche in Google Scholar
282. Jeremiasse, A. W.; Hamelers, H. V. M.; Saakes, M.; Buisman, C. J. N. Ni Foam Cathode Enables High Volumetric H2 Production in a Microbial Electrolysis Cell. Int. J. Hydrogen Energy 2010, 35, 12716–12723. https://doi.org/10.1016/j.ijhydene.2010.08.131.Suche in Google Scholar
283. Cheng, S.; Logan, B. E. Sustainable and Efficient Biohydrogen Production via Electrohydrogenesis. Proc. Natl. Acad. Sci. 2007, 104, 18871–18873; https://doi.org/10.1073/pnas.0706379104.Suche in Google Scholar PubMed PubMed Central
284. Heidrich, E. S.; Dolfing, J.; Scott, K.; Edwards, S. R.; Jones, C.; Curtis, T. P. Production of Hydrogen from Domestic Wastewater in a Pilot-Scale Microbial Electrolysis Cell. Appl. Microbiol. Biotechnol. 2013, 97, 6979–6989; https://doi.org/10.1007/s00253-012-4456-7.Suche in Google Scholar PubMed
285. Lu, L.; Xing, D.; Xie, T.; Ren, N.; Logan, B. E. Hydrogen Production from Proteins via Electrohydrogenesis in Microbial Electrolysis Cells. Biosens. Bioelectron. 2010, 25, 2690–2695. https://doi.org/10.1016/j.bios.2010.05.003.Suche in Google Scholar PubMed
286. Rozendal, R. A.; Hamelers, H. V. M.; Euverink, G. J. W.; Metz, S. J.; Buisman, C. J. N. Principle and Perspectives of Hydrogen Production through Biocatalyzed Electrolysis. Int. J. Hydrogen Energy 2006, 31, 1632–1640. https://doi.org/10.1016/j.ijhydene.2005.12.006.Suche in Google Scholar
287. Badia-Fabregat, M.; Rago, L.; Baeza, J. A.; Guisasola, A. Hydrogen Production from Crude Glycerol in an Alkaline Microbial Electrolysis Cell. Int. J. Hydrogen Energy 2019, 44, 17204–17213. https://doi.org/10.1016/j.ijhydene.2019.03.193.Suche in Google Scholar
288. Zhang, L.; Wang, Y.-Z.; Zhao, T.; Xu, T. Hydrogen Production from Simultaneous Saccharification and Fermentation of Lignocellulosic Materials in a Dual-Chamber Microbial Electrolysis Cell. Int. J. Hydrogen Energy 2019, 44, 30024–30030. https://doi.org/10.1016/j.ijhydene.2019.09.191.Suche in Google Scholar
289. Chen, J.; Xu, W.; Wu, X.; E, J.; Lu, N.; Wang, T.; Zuo, H. System Development and Environmental Performance Analysis of a Pilot Scale Microbial Electrolysis Cell for Hydrogen Production Using Urban Wastewater. Energy Convers. Manag. 2019, 193, 52–63. https://doi.org/10.1016/j.enconman.2019.04.060.Suche in Google Scholar
290. N, S.; Spurgeon, J.; Matheswaran, M.; Satyavolu, J. Simultaneous Biohydrogen Production with Distillery Wastewater Treatment Using Modified Microbial Electrolysis Cell. Int. J. Hydrogen Energy 2020, 45, 18266–18274. https://doi.org/10.1016/j.ijhydene.2019.06.134.Suche in Google Scholar
291. Hou, Y.; Zhang, R.; Yu, Z.; Huang, L.; Liu, Y.; Zhou, Z. Accelerated Azo Dye Degradation and Concurrent Hydrogen Production in the Single-Chamber Photocatalytic Microbial Electrolysis Cell. Bioresour. Technol. 2017, 224, 63–68. https://doi.org/10.1016/j.biortech.2016.10.069.Suche in Google Scholar PubMed
292. Wan, L.-L.; Li, X.-J.; Zang, G.-L.; Wang, X.; Zhang, Y.-Y.; Zhou, Q.-X. A Solar Assisted Microbial Electrolysis Cell for Hydrogen Production Driven by a Microbial Fuel Cell. RSC Adv. 2015, 5, 82276–82281; https://doi.org/10.1039/C5RA16919D.Suche in Google Scholar
293. Yang, N.; Hafez, H.; Nakhla, G. Impact of Volatile Fatty Acids on Microbial Electrolysis Cell Performance. Bioresour. Technol. 2015, 193, 449–455. https://doi.org/10.1016/j.biortech.2015.06.124.Suche in Google Scholar PubMed
294. Zhang, Y.; Angelidaki, I. Innovative Self-Powered Submersible Microbial Electrolysis Cell (SMEC) for Biohydrogen Production from Anaerobic Reactors. Water Res. 2012, 46, 2727–2736. https://doi.org/10.1016/j.watres.2012.02.038.Suche in Google Scholar PubMed
295. Asztalos, J. R.; Kim, Y. Enhanced Digestion of Waste Activated Sludge Using Microbial Electrolysis Cells at Ambient Temperature. Water Res. 2015, 87, 503–512. https://doi.org/10.1016/j.watres.2015.05.045.Suche in Google Scholar PubMed
296. Wang, A.; Sun, D.; Cao, G.; Wang, H.; Ren, N.; Wu, W.-M.; Logan, B. E. Integrated Hydrogen Production Process from Cellulose by Combining Dark Fermentation, Microbial Fuel Cells, and a Microbial Electrolysis Cell. Bioresour. Technol. 2011, 102, 4137–4143. https://doi.org/10.1016/j.biortech.2010.10.137.Suche in Google Scholar PubMed
297. Li, X.-H.; Liang, D.-W.; Bai, Y.-X.; Fan, Y.-T.; Hou, H.-W. Enhanced H2 Production from Corn Stalk by Integrating Dark Fermentation and Single Chamber Microbial Electrolysis Cells with Double Anode Arrangement. Int. J. Hydrogen Energy 2014, 39, 8977–8982. https://doi.org/10.1016/j.ijhydene.2014.03.065.Suche in Google Scholar
298. Zhen, G.; Kobayashi, T.; Lu, X.; Kumar, G.; Hu, Y.; Bakonyi, P.; Rózsenberszki, T.; Koók, L.; Nemestóthy, N.; Bélafi-Bakó, K.; Xu, K. Recovery of Biohydrogen in a Single-Chamber Microbial Electrohydrogenesis Cell Using Liquid Fraction of Pressed Municipal Solid Waste (LPW) as Substrate. Int. J. Hydrogen Energy 2016, 41, 17896–17906. https://doi.org/10.1016/j.ijhydene.2016.07.112.Suche in Google Scholar
299. Jayabalan, T.; Matheswaran, M.; Naina Mohammed, S. Biohydrogen Production from Sugar Industry Effluents Using Nickel Based Electrode Materials in Microbial Electrolysis Cell. Int. J. Hydrogen Energy 2019, 44, 17381–17388. https://doi.org/10.1016/j.ijhydene.2018.09.219.Suche in Google Scholar
300. Gunasekaran, M.; Merrylin, J.; Usman, T. M. M.; Kumar, G.; Kim, S.-H.; Banu, J. R. Chapter 30 - Biohydrogen Production from Industrial Wastewater. In Biomass, Biofuels, Biochemicals; Pandey, A.; Larroche, C.; Dussap, C.-G.; Gnansounou, E.; Khanal, S.K.; Ricke, S.B.T.-B.A.F.C.P. and G.B., Eds., 2nd ed.; Academic Press: USA, 2019; pp 733–760.10.1016/B978-0-12-816856-1.00030-0Suche in Google Scholar
301. Lu, L.; Ren, N.; Xing, D.; Logan, B. E. Hydrogen Production with Effluent from an Ethanol–H2-Coproducing Fermentation Reactor Using a Single-Chamber Microbial Electrolysis Cell. Biosens. Bioelectron. 2009, 24, 3055–3060. https://doi.org/10.1016/j.bios.2009.03.024.Suche in Google Scholar PubMed
302. Cucu, A.; Costache, T. A.; Divona, M.; Tiliakos, A.; Stamatin, I.; Ciocanea, A. Microbial Electrolysis Cell: Hydrogen Production Using Microbial Consortia from Romanian Waters. Dig. J. Nanomater. Biostruct. 2013, 8, 1179–1190.Suche in Google Scholar
303. Croese, E.; Jeremiasse, A. W.; Marshall, I. P. G.; Spormann, A. M.; Euverink, G.-J. W.; Geelhoed, J. S.; Stams, A. J. M.; Plugge, C. M. Influence of Setup and Carbon Source on the Bacterial Community of Biocathodes in Microbial Electrolysis Cells. Enzym. Microb. Technol. 2014, 61–62, 67–75. https://doi.org/10.1016/j.enzmictec.2014.04.019.Suche in Google Scholar PubMed
304. Yang, Q.; Jiang, Y.; Xu, Y.; Qiu, Y.; Chen, Y.; Zhu, S.; Shen, S. Hydrogen Production with Polyaniline/Multi-Walled Carbon Nanotube Cathode Catalysts in Microbial Electrolysis Cells. J. Chem. Technol. Biotechnol. 2015, 90, 1263–1269. https://doi.org/10.1002/jctb.4425.Suche in Google Scholar
305. Matsushima, H.; Nishida, T.; Konishi, Y.; Fukunaka, Y.; Ito, Y.; Kuribayashi, K. Water Electrolysis under Microgravity: Part 1. Experimental Technique. Electrochim. Acta 2003, 48, 4119–4125; https://doi.org/10.1016/S0013-4686(03)00579-6.Suche in Google Scholar
© 2025 the author(s), published by De Gruyter, Berlin/Boston
This work is licensed under the Creative Commons Attribution 4.0 International License.
Artikel in diesem Heft
- Frontmatter
- Advancements in CNT-based materials for optimized pharmaceutical removal via adsorption and photocatalysis
- Review on removal of heavy metals from industrial effluents by adsorption
- Cutting-edge techniques in low-temperature electrochemical water splitting: advancements in hydrogen production
- Biological activities of metal complexes with Schiff base
- Potential of organometallic complexes in medicinal chemistry
- Advancement in schiff base complexes for treatment of colon cancer
- Magnetic nanoparticles for efficient heavy metal removal: synthesis, adsorption capacity, and key experimental parameters
- Coal-based carbon/graphene quantum dots: formation mechanisms and applications
- Advancements in transition metal-catalyzed 1,2,3-triazole synthesis via azide–alkyne cycloaddition
- Gold complexes: a new frontier in the battle against lung cancer
Artikel in diesem Heft
- Frontmatter
- Advancements in CNT-based materials for optimized pharmaceutical removal via adsorption and photocatalysis
- Review on removal of heavy metals from industrial effluents by adsorption
- Cutting-edge techniques in low-temperature electrochemical water splitting: advancements in hydrogen production
- Biological activities of metal complexes with Schiff base
- Potential of organometallic complexes in medicinal chemistry
- Advancement in schiff base complexes for treatment of colon cancer
- Magnetic nanoparticles for efficient heavy metal removal: synthesis, adsorption capacity, and key experimental parameters
- Coal-based carbon/graphene quantum dots: formation mechanisms and applications
- Advancements in transition metal-catalyzed 1,2,3-triazole synthesis via azide–alkyne cycloaddition
- Gold complexes: a new frontier in the battle against lung cancer