Recent progress in electrochemical synthesis of hypochlorite and its future outlook

Liuyu Ji a, Junyang Ding a, Caiyun Wang *b, Yang Luo *cd, Qian Liu e, Guangzhi Hu f and Xijun Liu *a
aMOE Key Laboratory of New Processing Technology for Nonferrous Metals and Materials, School of Resources, Environment and Materials, Guangxi University, Nanning 530004, Guangxi, China. E-mail: xjliu@gxu.edu.cn
bGuangxi Vocational & Technical Institute of Industry, Nanning 530001, Guangxi, China. E-mail: wangcaiyun@gxgy.edu.cn
cDepartment of Physics, City University of Hong Kong, Kowloon, Hong Kong SAR 999077, China. E-mail: yluo24-c@my.cityu.edu.hk
dChina Huadian Corporation Ltd. (CHD), Beijing 100031, China
eInstitute for Advanced Study, Chengdu University, Chengdu 610106, Sichuan, China
fInstitute for Ecological Research and Pollution Control of Plateau Lakes School of Ecology and Environmental Science, Yunnan University, Kunming 650504, China

Received 8th April 2025 , Accepted 8th May 2025

First published on 9th May 2025


Abstract

Disinfection plays a critical role in ensuring the safety of drinking water during treatment. Sodium hypochlorite disinfection, a method that has been demonstrated to be both cost-effective and safe, exhibits considerable promise for widespread implementation when compared to alternative methods. Electrochemical synthesis of sodium hypochlorite solution has emerged as a preferred alternative to traditional chemical methods due to its numerous advantages, including high current efficiency, low energy consumption, ease of operation, accessible raw materials, high purity, and controllable safety. Nevertheless, the instability of the reaction in the electrochemical synthesis process poses a significant challenge to its broader implementation. This study explores the underlying principles of electrochemical synthesis for sodium hypochlorite solution and investigates the impact of various conditions on electrolysis efficiency. The objective of this study is to ascertain the most optimal electrolysis conditions. The study also examines various electrode materials for the anode and cathode, and it summarizes typical strategies for enhancing electrode performance. Furthermore, the study investigates the factors influencing the stability of sodium hypochlorite solution to enable precise regulation of its efficacy, thereby promoting the advancement of electrochemical synthesis technology for sodium hypochlorite solution.


1. Introduction

Disinfection, a common occurrence in daily life, plays a critical role in water treatment processes. Disinfection methodologies encompass a range of approaches, including chemical, physical, biological, and emerging disinfection technologies. Each disinfection method possesses its own set of advantages and disadvantages, and the selection of an optimal technique is typically contingent on factors such as water quality, treatment scale, available technology, and cost considerations. Chemical disinfection, in which chemical agents are introduced to eradicate pathogens in water, has become a prevalent method in contemporary water treatment. These methods include chlorine disinfection, chlorine dioxide disinfection, ozone disinfection, and hydrogen peroxide disinfection, to name a few.1–3

Chlorination has historically emerged as the most prevalent disinfection process. A comparison of the present method with other disinfection techniques reveals several unique advantages, including its low cost, mature technology, and ease of operation. It also provides residual disinfection capabilities, effectively preventing microbial regrowth. This extensive utilization is evidenced by its implementation in the majority of water treatment facilities globally.4,5 Liquid chlorine is currently the predominant disinfection technology used in this category. However, liquid chlorine disinfection involves complex processes and can produce by-products such as chloroform. It is imperative to note that liquid chlorine is a highly toxic and hazardous material with strong corrosive properties. In the event of a leak, liquid chlorine can cause significant personal injury and environmental contamination, and even explosion hazards. Consequently, stringent standards for the secure storage and transportation of liquid chlorine are essential.6,7 Consequently, numerous enterprises worldwide are actively seeking alternative solutions to liquid chlorine.

Sodium hypochlorite shares a similar disinfection mechanism to liquid chlorine and provides comparable disinfection efficacy to chlorine gas. Its complete miscibility with water at any proportion enables the formulation of chlorine-based disinfectants at varying concentrations, which can be tailored to specific use applications. In particular, sodium hypochlorite disinfectant demonstrates a high degree of compatibility with human tissues and does not generate by-products during the disinfection process. It offers advantages such as enhanced safety in transport and storage, minimal environmental harm, fewer safety hazards, broad applicability, and cost-effectiveness.8,9 A comparison of sodium hypochlorite disinfection with other disinfection methods reveals several significant advantages, including enhanced safety, cost-effectiveness, control of disinfection by-products, residual disinfection capacity, minimal impact on water quality, ease of operation, and broad scope of application. The product's efficacy, safety, and cost-effectiveness make it a suitable choice for a wide range of disinfection scenarios. Consequently, the substitution of liquid chlorine disinfection with sodium hypochlorite is emerging as a major trend in the future of water treatment. The application scenarios for hypochlorite solution are shown in Fig. 1.


image file: d5qm00299k-f1.tif
Fig. 1 Application scenarios of hypochlorite solution.

2. Introduction to hypochlorite solution

Sodium hypochlorite solution, often called bleach, is a cost-effective, safe, and highly efficient disinfectant and potent oxidizer with a wide range of applications. Initially, its applications were primarily in areas such as water treatment, particularly for disinfecting drinking water. As science and technology have advanced, the applications of sodium hypochlorite have expanded, resulting in a wider range of uses. In agriculture, it is employed to eliminate pesticide residues. In the medical field, the disinfecting properties of sodium hypochlorite solution are highly significant. In industry, it has become a critical component of many industrial processes.10,11

Sodium hypochlorite achieves its disinfecting and bactericidal effects via hypochlorous acid and chloride ions produced through hydrolysis.12 Hypochlorous acid (HOCl), a weak acid, dissociates into hypochlorite ions (–OCl) and protons (H+) depending on the pH of the solution. HOCl is widely regarded as the active agent for sterilization, while the concentration of –OCl plays a crucial role in determining cleaning effectiveness. Bactericidal activity depends on the ability of HOCl and –OCl to diffuse across microbial cell membranes. Ionized –OCl cannot penetrate microbial cell membranes due to the cell membrane's lipid bilayer (hydrophobic layer), which explains its weak bactericidal activity. On the other hand, HOCl can passively diffuse through the lipid bilayer of the cell membrane. Thus, the bactericidal activity of sodium hypochlorite is directly proportional to the concentration of HOCl.13,14 The mechanism of hypochlorous acid operates through two key aspects. First, its small molecular size and neutral charge allow it to effectively disrupt cell walls and viral envelopes, penetrate organisms, and oxidize nucleic acids, proteins, and enzymes in bacteria and viruses, ultimately leading to the destruction of pathogens. Second, the chloride ions generated during the reaction alter the osmotic pressure of bacteria and viruses, resulting in their deactivation and subsequent death.13,15

Sodium hypochlorite solution provides highly effective disinfection and sterilization, effectively eliminating bacteria from wastewater while also reducing color intensity. Furthermore, it is more flexible in storage and use, relatively safe, easier to operate, and carries no risk of significant leaks or air pollution. Furthermore, sodium hypochlorite solution generates no harmful by-products during disinfection, contributing to its environmental friendliness. Moreover, sodium hypochlorite solution is cost-effective, with low production costs and minimal requirements for complex equipment or costly maintenance during application.16

Currently, commercially available sodium hypochlorite solutions typically contain 5 to 12% available chlorine by mass, with varying concentrations suitable for different applications. However, its stability is constrained; at higher temperatures or during extended storage, decomposition and disproportionation reactions occur, leading to a gradual decrease in available chlorine concentration and limiting its applicability. To ensure its effectiveness, sodium hypochlorite should be stored in a cool, dry environment. Currently, the shelf life of commercially available sodium hypochlorite is limited to 3 months.17,18

3. Methods for the preparation of hypochlorite solution

Currently, chemical and electrochemical synthesis methods are the primary methods used to produce sodium hypochlorite solutions. Chemical methods usually require specific chemical reagents, leading to high procurement, transportation, and storage costs, as well as safety risks due to the use of hazardous chemicals. The production process is complicated, requiring precise control of reaction conditions, and production efficiency is relatively low. Additionally, chemical methods may introduce impurities during the reaction, compromising the purity of the product. The production of waste liquids, gases, and residues further contributes to environmental pollution. In contrast, the electrochemical synthesis method for sodium hypochlorite solutions utilizes easily accessible raw materials at a lower cost. The operating cost is mainly electricity consumption, and the production process is simpler, allowing for automated and continuous production with high efficiency, significantly lowering production costs. Moreover, sodium hypochlorite solutions produced by electrochemical synthesis are more stable and less likely to decompose than those made by chemical methods. It generates minimal waste and pollutants, enhancing its eco-friendliness. The electricity used is sourced from renewables (e.g., solar and wind energy), aligning with the “carbon neutrality” environmental goal, and there is a growing trend to replace chemical methods with this approach.19

3.1 Chemical routes

3.1.1 Alkaline chlorination. Owing to the advanced development of the chlor-alkali industry, alkaline chlorination is the predominant method used in the large-scale industrial production of sodium hypochlorite solution. The production process involves the precise control of chlorine gas flow and gas–liquid mixing rates, facilitating the reaction of purified chlorine gas with a chilled concentrated sodium hydroxide solution to produce sodium hypochlorite, sodium chloride, and water. Alternatively, industrial absorption of tail gas from chlor-alkali plants using alkali can also generate sodium hypochlorite as a by-product. Generally, the reaction temperature should be maintained below 15 °C, and the concentration of the sodium hydroxide solution is a key determinant of the available chlorine content, typically ranging from 30% to 40%.20–22
Primary reaction: 2NaOH + Cl2 → NaClO + NaCl + H2O

While this production method is simple, sodium hypochlorite is susceptible to decomposition, has limited long-term storage stability, requires frequent transportation, and experiences variability in quality.23,24

3.1.2 Bleaching powder double decomposition. The bleaching powder double decomposition method involves the reaction of bleaching powder (mainly consisting of calcium hypochlorite) with sodium carbonate to produce a sodium hypochlorite solution. The basic principle involves the chemical conversion of calcium hypochlorite in bleaching powder to sodium hypochlorite, accompanied by the formation of calcium carbonate precipitate.25,26
The primary reaction is: Ca(ClO)2 + CO2 + H2O → CaCO3↓ + 2HClO

This method uses readily available raw materials, involves higher economic costs, presents a relatively simple process flow, and is only suitable for small to medium-scale production.27

3.2 Electrochemical routes

The electrochemical method for sodium hypochlorite production employs electricity as the driving force, providing advantages such as high intensity, efficiency, reaction safety, and superior product purity. Commercially, the most prevalent types of electrolytic cells are membrane-less cells utilizing dilute brine or seawater as the electrolyte, and membrane-equipped chlor-alkali electrolytic cells. Experimental studies have shown that, under identical operating conditions, membrane-based electrolytic cells have approximately 8% reduction in current efficiency and higher energy consumption compared to membrane-less cells.28–31 Consequently, membrane-less electrolysis is more cost-effective and easier to operate.
3.2.1 Fundamental principles. As seawater passes through the electrolytic cell, H2O and Cl undergo electrolysis, and the resulting Cl2 and NaOH mix within the liquid flow, reacting chemically to form NaClO. The principle is depicted in Fig. 2a and the illustration of the chlor-alkali electrolysis cell is depicted in Fig. 2b.
In this process, the anode reaction is: 2Cl → Cl2 + 2e

The cathode reaction is: 2Na+ + 2H2O + 2e → 2NaOH + H2

The chemical reaction is: Cl2 + 2NaOH → NaClO + NaCl + H2O

The overall reaction is represented as: NaCl + H2O → NaClO + H2

image file: d5qm00299k-f2.tif
Fig. 2 (a) Operational mechanism of the sodium hypochlorite generator. (b) Schematic illustration of chlor-alkali electrolyzer.32 (c) Schematic of the seawater electrolysis system and reaction mechanisms of the production of hydrogen and free-Cl2 without recirculation (batch mode) and with recirculation.33 (d) Correlation between solution pH and chlorine products. (e) Schematic diagram of the sodium hypochlorite generator (supported by Jinan Oury Industrial Co., Ltd).

Additionally, Fig. 2c depicts a schematic diagram of the seawater electrolysis system and the reaction mechanism of anode-free chlorine in both non-circulating and circulating modes, utilizing an anion-exchange membrane (AEM).33 When the electrolyte in the electrolysis cell is non-circulating, hydrogen gas is generated at the cathode while sodium hypochlorite is produced at the anode. In the anode chamber, chloride ions (Cl) from seawater are oxidized to chlorine gas (Cl2), which subsequently reacts with sodium hydroxide (NaOH) produced at the cathode to yield sodium hypochlorite (NaClO, also known as free chlorine, free-Cl2). The anion-exchange membrane (AEM) is a barrier to inhibit the transport of multivalent cations (e.g., Mg2+ and Ca2+), thus preventing inorganic fouling. Under recirculating conditions, the electrolyte moves from the anode chamber to the cathode chamber, efficiently utilizing the sodium hypochlorite (free-Cl2) produced at the anode to sanitize the entire electrolysis cell, thus maintaining system stability through recirculation.

3.2.2 Factors influencing the effectiveness of electrolysis preparation.
3.2.2.1 pH of the electrolyte. In seawater, chloride ions undergo different chlorine evolution reactions (ClER) at the anode depending on the pH of the electrolyte, resulting in different products. At a pH below 3.0, chloride ions mainly produce chlorine gas via oxidation (2Cl → Cl2 + 2e, Eθ = 1.36 V vs. SHE). At pH values between 3.0 and 7.5, chloride ions react with water to produce hypochlorous acid (2Cl + 2H2O → 2HClO + 2H+ + 2e, Eθ = 1.48 V vs. SHE). At pH values above 7.5, chloride ions react with hydroxide ions to form hypochlorite (2Cl + 2OH → ClO + H2O + 2e, Eθ = 1.57 V vs. SHE).34,35 The correlation between pH and chlorine products is illustrated in Fig. 2d.
3.2.2.2 Temperature of the electrolyte. During the electrolytic production of sodium hypochlorite solution, the temperature of the electrolyte significantly affects the yield. A circulating cooling water system should be implemented to absorb heat and keep the reaction tank at a lower temperature, thereby improving the stability of the target product and reducing production costs.36–39
3.2.2.3 Concentration of the electrolyte. Electrolytes with high concentrations exhibit superior conductivity, thereby accelerating the electrolytic reaction and consequently enhancing the production of hypochlorite solution. Consequently, there will be a corresponding increase in production costs.40–43 Therefore, a comprehensive consideration of the relationship between production efficiency and cost is required to select the optimal electrolyte concentration.
3.2.2.4 Selection of current density. A higher current density is generally preferred to increase the production efficiency of the reaction. However, an excessively high current density can raise the temperature within the reaction tank, accelerating the decomposition of the hypochlorite solution, and causing various adverse effects.44,45 Hence, the current density should be reasonably regulated to determine an optimum value.
3.2.2.5 Selection of electrodes. In this process, the cathode can be constructed from stainless steel, which is known for its stainless and corrosion-resistant properties and is made by incorporating alloying elements such as Cr and Ni into ordinary steel. Both their electrolytic performance and corrosion resistance should be thoroughly evaluated when selecting electrode plates. The anode typically consists of plates containing metals such as titanium. Titanium boasts low density, high strength, and superior corrosion resistance, effectively mitigating electrode plate corrosion in the chlor-alkali industry and emerging as a critical material in this field.46–49 The details are elaborated on in Section 2.3.

In summary, when the pH value of the electrolyte is controlled to a level of approximately 11, the inlet water temperature is around 20 °C, the electrolyte concentration is approximately 7 g L−1, and the current density is around 1 A dm−2, the effective chlorine concentration and current efficiency are relatively high, and the operating costs are minimized. These conditions are optimal for the electrolytic production of sodium hypochlorite solution.

3.2.3 Industrial-scale sodium hypochlorite generation system for seawater applications. Numerous research setups for on-site NaClO production utilize NaCl solutions at concentrations ranging from 0.51 to 0.85 mol L−1. The essential elements of the seawater electrolysis system for sodium hypochlorite production on offshore platforms include the electrolysis power supply, electrolysis cell, sodium hypochlorite storage tank, rinsing and dewatering unit, acid cleaning system, hydrogen venting pipeline, fan, and various additional instruments.50–52 As illustrated in Fig. 2e:

(1) Electrolysis power supply: it provides the essential direct current required for electrolysis, enabling efficient electrochemical reactions to occur on the electrode plates within the electrolytic cell.

(2) Electrolytic cell: it is the main site for electrolysis, where electrode plates facilitate electrochemical reactions. As the core component of the system, the electrolytic cell enables the electrolysis of seawater using direct current to generate active chlorine. The overall efficiency of the system is significantly influenced by the performance of the cell.

(3) Sodium hypochlorite storage tank: it is specifically designed for the safe storage of the produced sodium hypochlorite solution. The tank features integrated manual and automatic diaphragm valves at the inlet and outlet, with ABS material for the diaphragm to ensure enhanced safety. A magnetic float level gauge is installed on the tank to monitor the sodium hypochlorite storage volume.

(4) Flushing and drainage function: to avoid the entry of larger solid particles from seawater into the sodium hypochlorite generation system, which could lead to blockages and equipment wear, an automatic flushing seawater filter with a 40 μm filtration precision is installed at the seawater pump inlet. When the pressure differential exceeds the predetermined threshold, the automatic backwash cycle initiates to cleanse the filter screen, removing any potential debris and sediments. This process ensures that the electrolytic cell stays clean and operates efficiently.

(5) Acid washing function: this process is utilized for the routine cleaning of the electrolytic cell to remove stubborn deposits such as calcium and magnesium, ultimately extending the lifespan of the equipment. Additionally, it can regulate the internal pH of the electrolytic cell to enhance electrolysis efficiency.

3.3 Selection of electrode materials

The electrolytic production of sodium hypochlorite is closely related to the reaction mechanism of the chlorine evolution reaction (CER), a key step in sodium hypochlorite synthesis. Chlorine gas produced by the chlorine evolution reaction is a critical intermediate in the synthesis of sodium hypochlorite. Chlorine gas reacts further in water to produce hypochlorous acid and hypochlorite ions, thus facilitating the synthesis of sodium hypochlorite. Therefore, the efficiency of the chlorine evolution reaction directly affects the yield of sodium hypochlorite.53,54 During the electrolysis process, the increased current efficiency of the chlorine evolution reaction leads to a higher yield of sodium hypochlorite. Electrodes (particularly the anode) serve as the central components of sodium hypochlorite generators and influence the chlorine evolution efficiency, energy consumption, and operating life of the equipment. Superior electrode materials should have low chlorine evolution potential, high current efficiency, excellent stability, extended service life, and low manufacturing costs.32
3.3.1 Cathode materials. The cathode remains passivated during electrolysis, ensuring excellent stability and corrosion resistance. Materials such as platinum (Pt), graphite, stainless steel, Ti, and coated titanium electrodes can be used as cathodes. However, the high cost of Pt and the low durability of graphite significantly limit their application. Additionally, coated titanium electrodes are mostly used as anode materials. As a result, stainless steel or Ti are now widely employed as cathode materials.47,55

Both stainless steel and titanium possess unique strengths and weaknesses. Stainless steel provides high strength and affordability but is susceptible to corrosion by hypochlorite solutions, causing black water issues in the generator and elevated solution turbidity. The presence of a Ti–O layer on the surface of Ti gives it high stability and superior durability compared to stainless steel, guaranteeing the quality of the produced water.56 Nevertheless, its operating cost is higher, and it has a tendency to form hydrides with hydrogen, resulting in detachment and a shorter service life (around 1–2 years).57 Thus, the choice of cathode materials should be tailored to specific circumstances.

3.3.2 Anode materials. In the process of electrolytic synthesis of sodium hypochlorite solution, the chlorine evolution reaction at the anode is inevitably accompanied by the oxygen evolution side reaction (OER).58–60 This not only reduces the current efficiency but also significantly shortens the service life of the anode. From the perspective of thermodynamics, the standard electrode potential of the oxygen evolution reaction (OER) is 1.23 V, which is more favorable than that of the chlorine evolution reaction (CER) (1.36 V). However, given that the CER is a two-electron process while the OER is a four-electron process, in terms of kinetics, the kinetics of the CER are significantly faster than those of the OER. Consequently, the CER is observed to be dominant above 1.36 V.58 An optimal anode material must have several key characteristics: (1) exceptional performance in chlorine evolution; (2) the ability to suppress oxygen evolution; (3) a long service life combined with a cost-effective price.61–63

Fig. 3 illustrates the evolution of chlorine evolution reaction (CER) catalyst materials, highlighting a shift from reliance on precious metals to a gradual decrease in their use. This transition culminates in the emergence of innovative non-precious metal catalysts, reflecting significant advances in CER catalyst research in recent years. During the 1950s, platinum (Pt) and magnetite (Fe3O4) were employed as anode materials, but their use declined over time due to high costs and limitations in current density. By the 1970s, graphite had become the dominant material in the industry due to its high electrical conductivity and accessibility. Nevertheless, graphite's stability was inadequate under the demanding conditions of chlor-alkali processes, making it unsuitable for long-term industrial use. Moreover, graphite was incompatible with membrane-based cells, as its particles tended to detach from the anode and obstruct the membrane. Subsequently, insoluble anodes were mainly produced using Pt and Ti, and it was found that substituting part of the Pt with Ir resolved stability and cost issues. Consequently, Pt–Ir/Ti anodes became the first titanium-based anodes to be commercialized on a large industrial scale. In 2000, Trasatti64 developed the first-generation of dimensionally stable anodes (DSAs), which consisted of 50% RuO2 and 50% TiO2 deposited on a titanium substrate. They exhibited improved CER activity, stability, and dissolution resistance compared to Pt–Ir/Ti electrodes. Additionally, the substitution of Ru-based coatings for Pt significantly reduced the cost of the anode. In 1967, Beer65 enhanced the DSA by reducing the Ru content from 50% to 30%, further reducing the reliance on precious metals. In 1986, the incorporation of IrO2 further stabilized RuO2 and improved the overall stability of the DSA. However, these materials have gradually been phased out of the market due to their limitations, such as short lifespans, coating delamination, poor electrochemical performance, and high production costs.66,67 In the 2020s, to further improve the activity, selectivity, and stability of MMO-based catalysts, another approach has been to incorporate new components from MMOs as CER electrocatalysts, resulting in the development of diverse new catalysts, including high-entropy oxides (HEOs), non-noble metal oxides (e.g., CoSb2Ox), and atomically dispersed catalysts (ADCs).68–71


image file: d5qm00299k-f3.tif
Fig. 3 Timeline of the development of CER catalysts.

The DSA electrode serves as the quintessential example of a composite material known as an enamel-electrocatalytic-semiconductor (CESC).71 The standard coating consists of titanium dioxide (TiO2) as the primary material, doped with ruthenium dioxide (RuO2), usually in a molar ratio of Ru[thin space (1/6-em)]:[thin space (1/6-em)]Ti = 30[thin space (1/6-em)]:[thin space (1/6-em)]70. While TiO2 is inherently non-conductive, RuO2 acts as an active electrocatalyst, greatly improving the electrode's performance. Although RuO2 alone has weak adhesion, the similar ionic radii of Ru4+ and Ti4+ allow doped RuO2 to form a co-crystalline structure with TiO2. Within this structure, Ru4+ substitutes for some Ti4+ and imparts semiconducting properties to the coating. Thus, developing the DSA electrode has been of great significance to the chlor-alkali industry. Early DSA electrodes included not only Ti/RuTiOx but also Ti/RuMnOx and Ti/RuSnOx, all of which are categorized as binary Ti/MMO electrodes (Fig. 4a). The first application of the relevant metals in DSA electrodes as CER catalysts is shown in Fig. 4b.


image file: d5qm00299k-f4.tif
Fig. 4 (a) Illustration of the binary Ti/MMO electrode structure.32 (b) Timeframes for using metals in DSA electrodes for CER catalysis.72

DSA-coated titanium anodes have been widely adopted in various industries, such as the chlor-alkali industry, wastewater treatment, metal electroplating, water electrolysis for hydrogen production, and cathodic protection, demonstrating high current efficiency and low energy consumption in the electrolytic production of NaClO. Nevertheless, their lifespan is typically limited to 2–3 months. The limited service life of early DSA electrodes can be explained by the following factors.73–76


3.3.2.1 Dissolution of RuO2. In a membrane-less electrolytic cell, the highly oxidizing sodium hypochlorite (NaClO) produced during electrolysis dissolves RuO2, transforming it into soluble high-valence compounds. At an anode potential of 1.387 V (vs. SHE), RuO2 is oxidized to the high-valence RuO4, which subsequently decomposes into RuO2·xH2O and O2 in the solution. These decomposition products can be deposited on the electrode surface or inside the electrolytic cell, resulting in the formation of black materials.
3.3.2.2 Anodic oxygen evolution. The RuO2–TiO2 coating is non-stoichiometric, comprising oxygen-deficient oxides (e.g., RuO2-x and TiO2-x), which are regarded as active sites for chlorine evolution. Due to the presence of oxygen vacancies, these vacancies become increasingly filled during anodic oxygen evolution, leading to a rise in electrode overpotential and resulting in passivation.
3.3.2.3 Substrate passivation. During electrolysis, reactive oxygen species or hydroxyl radicals can form on the anode surface, penetrate and adsorb onto the titanium substrate to create a low-conductivity TiO2 layer, leading to anode passivation.
3.3.2.4 Coating delamination. The Ru–Ti anode coating fabricated via conventional thermal decomposition usually displays a dense, bulky “mud-crack” morphology. As the anode operates, gas evolution reactions may persistently occur on the electrode surface, forming bubbles. When a bubble collapses, the surrounding liquid is continuously compressed, creating a micro-jet phenomenon where droplets impact the coating surface at velocities ranging from 100 to 500 m s−1. Such high-velocity impacts significantly damage the “mud-crack” structure of the coating, allowing the electrolyte to readily infiltrate the Ti substrate through these large cracks, resulting in Ti substrate corrosion and coating delamination.77
3.3.3 Summary of typical electrode materials. In addition to noble metal oxides, some non-noble metal materials also have performance advantages that cannot be ignored. Researchers have discovered that doping with transition metals not only lowers the cost of catalysts but also enhances the activity of the chlorine evolution reaction. Additionally, it improves the selectivity of Cl2 by adjusting the anode's composition and modulating its structure–activity relationship. For example, Saha et al.78 investigated the CER selectivity of first-row transition metal-doped RuO2 (such as Cu, Zn, Ni, Co, Fe, etc.) through DFT-based computational and experimental studies. The results demonstrated that doping RuO2 with elements that possess more d-electrons than Ru tends to reduce the binding strength of OER intermediates (e.g., HO, O, and HOO), thereby increasing the OER overpotential and providing more active sites for the CER. Moreover, the impact of doping on the binding strength of CER intermediates (ClO) is less significant, leading to enhanced CER selectivity. These findings offer novel insights into the design of efficient CER electrocatalysts. Furthermore, catalyst materials, including nanoporous materials and heterojunction composites, exhibit distinctive structural advantages.79,80 They can provide more active sites during the electrolysis reaction process, significantly enhancing the catalytic reaction activity. This work summarizes some of the electrode materials reported to date and provides a summary of the key data related to them, as shown in Table 1.
Table 1 Summary of typical electrode materials
Electrode materials ClER potentials (at 10 mA cm2) (V) Tafel slope (mV dec−1) RCT (Ω) Service lifetime (h) Cl2 selectivity (%) Mass activity (A gpt−1) Ref.
Ti/RuO2@TiO2–NTs 1.42 78 150 91 0.478 81
RuOx/2D TiOx 1.46 43.6 Smaller than RuO2 210 96.5 0.650 82
Ru0.3Sn0.35Ti0.35O2−x 1.32 47 640 83
RuO2–TiO2/TNTs/Ti 1.1 132 84
NF-RuO2–TiO2/Ti 1.04 (current density is 2 mA cm2) 41 0.87 85
Ru–Sn–Ti 1.11 40 2.48 240 95.7 86
RuO2–TiO2 NBs–Ti 1.13 (current density is 50 mA cm2) 48 408 57.1 87
Cu@FeN/Ti 1.61 1.58 75 60
Pt/F-CNTs 1.98 32.3 >98 3.81 88
Pt–C2N2 1.41 43 4.01 24 >99.6 2.12 89
Ti/SnO2–Sb2O5 180 90
Ti/RuO2–Sb2O5–SnO2 1.2 250 91
Pt/NPC 1.39 48 3.2 10 99 2.21 92
HPC 1.45 131 16.4 12 93
Ag0.15Ru0.85O2 1.47 68.4 50.1 150 86.5 94
CoSbyOx–Ti 1.89 178 240 92.2 95
Ru–S–TiO2 1.43 55 30.2 100 >95 2.31 96


3.4 Strategies for enhancing electrode performance

Given the challenges encountered by conventional electrode materials in real-world applications, several strategies can be implemented to significantly boost the performance of these materials, enhance their durability, and reduce production costs. These improvements are essential for advancing the development of sodium hypochlorite electro-synthesis technology.97
3.4.1 Developing multifunctional metal coatings. Multifunctional metal coatings represent a key strategy for enhancing electrode performance. In contrast to single-component coatings, Ti/MMO electrodes demonstrate better electrocatalytic performance and protect the Ti substrate from corrosion. In Ti/MMO electrodes, MMO consists of noble metal oxides (e.g., RuO2, IrO2) and inert components (e.g., TiO2, ZrO2), sometimes supplemented with common metal oxides (e.g., SnO2, Sb2O3) to boost electrocatalytic performance.98–102

Throughout history, the metallic elements used in DSA electrode coatings have varied depending the specific time period. Platinum group metal oxides (e.g., RuO2, IrO2, Rh2O3) have found extensive applications due to their excellent electrocatalytic activity and durability,103–109 and the properties of various platinum group metals are detailed in Table 2.

Table 2 Properties of platinum group metals
Element name Price Corrosion resistance Chlorine overpotential Oxygen overpotential Stability
Ruthenium (Ru) Relatively low Very strong Relatively low Relatively high High
Rhodium (Rh) Extremely high Extremely strong High Low High
Palladium (Pd) Moderate Slightly weak Relatively low Relatively high Slightly weak
Osmium (Os) Relatively high High Relatively high Relatively high High
Iridium (Ir) Relatively high Extremely strong Relatively high Relatively low Extremely high
Platinum (Pt) Relatively high Extremely strong Relatively high Relatively high High


The table shows that PdO exhibits a low chlorine evolution potential and a high oxygen evolution potential, suggesting its potential as an excellent catalyst material. Nevertheless, its corrosion resistance and stability are mediocre, and the practical lifespan of Pd-coated coated titanium electrodes is short, which hinders their effective utilization.110,111 RuO2 and IrO2 exhibit superior catalytic performance for chlorine and oxygen evolution, leading to extensive applications.112 Additionally, research has revealed that IrO2 coatings have far greater corrosion resistance than RuO2. Ir demonstrates a higher current efficiency in the electrolysis of low-concentration brine compared to Ru, achieving efficiencies of up to 90%. Incorporating a certain amount of Ir into the electrode substantially extends its lifespan (approximately 2–3 years, depending on the Ir content).91,113–115

At the same time, in the multifunctional design of electrode coatings, and considering the limited availability and high cost of noble metals, researchers have investigated the integration of widely available metal oxide semiconductors (such as oxides of Sn, Sb, Mn, Co, and others) to enhance the performance of the coatings.72,91 In particular, SnO2 can stabilize ruthenium-based/titanium-based coatings and improve their electrochemical activity.116–118 Incorporation of Sb2O3 improves the conductivity of SnO2, resulting in a more compact coating and mitigating the oxidation of the Ti substrate.90,119,120 This approach lowers the cost of Ti/MMO electrodes while maintaining their superior electrocatalytic performance.

Zhang et al.121 synthesized SnO2/TiO2 composites using a chemical co-precipitation approach. Fig. 5a shows the fabrication of the SnO2/TiO2 composite electrode and the electrocatalytic degradation process of methylene blue pollutants. The electrochemical properties of the electrode materials were evaluated at various calcination temperatures. The impedance test results indicate that the composite material calcined at 500 °C has the lowest impedance and the highest electron transfer rate (Fig. 5b), which is most favorable for the electrocatalytic degradation of methylene blue. Fig. 5c presents the degradation efficiency and reaction kinetics of the SnO2/TiO2 composite material for methylene blue. When calcined at lower temperatures, the incomplete growth of titanium dioxide particles on the surface results in limited degradation performance. After 240 min of electrolysis, the composite material calcined at 500 °C exhibited the highest degradation efficiency (96.6%) and the largest reaction rate constant (0.01316 min−1), demonstrating the fastest electrocatalytic decolorization rate and the highest catalytic capacity. The stability of the composite material was studied. Fig. 5d illustrates the change in the degradation efficiency of the SnO2/TiO2 composite material for methylene blue as a function of the number of cycles. After five degradation cycles, the composite material calcined at 500 °C maintained a degradation efficiency of over 93.3%, with a decrease of only 3.3%, indicating excellent stability.


image file: d5qm00299k-f5.tif
Fig. 5 (a) Schematic illustration for fabricating the SnO2/TiO2 composite electrodes and the electrocatalytic degradation process. (b) AC impedance diagram and the inset is the partially enlarged image and the fitted equivalent circuit model of the SnO2/TiO2 composite. (c) Curve of the degradation of methylene blue and the fitting curve of the first-order reaction kinetics for electrocatalytic degradation of SnO2/TiO2 composite. (d) Cyclic testing experiment on the degradation of methylene blue by SnO2/TiO2 composite. (e) SEM image of the Ti/SnO2–RuO2 anode. (f) Accelerated stability test of the Ti/SnO2–RuO2 anode at a current density of 500 mA cm2 in 0.5 M H2SO4 solution.

Chen et al.122 successfully prepared a Ti-based electrode with a SnO2–RuO2 active layer through thermal decomposition and used it for the effective electrocatalytic degradation of ACG. The SEM images presented in (e) reveal that Ti/SnO2–RuO2 exhibits a distinctive “mud-crack” structure, characterized by a high specific surface area. This structure increases the availability of active sites for electrocatalytic oxidation, thereby enhancing the degradation of organic wastewater. Fig. 5f illustrates the accelerated stability test of the Ti/SnO2–RuO2 electrode in a 0.50 M H2SO4 solution at a current density of 500 mA cm2 using direct current. The prepared electrode demonstrated a significantly enhanced stability lifespan of 40 hours before deactivation, representing a remarkable improvement compared to the Ti/SnO2 electrode, which deactivated after just 4.54 hours. This indicates that the incorporation of RuO2 has substantially improved the electrochemical stability of the Ti/SnO2–RuO2 electrode. These multi-component coating electrodes demonstrate efficient electrocatalytic performance and extended service lifespan in practical applications.

3.4.2 Incorporating an intermediate layer between the substrate and the coating. Researchers have demonstrated that incorporating a conductive intermediate layer between the electrode coating and the substrate effectively reduces substrate passivation and extends the lifespan of the electrode. The materials for the intermediate layer primarily consist of precious metals and their oxides (e.g., Pt, Ir), non-noble metal oxides (e.g., Sn, Sb, Mn), and combinations thereof. While the incorporation of an intermediate layer has been shown to enhance the lifespan, it concomitantly increases electrode resistance, leading to a reduction in chlorine evolution performance.123–126 Thus, the thickness and choice of elements for the intermediate layer are critical factors in electrode design.

Lang et al.127 synthesized the layered double hydroxide (LDH) electrocatalyst material (FeCo–LDH) via magnetron co-sputtering technology, with the reaction steps illustrated in Fig. 6a. The distinctive structure of the FeCo–LDH catalyst features a highly textured surface rich in nanopores, with nanosheets measuring approximately 5 nm in thickness (see Fig. 6b). This unique architecture significantly increases the active reaction area and reveals numerous active sites. Second, the FeCo–LDH catalyst demonstrates superior electrocatalytic performance, with a low overpotential of merely 300 mV at 10 mA cm−2 (Fig. 6c) and a Tafel slope of 43 mV dec−1 (Fig. 6d), suggesting rapid reaction kinetics and excellent electrocatalytic activity during catalysis. Furthermore, the FeCo–LDH catalyst demonstrates excellent stability during prolonged performance testing, showing no significant loss of activity (see Fig. 6e). This further emphasizes its potential reliability for practical use in real applications.


image file: d5qm00299k-f6.tif
Fig. 6 (a) Scheme for the construction of the FeCo–LDH catalyst layer. (b) AFM image and height profile of a nanosheet. (c) LSV curves, (d) Tafel slope measurement, and (e) potential variations with time for the Fe150Co50–LDH electrocatalyst at different current densities. (f) XRD diffraction patterns of RuO2–TiO2/TNTs/Ti and RuO2–TiO2/Ti electrodes. (g) SEM microphotographs of RuO2–TiO2/TNTs/Ti and RuO2–TiO2/Ti. (h)–(j) Polarization curves, cyclic voltammograms and accelerating corrosion of the materials affected by the presence or absence of the TNT interlayer.

Cao et al.84 fabricated a new RuO2–TiO2 composite electrode (Ti/TNTs/RuTiOx electrode) by employing TiO2 nanotube arrays (TNTs) as an intermediate layer via a thermal decomposition method. The thermally treated TNT intermediate layer provides a high specific surface area and strong adsorption capability, significantly enhancing the crystallinity of the ruthenium–titanium coating. This process minimizes surface cracks, effectively prevents electrolyte infiltration into the titanium substrate, and offers exceptional protection for the titanium substrate. Furthermore, the directly grown structure guarantees robust adhesion between the active components and the titanium substrate. Additionally, the presence of titanium nanotubes (TNTs) increases the number of active sites, leading to an electrode lifespan that is approximately double that of conventional Ti/RuTiOx electrodes. As illustrated in Fig. 6f, under identical preparation conditions, the diffraction peaks for RuO2 grains (2θ = 28.1°, 2θ = 35.1°) and rutile-phase TiO2 crystals (2θ = 27.4°, 36.1°) in the RuO2–TiO2/TNTs/Ti electrode exhibit significantly greater intensity compared to those in the RuO2–TiO2/Ti electrode. This observation indicates that the TNT intermediate layer enhances the crystallinity of both RuO2 and rutile-phase TiO2. Fig. 6g presents SEM images of the RuO2–TiO2/TNTs/Ti and RuO2–TiO2/Ti electrodes. Under identical preparation conditions, the electrode with the TNT intermediate layer exhibits a substantial reduction in “mud-crack” structures on its surface, effectively reducing the passivation of the Ti substrate. Fig. 6h–j illustrate the polarization curves, cyclic voltammetry (CV) profiles, and lifespan charts of the materials affected by the nanostructure with or without the TNT intermediate layer. In conclusion, the addition of the TNT intermediate layer significantly enhances both the catalytic activity and longevity of the RuO2–TiO2 composite electrode. This improvement is particularly notable in environments containing chloride, where the electrode demonstrates remarkable corrosion resistance and stability.

3.4.3 Enhancing electrode fabrication techniques. Thermal decomposition has become the most widely used method of electrode production due to its well-established technology, straightforward operation, and low equipment expenses. The process involves dissolving metal salts in alcohol-based organic solvents to create a precursor solution. This solution is then subjected to solvent evaporation and high-temperature calcination, a procedure that is repeated multiple times to produce the electrode coating. Nevertheless, electrodes produced by thermal decomposition often exhibit significant “sludge cracking” on the surface, which facilitates oxygen penetration into the substrate and leads to electrode passivation.128 Consequently, researchers have attempted to improve the technology to achieve high-quality catalytic coatings, with key optimization methods including the sol–gel process, magnetron sputtering, hydrothermal synthesis, electrodeposition, chemical vapor deposition, and ultrasonic spray pyrolysis.129–133
3.4.3.1 Sol–gel method. The sol–gel method enables the preparation of stoichiometric oxide components, with oxide electrodes uniformly distributed on the titanium substrate surface. However, it involves extensive reagent use and long processing times. At low sol concentrations, problems like weak bonding strength and delamination of the surface active layer may occur.134–137 Francisca et al.138 employed the sol–gel method to fabricate a ternary mixed metal oxide Ti/RuO2–ZrO2–Sb2O3 electrode (flowchart shown in Fig. 7a) and performed accelerated aging life tests on the electrode (shown in Fig. 7b). This approach can improve the catalytic activity of the electrode and extend its service life to 206 hours. Moreover, Liu et al.134 discovered that, unlike the thermal decomposition method, the sol–gel method facilitates a uniform distribution of active components, such as IrO2 and RuO2, within inactive components such as ZrO2. This results in anodes with unique microstructures and increased active surface areas, which significantly enhances electrocatalytic activity.
image file: d5qm00299k-f7.tif
Fig. 7 (a) Flow chart of the DSA electrode preparation using the modified Pechini method. (b) Behavior of the potential of a ternary mixture DSA electrode (Ti/RuO2–ZrO2 doped with Sb2O5) during an accelerated life test. (c) Cyclic voltammetry for anodes annealed at 600 °C, 100 mV s−1, and 1 mol L−1 H2SO4. (d) Model of the layer-by-layer degradation mechanism for the anode, (e) cyclic voltametric behavior of Ti/Ni–Sb–SnO2 anodes with different Ni doping concentrations in the precursor solution. (f) The calculation models of SnO2 doped with 12.5 at% Sb atoms, 12.5 at% Sb and 6.25 at% Ni atoms and 12.5 at% Sb and 12.5 at% Ni atoms. (g) Fermi energy level and work function of Ti/Ni–Sb–SnO2.

3.4.3.2 Magnetron sputtering method. Magnetron sputtering enables the fabrication of dense and uniform coatings, although the equipment is intricate and expensive.127,139,140 Yan et al.141 employed magnetron sputtering to finely tune the microstructure of the coating, fabricating Sb-doped Ti/SnO2 electrodes. The electrodes fabricated using this method feature a microrod structure, significantly differing from the porous or mud-cracked surfaces produced by other techniques. The microrod structure increases the specific surface area of the anode and the number of active sites, thereby reducing the actual current density and improving the density and adhesion of the coating. Additionally, magnetron sputtering can create coatings with uniform and well-crystallized structures, enhancing the anode's conductivity and stability. This unique structure significantly enhances the performance of the anode, resulting in an oxidation overpotential of approximately 2.2 V for the microrod-structured SnO2 anode in a 1.0 mol L−1 sulfuric acid solution (as illustrated in Fig. 7c). This performance is comparable to, or even superior to, that of Sb-doped SnO2 anodes prepared by conventional methods. Moreover, a layer-by-layer degradation mechanism model for the anode has been introduced in the literature (depicted in Fig. 7d), which effectively explains the periodic oscillations observed during accelerated life testing of the microrod-structured anode and provides important theoretical foundations for understanding the anode failure mechanism.
3.4.3.3 Ultrasonic spray pyrolysis method. The ultrasonic spray pyrolysis technique is advantageous due to its low equipment cost and the ease with which it allows for the control of decomposition conditions. Moreover, the generation of ultrafine droplets generated through ultrasonic atomization has been demonstrated to refine the microstructure of the film surface, thereby enhancing the performance of the film.142–144 Chen et al.145 employed spray pyrolysis to fabricate Ti/Ni–Sb–SnO2 anodes, demonstrating excellent density. In contrast to conventional thermal decomposition or spin-coating methods, the surface displayed no noticeable cracks, enhancing the anode's durability and service life. Fig. 7e illustrates the cyclic voltammetry behavior of Ti/Ni–Sb–SnO2 electrodes with different nickel doping concentrations in 0.5 M sulfuric acid solution. As shown, the oxygen evolution onset potential for the undoped Ti/Sb–SnO2 electrode is around 2.0–2.4 V (vs. NHE), while nickel doping raises the onset potential to above 2.4 V, with a slight increase as the nickel concentration increases (reaching up to 2.5 V). This finding indicates that nickel doping enhances the anode's oxidation capacity, thereby facilitating the enhanced efficiency of electrochemical treatment for organic pollutants. To further understand the impact of nickel doping on the oxygen evolution onset potential, the team calculated the work function of Sb–SnO2 with nickel atomic ratios ranging from 0 to 12.5 at%. The computational model is illustrated in Fig. 7f and calculations were performed using density functional theory (DFT, Fig. 7g). The results indicate that an increase in the nickel doping concentration results in a higher work function of SnO2. This, in turn, aids in elevating the anode's oxidation potential and strengthening its oxidation capability. This finding offers further theoretical backing for the experimental results.

Furthermore, a variety of synthesis methods exist, each exhibiting distinct strengths and weaknesses. These methods are constantly being optimized and explored. For example, hydrothermal synthesis can produce electrodes with distinctive structures by controlling the concentration of hydrochloric acid and has demonstrated superior performance.85 Electrodeposition has been demonstrated to yield electrodes with strong adhesion, uniform thickness, and crack-free surfaces. This methodology also enables precise tuning of the coating's structure and performance. However, it requires significant amounts of metal precursor salts, especially for the deposition of precious metals, resulting in increased costs and environmental concerns such as water contamination.146 Chemical vapor deposition (CVD) is a widely used technique for the fabrication of high-purity, high-density films. This methodology allows for the creation of gradient concentrations or variable compositions by modifying the gas phase composition. Nonetheless, the high cost of gas phase precursors limits its widespread use in DSA electrode synthesis.147

4. Efficacy of hypochlorite solution

The production capacity of an electrolytic sodium hypochlorite solution system is typically evaluated based on the chlorine content available per unit of time. However, seawater contains various salts in addition to NaCl, including Ca2+ and Mg2+. Calcium and magnesium ions form calcium carbonate and magnesium carbonate on the surface of the cathode, where they adhere firmly. These deposits increase the cell voltage and decrease the yield of sodium hypochlorite. This directly impacts the energy consumption of the entire chlorine generation system.148–151 At the same time, different electrode materials and electrolytic cell designs can affect current efficiency, resulting in the actual yield not meeting the theoretical value. Additionally, the conditions under which reactions occur also play a critical role in determining the stability of the solution.

Fig. 8 illustrates the electrochemical reactions occurring on the surface of the cathode electrocatalyst within the reverse electrodialysis (RED) system, as well as the mechanisms behind the formation of inorganic scaling on different electrocatalysts. Diagram a illustrates the nucleation and growth process of scaling (Mg(OH)2 deposition) on bulk Pt cathode electrocatalysts. On the cathode surface, water undergoes electrolysis to produce hydroxide ions (OH). These OH ions react with multivalent cations in seawater (e.g., Mg2+ and Ca2+) to form hydroxides (such as Mg(OH)2), which progressively deposited on the electrocatalyst surface. The kinetics of crystal growth in precipitates have a significant effect on the properties of cathode electrocatalysts. These properties include the scaling factor (whether at the nanoscale or bulk level), porosity, crystallinity, and surface chemistry. Consequently, these factors can lead to a reduction in electrocatalytic activity and charge transfer efficiency. Diagram b displays the morphological variations in inorganic scaling formed by different cathode electrocatalysts (e.g., bulk Pt, nano Pt/C, heat-treated carbon, untreated carbon, and acid-air plasma-treated carbon) within the RED system, highlighting the influence of electrocatalyst design on scaling behavior. Hence, therefore, the selection of suitable electrode materials can effectively mitigate the effects of precipitates on electrolysis efficiency.


image file: d5qm00299k-f8.tif
Fig. 8 (a) Schematic of the electrochemical reactions at the cathodic electrocatalyst during real-time RED evaluation. (b) Schematic of the scale formation on bulk Pt, nano Pt/CB/CC, HT CB/CC, bare CB/CC, and AA CB/CC.152

4.1 Measurement of available chlorine content

Numerous methods exist for measuring the available chlorine content in solutions, including the test strip method, iodometric titration, colorimetry, and gas chromatography. Nonetheless, these methods are either labor-intensive and time-consuming or require high reagent consumption and costly equipment.153 In contrast, UV-visible spectrophotometry provides notable advantages in sensitivity and accuracy, allowing chlorine content to be measured quickly and accurately using a simple procedure. It offers swift analysis, user-friendly operation, excellent reproducibility, and high analytical precision, contributing to its widespread application in water quality assessment across both laboratory and industrial settings. A schematic diagram of the available chlorine method is shown in Fig. 9.
image file: d5qm00299k-f9.tif
Fig. 9 Schematic diagram of the method for measuring the available chlorine content.
4.1.1 Test strip method. The test strip method involves dipping a chemically treated test strip into a water sample and assessing the available chlorine concentration through color changes. This method is very simple to use, ideal for quick on-site screening, and highly cost-effective, but it has low precision and can only give a rough concentration range. Moreover, ambient humidity and temperature greatly affect the test strip method.
4.1.2 Iodometric titration method. The principle of iodometric titration involves the reaction of available chlorine with potassium iodide in an acidic solution to produce iodine, which is subsequently titrated with a sodium thiosulfate standard solution to calculate the available chlorine content. This method is mature, relatively simple, applicable, and appropriate for routine laboratory analysis. However, it requires specialized chemical reagents and titration apparatus, requires subjective factors that may influence the operator's advanced technical skills, and requires determination of the titration endpoint.
4.1.3 Iodometric photometric method. The iodometric photometric method entails the reaction of available chlorine with potassium iodide in an acidic solution to release iodine, which is then measured colorimetrically using photometry to determine the available chlorine content. This method is applicable for determining high and low concentrations of available chlorine, integrating the high sensitivity and selectivity of photometry to yield more accurate measurements. However, it requires equipment such as a colorimeter, which increases the cost, and requires more extensive sample preparation, which can be affected by sample color or turbidity.154
4.1.4 Gas chromatography method. Gas chromatography effectively isolates chloride components from the sample using a chromatographic column. It measures the concentration of each component with a detector, such as an electron capture detector (ECD) or a thermal conductivity detector (TCD). Finally, it determines the available chlorine content by referencing a standard curve or applying a correction factor. The thermal conductivity detector operates by measuring the difference in thermal conductivity between the separated components and the carrier gas. Variations in gas components and concentrations lead to changes in the temperature and resistance of the thermal sensor, with the detection signal magnitude proportional to the component concentration. This method offers exceptional sensitivity and selectivity, along with simple operation and reliable results. It generates more representative analytical data, but requires specialized equipment and trained personnel.155
4.1.5 UV spectrophotometry method. The principle of UV spectrophotometry involves the reaction of chloride ions with specific reagents to generate colored products that display characteristic absorption peaks in the UV-visible spectrum. The absorbance of these colored products is proportional to the concentration of chloride ions, allowing the chloride ion content to be accurately determined from the absorbance and a standard curve. This detection method offers several advantages, including high sensitivity for identifying low concentrations of chloride ions, simple experimental procedures that are ideal for rapid laboratory testing, and the ability to produce highly accurate results that allow accurate quantitative analysis based on a standard curve.156–158

4.2 Enhancing the stability of the solution

4.2.1 pH effects. Under alkaline conditions (pH > 7), the monovalent chlorine present in sodium hypochlorite solution primarily exists as hypochlorite ions (ClO), which is the most stable form and demonstrates a slow decomposition rate. Additionally, a higher OH concentration in the solution promotes the reversible hydrolysis reaction of sodium hypochlorite to shift to the left, minimizing the degradation of available chlorine. Nevertheless, an excessively high pH can reduce the bactericidal effectiveness of sodium hypochlorite solution.159–162 To ensure optimal stability and bactericidal efficacy, the pH of the disinfectant is typically maintained between 7 and 8 by adjusting the free alkali content.
4.2.2 Light effects. Due to its strong oxidizing properties, sodium hypochlorite is subject to natural decomposition at room temperature, a process that is significantly accelerated by exposure to light. The reaction is represented as image file: d5qm00299k-t1.tif. According to experimental data, cumulative sunlight exposure exceeding 20 hours under natural conditions reduces the available chlorine content to below 10%. With increasing temperature and light intensity, the motion of organic chlorine molecules accelerates, and the activation energy for decomposition decreases, further hastening the decomposition rate. This process results in precipitation during storage or use, ultimately affecting the available chlorine content.163 Hence, light protection is essential during the production, transportation, and storage of sodium hypochlorite solution. For optimal storage, it is advisable to use opaque or dark packaging materials, including rigid PVC, polyethylene, steel containers, or brown bottles.
4.2.3 Solution concentration effects. The concentration of hypochlorite solution is inversely correlated with the tendency of the decomposition reaction. Higher chlorine content in sodium hypochlorite solution increases the likelihood of disproportionation reactions, resulting in accelerated decomposition and reduced stability.164,165 Low concentrations can effectively slow the decomposition rate and improve stability. In practice, to balance performance and cost effectiveness, the initial concentration of sodium hypochlorite solution should be reduced, or the solution should be diluted to a lower concentration for storage. This method effectively slows the rate of decomposition of the sodium hypochlorite solution, minimizing the loss of active ingredients and increasing its stability during extended storage. Moreover, lowering the concentration reduces safety risks during storage by preventing potential chemical reactions or leaks due to high concentrations.
4.2.4 Stabilizer effects. To enhance the stability of sodium hypochlorite, it is common practice to incorporate an appropriate amount of stabilizer into its aqueous solution. The choice of stabilizers must ensure that they do not diminish the disinfecting efficacy of sodium hypochlorite. Typical stabilizers include inorganic compounds (e.g., sodium silicate, boron nitride, sodium bromide, disodium EDTA) and organic compounds (e.g. β-cyclodextrin, cellulose, sulfamic acid, phenanthroline). Research shows that, under the same experimental conditions, sodium hypochlorite solutions containing stabilizers effectively maintain a stable level of available chlorine compared to a control group without stabilizers. This significantly extends both the shelf life and half-life of the solution, thereby enhancing its practicality and economic value.

In addition, factors such as temperature, storage conditions, metal ions and contaminants, and storage duration can all affect the available chlorine content in sodium hypochlorite solutions. It is essential to maintain optimal conditions to preserve the available chlorine content as much as possible and improve the stability of sodium hypochlorite solution. A schematic diagram of methods to increase the stability of sodium hypochlorite solution is shown in Fig. 10.


image file: d5qm00299k-f10.tif
Fig. 10 Schematic diagram of methods to enhance the stability of sodium hypochlorite solution.

5. Conclusions and prospects

Electrolytic synthesis of sodium hypochlorite solution is an efficient, safe, and environmentally friendly production method with extensive applications in water treatment, healthcare, and other sectors. As global attention toward environmental protection and public health continues to grow, the electrolytic synthesis of sodium hypochlorite solution technology, with its distinct advantages, can effectively mitigate pollutant emissions and safety risks associated with traditional chemical synthesis. This technology is progressively replacing other synthesis methods, emerging as a new industry trend, delivering substantial economic and environmental benefits, and is poised to play a significant role in more emerging areas.

Significant advancements have been made in the electrolytic synthesis of sodium hypochlorite solution technology, particularly in the areas of equipment design, process optimization, and application development. The development of electrode materials has advanced significantly from unitary to diversified types, with ongoing research expected to produce new electrode materials featuring high current efficiencies, extended lifespans, and low costs. At the same time, ensuring the stability of the sodium hypochlorite solution is imperative and cannot be ignored. Strategies such as adjusting storage conditions and incorporating external additives have proven to be effective in ensuring the stability of the solution.

The electrolytic synthesis of sodium hypochlorite solution technology is poised for continued evolution. First, advances in technology and continued innovation will lead to more efficient and energy-saving production processes while improving product purity and stability. Second, the diversification of market demand indicates that electrolytic synthesis of sodium hypochlorite solution technology has tremendous market potential and will play an increasingly important role in areas such as food processing, textiles and renewable energy, providing new momentum for industry growth. Lastly, the national focus on environmental safety and the push for environmental policies will steer the industry toward greener and more sustainable development. Achieving a mutually beneficial outcome for both economic and environmental interests is possible through technological innovations in multiple domains, including energy conservation and emission reduction, resource recycling, green chemistry, and intelligent automation. Electrolytic synthesis technology fulfills the requirements for sustainable development and is poised to become the dominant process for sodium hypochlorite solution production.

In conclusion, the technology for electrolytic synthesis of sodium hypochlorite solution strongly supports current market needs, highlights significant environmental and economic benefits, and provides a solid foundation for future sustainable development. This technology is poised to drive industry advancements and expand the application of sodium hypochlorite solution into broader markets.

Data availability

Data will be made available on request.

Conflicts of interest

The authors declare no competing interests.

Acknowledgements

This work was financially supported by the Guangxi Natural Science Fund for Distinguished Young Scholars (2024GXNSFFA010008) and the National Natural Science Foundation of China (22469002).

References

  1. A. L. Severing, J. D. Rembe and V. Koester, et al., Safety and efficacy profiles of different commercial sodium hypochlorite/hypochlorous acid solutions (NaClO/HClO): antimicrobial efficacy, cytotoxic impact and physicochemical parameters in vitro, J. Antimicrob. Chemother., 2019, 74(2), 365–372 CrossRef CAS PubMed.
  2. G. A. De Queiroz and D. F. Day, Antimicrobial activity and effectiveness of a combination of sodium hypochlorite and hydrogen peroxide in killing and removing Pseudomonas aeruginosa biofilms from surfaces, J. Appl. Microbiol., 2007, 103(4), 794–802 CrossRef CAS PubMed.
  3. Y. Li, M. Yang and X. Zhang, et al., Two-step chlorination: A new approach to disinfection of a primary sewage effluent, Water Res., 2017, 108, 339–347 CrossRef CAS PubMed.
  4. M. Liu and I. Kasuga, Impact of chlorine disinfection on intracellular and extracellular antimicrobial resistance genes in wastewater treatment and water reclamation, Sci. Total Environ., 2024, 949, 175046 CrossRef CAS PubMed.
  5. M. Y. Xu, Y. L. Lin and T. Y. Zhang, et al., Chlorine dioxide-based oxidation processes for water purification: A review, J. Hazard. Mater., 2022, 436, 129195 CrossRef CAS PubMed.
  6. H. Xi, K. E. Ross and J. Hinds, et al., Efficacy of chlorine-based disinfectants to control Legionella within premise plumbing systems, Water Res., 2024, 259, 121794 CrossRef CAS PubMed.
  7. T. A. Ammar, K. Y. Abid and A. A. El-Bindary, et al., Comparison of commercial analytical techniques for measuring chlorine dioxide in urban desalinated drinking water, J. Water Health, 2015, 13(4), 970–984 CrossRef CAS PubMed.
  8. C. Y. Chang, Y. H. Hsieh and S. S. Hsu, et al., The formation of disinfection by-products in water treated with chlorine dioxide, J. Hazard. Mater., 2000, 79(1–2), 89–102 CrossRef CAS PubMed.
  9. T. A. Ammar, K. Y. Abid and A. A. El-Bindary, et al., Chlorine dioxide bulk decay prediction in desalinated drinking water, Desalination, 2014, 352, 45–51 CrossRef CAS.
  10. C. M. C. Andrés, J. M. P. D. L. Lastra and C. Andrés Juan, et al., Chlorine Dioxide: Friend or Foe for Cell Biomolecules? A Chemical Approach, Int. J. Mol. Sci., 2022, 23(24), 15660 CrossRef PubMed.
  11. N. Kishimoto, State of the Art of UV/Chlorine Advanced Oxidation Processes: Their Mechanism, Byproducts Formation, Process Variation, and Applications, J. Water Environ. Nanotechnol., 2019, 17(5), 302–335 CrossRef.
  12. L. V. Ashby, R. Springer and M. B. Hampton, et al., Evaluating the bactericidal action of hypochlorous acid in culture media, Free Radicals Biol. Med., 2020, 159, 119–124 CrossRef CAS PubMed.
  13. S. B. Young and P. Setlow, Mechanisms of killing of Bacillus subtilis spores by hypochlorite and chlorine dioxide, J. Appl. Microbiol., 2003, 95(1), 54–67 CrossRef CAS PubMed.
  14. Y. Jiang, Y. Qiao and R. Jin, et al., Application of chlorine dioxide and its disinfection mechanism, Arch. Microbiol., 2024, 206(10), 400 CrossRef CAS PubMed.
  15. S. Fukuzaki, Mechanisms of Actions of Sodium Hypochlorite in Cleaning and Disinfection Processes, Biocontrol Sci., 2006, 11(4), 147–157 CrossRef CAS PubMed.
  16. D. Wang, X. Chen and J. Luo, et al., Comparison of chlorine and chlorine dioxide disinfection in drinking water: Evaluation of disinfection byproduct formation under equal disinfection efficiency, Water Res., 2024, 260, 121932 CrossRef CAS PubMed.
  17. W. T. Cunningham and A. Y. Balekjian, Effect of temperature on collagen-dissolving ability of sodium hypochlorite endodontic irrigant, Oral Surg., Oral Med., Oral Pathol., 1980, 49(2), 175–177 CrossRef CAS PubMed.
  18. T. M. Fabian and S. E. Walker, Stability of sodium hypochlorite solutions, Am. J. Health-Syst. Pharm., 1982, 39(6), 1016–1017 CrossRef CAS.
  19. S. Wu, F. Shen and P. Zhao, et al., Designing a Bipolar Membrane Electrolyzer for NaCl Electrolysis to Produce High-Quality NaClO, J. Phys. Chem. C, 2023, 127(31), 15177–15184 CrossRef CAS.
  20. S. Ahmed and P. Foller, Responding to a Harsh Business Environment: A New Diaphragm for the Chlor-Alkali Industry, Electrochem. Soc. Interface, 2003, 12(4), 34–39 CrossRef.
  21. I. Garcia-Herrero, M. Margallo and R. Onandía, et al., Connecting wastes to resources for clean technologies in the chlor-alkali industry: a life cycle approach, Clean Technol. Environ. Policy, 2018, 20(2), 229–242 CrossRef CAS.
  22. I. Requena-Leal, M. Carvela and C. M. Fernández-Marchante, et al., On the use of chlor-alkali technology to power environmental electrochemical treatment technologies, Curr. Opin. Electrochem., 2024, 45, 101461 CrossRef CAS.
  23. X. Zhang, Y. Wang and S. Wei, et al., Assessing the chlorine metabolism and its resource efficiency in chlor-alkali industrial symbiosis - A case of Shanghai Chemical Industry Park, J. Cleaner Prod., 2022, 380, 134934 CrossRef CAS.
  24. L. C. Brée, A. Bulan and R. Herding, et al., Techno-Economic Comparison of Flexibility Options in Chlorine Production, Ind. Eng. Chem. Res., 2020, 59(26), 12186–12196 CrossRef.
  25. C. W. Bunn, L. M. Clark and I. L. Clifford, The constitution and formation of bleaching powder, Proc. R. Soc. London, Ser. A, 1935, 151(872), 141–167 CAS.
  26. D. M. Bibby and N. B. Milestone, The decomposition of high grade bleaching powder (calcium hypochlorite), J. Chem. Technol. Biotechnol., Chem. Technol., 1984, 34(8), 423–430 CrossRef.
  27. Y. Tang, H. Zhang and J. Yan, et al., Assessing the efficacy of bleaching powder in disinfecting marine water: Insights from the rapid recovery of microbiomes, Water Res., 2023, 241, 120136 CrossRef CAS PubMed.
  28. K. Sravan Kumar, S. Mateo and A. R. De La Osa, et al., Advancements in membrane-less electrolysis configurations: Innovations and challenges, Curr. Opin. Electrochem., 2025, 49, 101602 CrossRef CAS.
  29. J. Cho, T. L. Doan and S. Lee, et al., Porous transport electrodes for oxygen evolution reaction in proton exchange membrane water electrolysis -cells: Materials, designs, and diagnoses, Chem. Eng. J., 2025, 507, 160473 CrossRef CAS.
  30. J. Van Zee, R. E. White and A. T. Watson, Simple Models for Diaphragm-Type Chlorine/Caustic Cells: I. Dynamic Behavior, J. Electrochem. Soc., 1986, 133(3), 501–507 CrossRef CAS.
  31. L. Wang, K. Linowski and H. Liu, Scalable membrane-less microbial electrolysis cell with multiple compact electrode assemblies for high performance hydrogen production, Chem. Eng. J., 2023, 454, 140257 CrossRef CAS.
  32. J. Kim, M. Usama and K. S. Exner, et al., Renaissance of Chlorine Evolution Reaction: Emerging Theory and Catalytic Materials, Angew. Chem., Int. Ed., 2024, e202417293 Search PubMed.
  33. E. Jwa, W. Lee and S. Choi, et al., In situ disinfection and green hydrogen production using carbon-based cathodes in seawater electrolysis, Desalination, 2024, 580, 117580 CrossRef CAS.
  34. T. Prasert, F. Kurisu and P. Phungsai, Characterizing the precursors of byproducts formed by chlorine and chlorine dioxide disinfection using unknown screening analysis with Orbitrap mass spectrometry, Sci. Total Environ., 2025, 959, 178296 CrossRef CAS PubMed.
  35. B. Liu and J. Zhang, A ruthenium oxide and iridium oxide coated titanium electrode for pH measurement, RSC Adv., 2020, 10(43), 25952–25957 RSC.
  36. Y. Liu, L. Yu and L. Liu, et al., Tailoring the vanadium/proton ratio of electrolytes to boost efficiency and stability of vanadium flow batteries over a wide temperature range, Appl. Energy, 2021, 301, 117454 CrossRef CAS.
  37. T. Morishige, G. M. Haarberg and H. Gudbrandsen, et al., Effects of Composition and Temperature on Current Efficiency for Aluminium Electrolysis from Cryolite-Based Molten Alumina Electrolytes, ECS Trans., 2017, 77(11), 997–1002 CrossRef CAS.
  38. Y. Cheng, X. Shi and Y. Lv, et al., Effect of Electrolyte Temperature on Plasma Electrolytic Oxidation of Pure Aluminum, Metals, 2024, 14(6), 615 CrossRef CAS.
  39. A. M. Gheitaghy, H. Saffari and D. Ghasimi, et al., Effect of electrolyte temperature on porous electrodeposited copper for pool boiling enhancement, Appl. Therm. Eng., 2017, 113, 1097–1106 CrossRef CAS.
  40. S. Wang, Z. Xu and X. Wu, et al., Analyses and optimization of electrolyte concentration on the electrochemical performance of iron-chromium flow battery, Appl. Energy, 2020, 271, 115252 CrossRef CAS.
  41. M. Ghufron, Istiroyah and C. A. Perwita, et al., Influence of electrolyte concentration on static and dynamic Lead-Acid battery, J. Phys.:Conf. Ser., 2020, 1595(1), 012012 CrossRef CAS.
  42. X. Tu, Y. Ma and C. Yang, et al., Trace Amount of High-Concentration Electrolyte for High-Performance Aqueous Zn Metal Anodes, Energy Fuels, 2024, 38(16), 15789–15796 CrossRef CAS.
  43. F. Shen, S. Wu and P. Zhao, et al., Bipolar membrane Electrolyzer for CO2 electro-reduction to CO in organic electrolyte with NaClO produced as byproduct, Electrochim. Acta, 2024, 483, 144056 CrossRef CAS.
  44. R. Meirbekova, G. M. Haarberg and J. Thonstad, et al., Influence of Phosphorus on Current Efficiency in Aluminum Electrolysis at Different Current Densities, J. Electrochem. Soc., 2017, 164(7), E161–E165 CrossRef CAS.
  45. G. S. W. Hsu and S. Y. Hsu, Effects of electrode gap and electric current on chlorine generation of electrolyzed deep ocean water, J. Food Drug Anal., 2018, 26(2), 512–517 CrossRef CAS PubMed.
  46. Y. J. Chen, Y. H. Li and C. Y. Chen, Studying the Effect of Electrode Material and Magnetic Field on Hydrogen Production Efficiency, Magnetochemistry, 2022, 8(5), 53 CrossRef CAS.
  47. J. He, Z. Wang and W. Zhou, et al., Electrolytic Characteristics of Microhole Array Manufacturing Using Polyacrylamide Electrolyte in 304 Stainless Steel, Micromachines, 2023, 14(10), 1808 CrossRef PubMed.
  48. R. Ming, Y. Zhu and L. Deng, et al., Effect of electrode material and electrolysis process on the preparation of electrolyzed oxidizing water, New J. Chem., 2018, 42(14), 12143–12151 RSC.
  49. M. J. Lavorante and J. I. Franco, Performance of stainless steel 316L electrodes with modified surface to be use in alkaline water electrolyzers, Int. J. Hydrogen Energy, 2016, 41(23), 9731–9737 CrossRef CAS.
  50. A. F. Adamson, B. G. Lever and W. F. Stones, The production of hypochlorite by direct electrolysis of sea water: Electrode materials and design of cells for the process, J. Appl. Chem., 1963, 13(11), 483–495 CrossRef CAS.
  51. S. Lakshmanan and T. Murugesan, The chlor-alkali process: Work in Progress, Clean Technol. Environ. Policy, 2014, 16(2), 225–234 CrossRef CAS.
  52. R. Belhadj Ammar, T. Ounissi and L. Baklouti, et al., A New Method Based on a Zero Gap Electrolysis Cell for Producing Bleach: Concept Validation, Membranes, 2022, 12(6), 602 CrossRef CAS PubMed.
  53. F. Tanos, A. Razzouk and G. Lesage, et al., A Comprehensive Review on Modification of Titanium Dioxide-Based Catalysts in Advanced Oxidation Processes for Water Treatment, ChemSusChem, 2024, 17(6), e202301139 CrossRef CAS PubMed.
  54. A. S. O. Gomes, M. Busch and M. Wildlock, et al., Understanding Selectivity in the Chlorate Process: A Step towards Efficient Hydrogen Production, ChemistrySelect, 2018, 3(23), 6683–6690 CrossRef CAS.
  55. C. Hu, J. Liu and M. Zhang, et al., A novel CF/Ti/β-PbO2 composite anode for zinc electrowinning: preparation, electrochemical properties and application, J. Mater. Chem. A, 2023, 11(3), 1403–1418 RSC.
  56. K. Lyu, W. Zhang and Y. Chen, et al., A Novel Titanium-Based Dimensionally Stable Anode toward Energy-Efficient and Clean Electrowinning of Metallic Manganese, ACS Sustainable Chem. Eng., 2024, 12(28), 10593–10603 CrossRef CAS.
  57. K. M. Macounová, R. K. Pittkowski and R. Nebel, et al., Selectivity of Ru-rich Ru–Ti–O oxide surfaces in parallel oxygen and chlorine evolution reactions, Electrochim. Acta, 2022, 427, 140878 CrossRef.
  58. K. S. Exner, J. Anton and T. Jacob, et al., Controlling Selectivity in the Chlorine Evolution Reaction over RuO2-Based Catalysts, Angew. Chem., Int. Ed., 2014, 53(41), 11032–11035 CrossRef CAS PubMed.
  59. T. Zhang, S. Liu and F. Wang, et al., Modulation strategies of electrocatalysts for 5-hydroxymethylfurfural oxidation-assisted water splitting, Microstructures, 2024, 4(3), 2024043 CAS.
  60. A. Badreldin, E. Youssef and A. El-Ghenymy, et al., Harnessing Iron Nitride Matrices through Incorporated Cu-Based Moieties for Chlorine Evolution Reaction, ACS Appl. Energy Mater., 2024, 7(20), 9156–9167 CrossRef CAS.
  61. K. Wu, T. Peng and J. Cao, Rod-Like Fe, Bi Co-doped Co3 O4 as an Efficient Electrocatalyst for Chlorine Evolution Reaction and Ammonia-Nitrogen-Oxidation in Water, ChemistrySelect, 2025, 10(9), e202405544 CrossRef CAS.
  62. J. Hossen and N. Nakatani, Theoretical study on the carbon nanomaterial-supported Pt complex electrocatalysts for efficient and selective chlorine evolution reaction, J. Comput. Chem., 2024, jcc.27466 CrossRef PubMed.
  63. J. Fan, L. Yang and W. Zhu, Low-dimensional lateral heterojunctions made of hexagonal boron nitride and carbon materials as efficient electrocatalysts for the chlorine evolution reaction: a study of DFT and machine learning, J. Mater. Chem. A, 2024, 12(7), 4258–4267 RSC.
  64. S. Trasatti, Electrocatalysis: understanding the success of DSA, Electrochim. Acta, 2000, 45(15–16), 2377–2385 CrossRef CAS.
  65. H. B. Beer, The Invention and Industrial Development of Metal Anodes, J. Electrochem. Soc., 1980, 127(8), 303C–307C CrossRef CAS.
  66. R. Kötz and S. Stucki, Stabilization of RuO2 by IrO2 for anodic oxygen evolution in acid media, Electrochim. Acta, 1986, 31(10), 1311–1316 CrossRef.
  67. Y. Takasu, W. Sugimoto and Y. Nishiki, et al., Structural analyses of RuO2–TiO2/Ti and IrO2–RuO2–TiO2/Ti anodes used in industrial chlor-alkali membrane processes, J. Appl. Electrochem., 2010, 40(10), 1789–1795 CrossRef CAS.
  68. X. Liu, M. Chen and J. Ma, et al., Advances in the synthesis strategies of carbon-based single-atom catalysts and their electrochemical applications, China Powder Sci. Technol., 2024, 30(05), 35–46 Search PubMed.
  69. F. Qiu, L. Wang and Y. Fan, et al., Review on Structural Adjustment Strategies of Titanium-Based Metal Oxide Dimensionally Stable Anodes in Electrochemical Advanced Oxidation Technology, Adv. Eng. Mater., 2024, 26(8), 2400122 CrossRef CAS.
  70. S. Hong, T. K. Lee and M. R. Hoffmann, et al., Enhanced chlorine evolution from dimensionally stable anode by heterojunction with Ti and Bi based mixed metal oxide layers prepared from nanoparticle slurry, J. Catal., 2020, 389, 1–8 CrossRef CAS PubMed.
  71. V. Krstić and B. Pešovski, Reviews the research on some dimensionally stable anodes (DSA) based on titanium, Hydrometallurgy, 2019, 185, 71–75 CrossRef.
  72. H. Dong, W. Yu and M. R. Hoffmann, Mixed Metal Oxide Electrodes and the Chlorine Evolution Reaction, J. Phys. Chem. C, 2021, 125(38), 20745–20761 CrossRef CAS.
  73. R. Chen, V. Trieu and A. R. Zeradjanin, et al., Microstructural impact of anodic coatings on the electrochemical chlorine evolution reaction, Phys. Chem. Chem. Phys., 2012, 14(20), 7392 RSC.
  74. Z. Deng, S. Xu and C. Liu, et al., Stability of dimensionally stable anode for chlorine evolution reaction, Nano Res., 2023, 949–959 Search PubMed.
  75. A. Cornell, Ruthenium based DSA in chlorate electrolysis—critical anode potential and reaction kinetics, Electrochim. Acta, 2003, 48(5), 473–481 CrossRef CAS.
  76. J. Li, L. Chang and B. Chen, et al., Study on the Performance and Corrosion Failure Process of Porous Titanium-Based Coated Electrodes, Lubricants, 2022, 10(11), 282 CrossRef CAS.
  77. A. Fujishima, T. N. Rao and D. A. Tryk, TiO2 photocatalysts and diamond electrodes, Electrochim. Acta, 2000, 45(28), 4683–4690 CrossRef CAS.
  78. S. Saha, K. Kishor and R. G. S. Pala, Increasing electrochemical chlorine selectivity over oxygen selectivity through the optimal weakening of oxygen bonds in transition metal-doped RuO2, Catal. Sci. Technol., 2024, 14(16), 4566–4574 RSC.
  79. W. Liu, M. Zhou and J. Zhang, et al., Construction of a CoP/MnP/Cu3 P heterojunction for efficient methanol oxidation-assisted seawater splitting, Mater. Chem. Front., 2025, 9(6), 953–964 RSC.
  80. J. Choi, S. Im and J. Choi, et al., Recent advances in 2D structured materials with defect-exploiting design strategies for electrocatalysis of nitrate to ammonia, Energy Mater., 2024, 4(2), 400020 CrossRef CAS.
  81. J. Xu, J. Song and H. Li, et al., Rational surface Design and electron Regulation of co-deposition Ru and Ti on TiO2 nanotubes as self-supporting electrode for high-performance chlorine evolution reaction, J. Colloid Interface Sci., 2025, 680, 632–639 CrossRef CAS PubMed.
  82. J. Ji, J. Liu and L. Shi, et al., Ruthenium Oxide Clusters Immobilized in Cationic Vacancies of 2D Titanium Oxide for Chlorine Evolution Reaction, Small Struct., 2024, 5(3), 2300240 CrossRef CAS.
  83. W. I. Choi, J. Jo and S. Choi, et al., Thin Titanium Nitride Layer-Inserted Porous Ru0.3Sn0.35Ti0.35O2x Anode for Stable Chlorine Evolution Reaction, ACS Appl. Energy Mater., 2024, 7(21), 9962–9973 CrossRef CAS.
  84. H. Cao, G. Zhao and G. Hou, et al., A Study on the Catalytic Activity and Service Lifetime of RuO2-TiO2 Composite Electrode with TNTs as Interlayer, ChemistrySelect, 2019, 4(36), 10965–10971 CrossRef CAS.
  85. K. Xiong, L. Peng and Y. Wang, et al., In situ growth of RuO2–TiO2 catalyst with flower-like morphologies on the Ti substrate as a binder-free integrated anode for chlorine evolution, J. Appl. Electrochem., 2016, 46(8), 841–849 CrossRef CAS.
  86. H. Gong, B. Zhu and D. Zhang, et al., Ru dopant induced high selectivity and stability of ternary RuSnTi electrode toward chlorine evolution reaction, Appl. Catal., B, 2024, 349, 123892 CrossRef CAS.
  87. M. Hu, T. Yu and K. Tan, et al., Ultralow Ru loading RuO2–TiO2 with strong oxide-support interaction for efficient chlorine evolution and ammonia-nitrogen-elimination, Chem. Eng. J., 2023, 465, 143001 CrossRef CAS.
  88. X. Shao, A. Maibam and F. Cao, et al., Coordination Environment and Distance Optimization of Dual Single Atoms on Fluorine-Doped Carbon Nanotubes for Chlorine Evolution Reaction, Angew. Chem., Int. Ed., 2024, e202406273 CAS.
  89. M. Ha, P. Thangavel and N. K. Dang, et al., High-Performing Atomic Electrocatalyst for Chlorine Evolution Reaction, Small, 2023, 19(20), 2300240 CrossRef CAS PubMed.
  90. R. J. Watts, M. S. Wyeth and D. D. Finn, et al., Optimization of Ti/SnO2–Sb2O5 anode preparation for electrochemical oxidation of organic contaminants in water and wastewater, J. Appl. Electrochem., 2007, 38(1), 31–37 CrossRef.
  91. S. Chen, Y. Zheng and S. Wang, et al., Ti/RuO2–Sb2O5–SnO2 electrodes for chlorine evolution from seawater, Chem. Eng. J., 2011, 172(1), 47–51 CrossRef CAS.
  92. Y. Huang, K. D. Seo and K. A. Jannath, et al., Heteroatoms doped carbon decorated with tiny amount Pt nanoparticles as a bifunctional catalyst for hydrogen and chlorine generation from seawater, Carbon, 2022, 196, 621–632 CrossRef CAS.
  93. H. Qu, B. Li and Y. Ma, et al., Defect-Enriched Hollow Porous Carbon Nanocages Enable Highly Efficient Chlorine Evolution Reaction, Adv. Mater., 2023, 35(28), 2301359 CrossRef CAS PubMed.
  94. Y. Cheng, P. Meng and L. Li, et al., Boosting selective chlorine evolution reaction: Impact of Ag doping in RuO2 electrocatalysts, J. Colloid Interface Sci., 2025, 685, 97–106 CrossRef CAS PubMed.
  95. Z. Mahidashti and M. Rezaei, Design of CoSbyOx/Ti electrocatalyst via facile thermal decomposition method for durable and high temperature chlorine evolution reaction in acidic media, Electrochim. Acta, 2023, 472, 143426 CrossRef CAS.
  96. A. R. Jadhav, X. Liu and P. Silambarasan, et al., Stable and efficient chlorine evolution reaction with atomically dispersed Ru on surface tensile strained TiO2, Appl. Catal., B, 2024, 359, 124456 CrossRef CAS.
  97. Y. Wang, Y. Liu and D. Wiley, et al., Recent advances in electrocatalytic chloride oxidation for chlorine gas production, J. Mater. Chem. A, 2021, 9(35), 18974–18993 RSC.
  98. A. R. Zeradjanin, N. Menzel and W. Schuhmann, et al., On the faradaic selectivity and the role of surface inhomogeneity during the chlorine evolution reaction on ternary Ti–Ru–Ir mixed metal oxide electrocatalysts, Phys. Chem. Chem. Phys., 2014, 16(27), 13741–13747 RSC.
  99. K. R. Rasmi, S. C. Vanithakumari and R. P. George, et al., Active Nano Metal Oxide Coating for Bio-fouling Resistance, Trans. Indian Inst. Met., 2018, 71(6), 1323–1329 CrossRef CAS.
  100. Z. Ahaliabadeh, V. Miikkulainen and M. Mäntymäki, et al., Understanding the Stabilizing Effects of Nanoscale Metal Oxide and Li–Metal Oxide Coatings on Lithium-Ion Battery Positive Electrode Materials, ACS Appl. Mater. Interfaces, 2021, 13(36), 42773–42790 CrossRef CAS PubMed.
  101. A. S. Jamadar, R. Sutar and S. Patil, et al., Progress in metal oxide-based electrocatalysts for sustainable water splitting, Mater. Rep.: Energy, 2024, 4(3), 100283 CAS.
  102. K. Wang, P. Liu and M. Wang, et al., Modulating d-d orbitals coupling in PtPdCu medium-entropy alloy aerogels to boost pH-general methanol electrooxidation performance, Chin. Chem. Lett., 2025, 36(4), 110532 CrossRef CAS.
  103. N. Seriani and F. Mittendorfer, Platinum-group and noble metals under oxidizing conditions, J. Phys.: Condens. Matter, 2008, 20(18), 184023 CrossRef.
  104. F. Zhang and J. Chen, Structural design and application of fiber-based electrocatalytic material, China Powder Sci. Technol., 2024, 30(4), 161–170 Search PubMed.
  105. S. Iftekhar, G. Heidari and N. Amanat, et al., Porous materials for the recovery of rare earth elements, platinum group metals, and other valuable metals: a review, Environ. Chem. Lett., 2022, 20(6), 3697–3746 CrossRef CAS.
  106. X. Li and W. Hu, Review of design strategies for particle catalysts used in electrocatalytic CO2 reduction in acidic media, China Powder Sci. Technol., 2025, 31(2), 81–90 Search PubMed.
  107. J. Yang, T. Liang and B. Pan, et al., A spherical adsorbent produced from a bagasse biochar chitosan assembly for selective adsorption of platinum-group metals from wastewater, Int. J. Biol. Macromol., 2024, 266, 131142 CrossRef CAS PubMed.
  108. J. Krajczewski, R. Ambroziak and A. Kudelski, Formation and selected catalytic properties of ruthenium, rhodium, osmium and iridium nanoparticles, RSC Adv., 2022, 12(4), 2123–2144 RSC.
  109. H. Liu, Z. Zhang and M. Li, et al., Iridium Doped Pyrochlore Ruthenates for Efficient and Durable Electrocatalytic Oxygen Evolution in Acidic Media, Small, 2022, 18(30), 2202513 CrossRef CAS PubMed.
  110. A. Visintin, C. A. Tori and G. Garaventta, et al., The effect of palladium coating on hydrogen storage alloy electrodes for nickel/metal hydride batteries, J. Braz. Chem. Soc., 1997, 8(2), 125–129 CrossRef CAS.
  111. T. Miyazawa, M. Kurihara and S. Ohno, et al., Oxygen-free palladium/titanium coating, a novel nonevaporable getter coating with an activation temperature of 133 °C, J. Vac. Sci. Technol., A, 2018, 36(5), 051601 CrossRef.
  112. H. Mazhari Abbasi, K. Jafarzadeh and S. M. Mirali, An investigation of the effect of RuO2 on the deactivation and corrosion mechanism of a Ti/IrO2 + Ta2O5 coating in an OER application, J. Electroanal. Chem., 2016, 777, 67–74 CrossRef CAS.
  113. I. V. Pchelnikov, R. V. Fedotov and A. Y. Cherkesov, To selection of low-wear anode coatings based on iridium, ruthenium, titanium and tantalum oxides for direct electrolysis of natural waters, IOP Conf. Ser.:Mater. Sci. Eng., 2019, 687(6), 066050 CAS.
  114. W. Wu and Z. Chen, Micropore formation mechanism in iridium coating after high-temperature treatment, Surf. Interface Anal., 2016, 48(6), 353–359 CrossRef CAS.
  115. M. Wang, X. Ma and L. Xue, et al., Influence of iridium-tantalum interlayers on the properties of ruthenium-iridium-tin metal oxide anodes, J. Electroanal. Chem., 2024, 968, 118498 CrossRef CAS.
  116. J. Gaudet, A. C. Tavares and S. Trasatti, et al., Physicochemical Characterization of Mixed RuO2–SnO2 Solid Solutions, Chem. Mater., 2005, 17(6), 1570–1579 CrossRef CAS.
  117. C. Iwakura and K. Sakamoto, Effect of Active Layer Composition on the Service Life of (SnO2 and RuO2) – Coated Ti Electrodes in Sulfuric Acid Solution, J. Electrochem. Soc., 1985, 132(10), 2420–2423 CrossRef CAS.
  118. Q. M. Al-Bataineh, L. A. Alakhras and A. A. Ahmad, et al., Tin-doped titanium dioxide film-enhanced electrocatalytic hydrogen evolution, Mater. Sci. Eng., B, 2024, 310, 117753 CrossRef CAS.
  119. L. Zhang, L. Xu and J. He, et al., Preparation of Ti/SnO2–Sb electrodes modified by carbon nanotube for anodic oxidation of dye wastewater and combination with nanofiltration, Electrochim. Acta, 2014, 117, 192–201 CrossRef CAS.
  120. M. I. Alvarado-Ávila, E. Toledo-Carrillo and J. Dutta, Cerium Oxide on a Fluorinated Carbon-Based Electrode as a Promising Catalyst for Hypochlorite Production, ACS Omega, 2022, 7(42), 37465–37475 CrossRef PubMed.
  121. J. Zhang, M. Zhang and H. Wang, et al., Fabrication of SnO2/TiO2 composite by a chemical co-precipitation method for efficient electrocatalytic oxidation of methylene blue, Braz. J. Chem. Eng., 2024, 42(1), 171–181 CrossRef.
  122. S. Chen, L. Zhou and T. Yang, et al., Thermal decomposition based fabrication of dimensionally stable Ti/SnO2–RuO2 anode for highly efficient electrocatalytic degradation of alizarin cyanin green, Chemosphere, 2020, 261, 128201 CrossRef CAS PubMed.
  123. M. Zhang, W. Jin and F. Yang, et al., Engineering a Nanocomposite Interlayer for a Novel Ceramic-Based Forward Osmosis Membrane with Enhanced Performance, Environ. Sci. Technol., 2020, 54(12), 7715–7724 CrossRef CAS PubMed.
  124. Q. Bi, W. Guan and Y. Gao, et al., Study of the mechanisms underlying the effects of composite intermediate layers on the performance of Ti/SnO2–Sb–La electrodes, Electrochim. Acta, 2019, 306, 667–679 CrossRef CAS.
  125. Y. Hao, D. Xiong and W. Liu, et al., Controllably Designed “Vice-Electrode” Interlayers Harvesting High Performance Lithium Sulfur Batteries, ACS Appl. Mater. Interfaces, 2017, 9(46), 40273–40280 CrossRef CAS PubMed.
  126. J. Choi, H. Lim and S. Surendran, et al., Dichalcogenides as Emerging Electrocatalysts for Efficient Ammonia Synthesis: A Focus on Mechanisms and Theoretical Potentials, Adv. Funct. Mater., 2025, 2422585 CrossRef.
  127. Z. Lang, G. L. Song and X. Liao, et al., Conversion of magnetron-sputtered sacrificial intermediate layer into a stable FeCo–LDH catalyst for oxygen evolution reaction, Nano Res., 2024, 17(5), 4307–4313 CrossRef CAS.
  128. P. Shih, C. Zhang and H. Raheja, et al., Polymer Entrapment Flash Pyrolysis for the Preparation of Nanoscale Iridium-Free Oxygen Evolution Electrocatalysts, ChemNanoMat, 2020, 6(6), 930–936 CrossRef CAS.
  129. M. Mierzwa, E. Lamouroux and A. Walcarius, et al., Porous and Transparent Metal-oxide Electrodes: Preparation Methods and Electroanalytical Application Prospects, Electroanalysis, 2018, 30(7), 1241–1258 CrossRef CAS.
  130. L. Liu, Y. Chen and J. Chen, et al., Self-Supporting Electrode Fabricated by Flowing Synthesis for Efficient Hydrogen Evolution Reaction, ACS Sustainable Chem. Eng., 2023, 11(14), 5506–5514 CrossRef CAS.
  131. B. Narayanan, K. Lakshmanan and S. Gurusamy, Fabrication of highly efficient FeNi-based electrodes using thermal plasma spray for electrocatalytic oxygen evolution reaction, Surf. Interfaces, 2024, 46, 104091 CrossRef CAS.
  132. H. Qi, L. Wang and W. Sun, et al., Preparation of Ti/IrRuSnSbOx Electrodes by a Hydrosol Process for Efficient and Stable Chlorine Evolution Reaction, Ind. Eng. Chem. Res., 2023, 62(29), 11517–11526 CrossRef CAS.
  133. Y. Niu, Z. Chi and M. Li, Advancements in biomass gasification research utilizing iron-based oxygen carriers in chemical looping: A review, Mater. Rep.: Energy, 2024, 4(3), 100282 CAS.
  134. B. Liu, S. Wang and C. Wang, et al., Surface morphology and electrochemical properties of RuO2-doped Ti/IrO2-ZrO2 anodes for oxygen evolution reaction, J. Alloys Compd., 2019, 778, 593–602 CrossRef CAS.
  135. N. Royaei, T. Shahrabi and Y. Yaghoubinezhad, Optimization the selectivity property of graphene oxide modified dimensionally stable anode (DSA) produced by the sol–gel method, J. Sol-Gel Sci. Technol., 2019, 90(3), 547–564 CrossRef CAS.
  136. R. A. Raimundo, T. R. Silva and J. R. D. Santos, et al., Nickel oxide nanocatalyst obtained by a combined sol-gel and hydrothermal method for oxygen evolution reaction, MRS Commun., 2023, 13(2), 276–282 CrossRef CAS.
  137. S. Tabassum, E. Yamasue and H. Okumura, et al., Electrical stability of Al-doped ZnO transparent electrode prepared by sol-gel method, Appl. Surf. Sci., 2016, 377, 355–360 CrossRef CAS.
  138. F. A. Rodríguez, E. P. Rivero and I. González, Adapted Pechini method to prepare DSA type electrodes of RuO2-ZrO2 doped with Sb2O5 over titanium plates, MethodsX, 2018, 5, 1613–1617 CrossRef PubMed.
  139. S. Yang, M. Xiao and M. Xiao, et al., The enhanced properties of ZnCoNi-S electrode by magnetron sputtering of Au nanoparticles for high performance supercapacitors, J. Power Sources, 2024, 623, 235424 CrossRef CAS.
  140. H. Larhlimi, A. Ghailane and M. Makha, et al., Magnetron sputtered titanium carbide-based coatings: A review of science and technology, Vacuum, 2022, 197, 110853 CrossRef CAS.
  141. B. Yan, A. Chen and C. Shao, et al., Microrod structure and properties of Sb-doped Ti/SnO2 anodes prepared by magnetron sputtering, Sci. Bull., 2015, 60(24), 2135–2139 CrossRef CAS.
  142. Y. S. Ocak, D. Batibay and S. Baturay, Optical and electrical properties of Ni-doped CdO thin films by ultrasonic spray pyrolysis, J. Mater. Sci.: Mater. Electron., 2018, 29(20), 17425–17431 CrossRef CAS.
  143. Z. Ye, J. Yang and B. Li, et al., Amorphous Molybdenum Sulfide/Carbon Nanotubes Hybrid Nanospheres Prepared by Ultrasonic Spray Pyrolysis for Electrocatalytic Hydrogen Evolution, Small, 2017, 13(21), 1700111 CrossRef PubMed.
  144. M. Hamdani, M. Pereira and J. Douch, et al., Physicochemical and electrocatalytic properties of Li-Co3O4 anodes prepared by chemical spray pyrolysis for application in alkaline water electrolysis, Electrochim. Acta, 2004, 49(9–10), 1555–1563 CrossRef CAS.
  145. A. Chen, X. Zhu and J. Xi, et al., Effects of nickel doping on the preferred orientation and oxidation potential of Ti/Sb SnO2 anodes prepared by spray pyrolysis, J. Alloys Compd., 2016, 684, 137–142 CrossRef CAS.
  146. Y. Duan, Y. Chen and Q. Wen, et al., Fabrication of dense spherical and rhombic Ti/Sb–SnO2 electrodes with enhanced electrochemical activity by colloidal electrodeposition, J. Electroanal. Chem., 2016, 768, 81–88 CrossRef CAS.
  147. P. Yao, X. Chen and H. Wu, et al., Active Ti/SnO2 anodes for pollutants oxidation prepared using chemical vapor deposition, Surf. Coat. Technol., 2008, 202(16), 3850–3855 CrossRef CAS.
  148. A. A. Al-Hamzah, E. J. Smith and C. M. Fellows, Inhibition of Homogeneous Formation of Magnesium Hydroxide by Low-Molar-Mass Poly(acrylic acid) with Different End-Groups, Ind. Eng. Chem. Res., 2015, 54(7), 2201–2207 CrossRef CAS.
  149. D. Hasson, R. Semiat and D. Bramson, et al., Suppression of CaCO3 scale deposition by anti-scalants, Desalination, 1998, 118(1–3), 285–296 CrossRef CAS.
  150. M. Maril, J. L. Delplancke and C. Salvo, et al., Effect of Mg and Ca precipitates in the stability of Fe-Mn-Ni anodes used in alkaline seawater electrolysis, J. Electroanal. Chem., 2024, 967, 118422 CrossRef CAS.
  151. J. Han, Exploring the Interface of Porous Cathode/Bipolar Membrane for Mitigation of Inorganic Precipitates in Direct Seawater Electrolysis, ChemSusChem, 2022, 15(11), e202200372 CrossRef CAS PubMed.
  152. E. Jwa, H. Kim and J. Y. Nam, et al., Design strategy of cathodic electrocatalysts to effectively suppress inorganic fouling for long-term stability of reverse electrodialysis, Chem. Eng. J., 2023, 476, 146521 CrossRef CAS.
  153. F. Maya, J. M. Estela and V. Cerdà, Spectrophotometric determination of chloride in waters using a multisyringe flow injection system, Talanta, 2008, 74(5), 1534–1538 CrossRef CAS PubMed.
  154. C. Q. Ni and W. Q. Xie, Quantifying sodium hypochlorite content in hypochlorite-based disinfectants via phase-conversion headspace technique, J. Chromatogr. A, 2024, 1721, 464812 CrossRef CAS PubMed.
  155. Ł. Dąbrowski, Multidetector systems in gas chromatography, TrAC, Trends Anal. Chem., 2018, 102, 185–193 CrossRef.
  156. R. Arai, T. Sakai and Y. Nishida, et al., In situ temperature-compensated ultraviolet spectrophotometry to estimate nitrate and chloride concentrations in estuarine seawater with different salinity and composition, J. Environ. Manage., 2023, 344, 118689 CrossRef CAS PubMed.
  157. M. H. Abu Bakar, N. H. Azeman and N. N. Mobarak, et al., Uv-vis spectroscopy for simple chloride content analysis in edible oil using charged amino-functionalized carbon quantum dots fluorophore reagent, Spectrochim. Acta, Part A, 2024, 317, 124419 CrossRef CAS PubMed.
  158. M. Chen, J. Ma and Y. Feng, et al., Advanced characterization enables a new era of efficient carbon dots electrocatalytic reduction, Coord. Chem. Rev., 2025, 535, 216612 CrossRef CAS.
  159. T. B. Sil, D. Malyshev and M. Aspholm, et al., Boosting hypochlorite's disinfection power through pH modulation, BMC Microbiol., 2025, 25(1), 101 CrossRef CAS PubMed.
  160. K. Li, Q. Su and S. Li, et al., Aging of PVDF and PES ultrafiltration membranes by sodium hypochlorite: Effect of solution pH, J. Environ. Sci., 2021, 104, 444–455 CrossRef CAS PubMed.
  161. M. Yang, T. Okazaki and M. Zhang, Removal of Carbon Nanotubes from Aqueous Solutions by Sodium Hypochlorite: Effects of Treatment Conditions, Toxics, 2021, 9(9), 223 CrossRef CAS PubMed.
  162. Z. Zhang, D. Ren and D. Zhang, et al., Electrolyte dependence for the electrochemical decarboxylation of medium-chain fatty acids (n-octanoic acid) into fuel on Pt electrode, Mater. Rep.: Energy, 2024, 4(2), 100244 CAS.
  163. P. P. Wright, B. Kahler and L. J. Walsh, The effect of temperature on the stability of sodium hypochlorite in a continuous chelation mixture containing the chelator clodronate, Aust. Endod. J., 2020, 46(2), 244–248 CrossRef PubMed.
  164. W. Liu, M. Zhou and J. Zhang, et al., Construction of a CoP/MnP/Cu3P heterojunction for efficient methanol oxidation-assisted seawater splitting, Mater. Chem. Front., 2025, 9, 953–964 RSC.
  165. L. Hou, X. Peng and S. Lyu, et al., Advancements in MXene-based nanohybrids for electrochemical water splitting, Chin. Chem. Lett., 2025, 36, 110392 CrossRef CAS.

This journal is © the Partner Organisations 2025
Click here to see how this site uses Cookies. View our privacy policy here.