Next Article in Journal
Towards GreenPLM—Key Sustainable Indicators Selection and Assessment Method Development
Previous Article in Journal
Active-Arm Bridge with Two-Channel DC Voltage Source for Comparison of High-Resistance Standards in 1:1 Ratio
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Application of Capacitive Deionization in Water Treatment and Energy Recovery: A Review

1
School of Resources and Environmental Engineering, Wuhan University of Technology, Wuhan 430070, China
2
Hubei Key Laboratory of Mineral Resources Processing and Environment, Wuhan 430070, China
3
State Environmental Protection Key Laboratory of Mineral Metallurgical Resources Utilization and Pollution Control, Wuhan University of Science and Technology, Wuhan 430081, China
4
Hubei Collaborative Innovation Center for High Efficient Utilization of Vanadium Resources, Wuhan University of Science and Technology, Wuhan 430081, China
*
Author to whom correspondence should be addressed.
Energies 2023, 16(3), 1136; https://doi.org/10.3390/en16031136
Submission received: 28 November 2022 / Revised: 3 January 2023 / Accepted: 13 January 2023 / Published: 19 January 2023
(This article belongs to the Topic Capacitive Deionization Technology for Water Treatment)
(This article belongs to the Section B: Energy and Environment)

Abstract

:
Water resources are the basis for human survival and development. However, human beings face severe challenges of water pollution and freshwater shortage. With the critical advantages of low energy consumption, high efficiency, low cost, green and pollution-free, and renewable electrodes, capacitive deionization (CDI) has become an up-and-coming water treatment technology. After decades of development, the application of CDI has expanded from seawater desalination to many fields. However, the existing literature still needs a comprehensive overview of the multi-functional application of CDI technology in water treatment. Therefore, our work critically reviewed the latest research progress of CDI in water treatment to meet the technical requirements of various application fields. This paper first summarizes the various applications of CDI in water treatment, focusing on CDI’s representative research results in heavy metal removal, organic contaminants removal, water softening, phosphate and nitrate removal, and water disinfection. In addition, we also discussed the latest research progress of energy recovery and energy consumption assessment for the CDI process. Finally, this paper discusses the challenges and future opportunities facing CDI technology.

1. Introduction

The rapid development of industry and agriculture in the world has caused severe water pollution. At the same time, the increase in population numbers has led to a shortage of water resources. According to statistics, 80% of the wastewater produced in the world is discharged into the environment without purification treatment. These water sources usually contain high concentrations of TDS, heavy metals, and organic and inorganic contaminants [1,2]. Moreover, the earth’s water resources consist of 97% sea and 3% fresh water [3,4]. It is worth noting that seawater can only be used after being treated. The growing human demand for freshwater resources has exacerbated the global shortage. Water pollution and freshwater shortage are seriously threatening human survival. Therefore, the purification of water resources has become an urgent problem for human beings. At present, water treatment technology has been paid more and more attention to. Researchers have developed various water treatment technologies for seawater desalination and sewage purification, such as reverse osmosis, electrodialysis, distillation, adsorption, chemical precipitation, ion exchange, and other technologies [5,6,7,8,9]. However, these methods have some limitations in production and application, such as high energy consumption, high equipment consumption rate, easy-to-produce secondary pollution, and other problems. CDI is a new electro-sorption technology. CDI has the advantages of environmental friendliness, low cost, low energy consumption, high efficiency, easy electrode regeneration, and no secondary pollution. More importantly, based on the reversible electrochemical process, the energy consumed by capacitor deionization can be partially recovered, and the energy consumption and economic cost of the system can be reduced.
Figure 1 shows the number of research papers on the application of CDI in water treatment from 2010 to 2022, as searched using the Web of Science. Since 2010, the experimental papers on CDI technology in water purification research have been overgrown, which strongly shows the rapid development of CDI technology. Therefore, in the face of the energy crisis and global warming, CDI technology will likely provide humans with the critical method of using fresh water.
Figure 1. The number of research papers published on the application of CDI technology in water treatment from 2010 to 2022.
Figure 1. The number of research papers published on the application of CDI technology in water treatment from 2010 to 2022.
Energies 16 01136 g001
The principle of CDI technology is mainly based on the theory of electric double-layer adsorption, including adsorption and desorption processes [10]. In the adsorption process, an external voltage (<1.8 V) is applied to the two parallel plate electrodes, and an electrostatic field is formed between the two electrodes. When the ions in the solution flow through the electrode, they migrate to the electrode under the action of electric field force and concentration gradient and adsorb on the electrode material surface to form a double layer to achieve the purpose of ion removal [11]. When the adsorbed ion is nearly saturated, the voltage between the electrodes or the reverse voltage is removed. The ion is released into the solution to achieve electrode regeneration [12,13,14]. The CDI architecture diagram and removal mechanism are shown in Figure 2a. In recent years, people have begun to explore efficient electrode materials and innovate the battery structure to improve the removal rate and energy efficiency. With the rapid development of CDI technology, especially the continuous emergence of new CDI devices, representative advanced CDI systems such as membrane capacitor deionization (MCDI) technology, flow electrode capacitance deionization (FCDI) technology, and hybrid capacitor deionization technology (HCDI) have emerged [15]. The specific configurations are shown in Figure 2.
The conventional CDI system has a “common ion effect” [16]. When a voltage is applied to the electrode, ions with an opposite charge will be adsorbed on the electrode. At the same time, ions with the same charge as the electrode will enter the solution to be measured through the electrode’s pores, lowering the ion removal efficiency [13,17]. In order to improve the charge efficiency and ion removal ability of carbon materials by reducing the influence of the common ion effect, Lee et al. [18] developed a new desalination technique for MCDI by placing ion exchange membranes on one or more electrodes. Compared with the conventional CDI, MCDI avoids the common ion effect, improves charge efficiency and energy efficiency, and exhibits excellent desalination performance [19,20]. MCDI has become one of the most interesting new configurations in CDI technology.
However, MCDI technology still did not meet business needs, so FCDI technology came into being. For a long time, the optimization research of CDI focused on the fixed electrode and improved the electrode performance through the innovation of electrode materials. However, the adsorption capacity of fixed electrode materials in CDI technology is limited, and the system cannot operate continuously for a long time because of the desorption process during operation [21]. Therefore, the innovation of electrode materials has yet to effectively solve the limitation of electrode materials, which seriously restricts the application of CDI technology in treating high-concentration brine. Based on MCDI, Jeon et al. [22] developed Flow electrode capacitive deionization (FCDI) in 2013 by using porous material suspension as a mobile electrode. The appearance of FCDI technology breaks the limit of the fixed electrode, and the mobile electrode replaces the fixed electrode to adsorb ions in water. During the operation of the FCDI, the electrode suspension is continuously circulated and regenerated by desorption outside the electrode chamber. FCDI improves desalination performance and provides a system for long-term continuous operation [23,24,25].
Another new system has emerged with the development and application of MCDI, FCDI, and other structures. Lee’s team reported another new seawater desalination technology called Hybrid capacitive deionization (HCDI) [20,26]. HCDI system combines CDI and asymmetric capacitance system of battery material and is made of two different electrode materials, a carbon electrode and the battery material [27,28]. During desalination, sodium ions are retained in the redox electrode through a chemical reaction. Chlorine ions are trapped in the double-layer electric layer formed by the surface of the porous carbon electrode. This design increases the efficiency of desalination.
CDI technology has unique advantages in water treatment. With the development and improvement of electrode materials and the continuous optimization of different types of CDI structures, the functional application areas of CDI technology have been widely expanded to include brackish water desalination, seawater desalination, heavy metal removal, water disinfection, organic contaminants removal, energy recovery, and other areas. Other review papers summarized CDI electrode materials and CDI battery structure. However, the existing literature still needs a comprehensive overview of the multi-functional application of CDI technology in water treatment. Therefore, this paper comprehensively summarizes the latest research advances in CDI technology in water treatment. It focuses on representative applications in various fields, providing a specific theoretical basis and the latest research results for relevant researchers. At the same time, it further promotes the application and development of CDI technology in various fields. It summarizes and examines the future development prospects and challenges of CDI technology in scientific research and practical applications.

2. Application Progress of CDI Technology in Water Treatment

2.1. Water Desalination

In 1960, Blair and Murphy first proposed the concept of desalination, called “electrochemical desalination of water” [29]. They conducted pioneering research on the application of CDI technology in seawater desalination. Porous-activated carbon is the earliest electrode material used in CDI technology, but it has some defects, such as poor conductivity, wettability, and low capacity. Therefore, researchers have improved the adsorption performance of the electrode through modification, activation, doping, and other methods. Ryoo et al. [30,31] doped the activated carbon electrode with titanium dioxide. The existence of titanium dioxide significantly increases the electro-adsorption of electrode materials on NaCl, which proves that the desalination performance of activated carbon materials can be improved by doping some substances. After a period of research on conventional activated carbon, researchers were not satisfied with its adsorption capacity. Therefore, other carbon materials are constantly being developed to improve the desalination performance of electrode materials. Manganese dioxide (MnO2) is a TMO material with high theoretical specific capacitance (>1300 F/g), but the lower electrical conductivity limits the practical application of MnO2 [32,33,34,35,36]. MnO2 is commonly compounded with carbon to enhance the electro-adsorption performance and is used as an HCDI electrode. Biomass carbonization usually produces biochar under high temperatures and in an oxygen-free atmosphere, so it is considered a new substitute for traditional activated carbon. Adorna et al. [37] prepared coconut shell-derived activated biochar (AB) with a high specific surface area using waste coconut shells and combined AB with MnO2 to prepare AB-MnO2 nanocomposites. The material exhibited excellent CDI performance at 1.2 V, with a specific electro-sorption capacity of 68.4 mg/g for Na+ removal. Other researchers have synthesized MnO2 nanowires with different pore crystalline structures using hydrothermal methods and investigated the effect of crystal structure on desalination performance [38]. As shown in Figure 3a, the crystal structure has a significant influence on the desalination performance of HCDI. It can be adjusted according to the ion hydration radius to optimize desalination performance further.
Graphene has excellent conductivity and a hierarchical porous structure compared with other carbon materials. Its specific surface area reaches 2630 m2/g, showing that graphene can have excellent desalination performance when used as a CDI electrode [39,40,41,42,43]. Zhang et al. [44] fabricated graphene electrodes with excellent desalination performance by assembling core-shell composites with incomplete graphene oxide shells into bulk membranes using the compression molding method and heat treating them under an NH3 atmosphere (Figure 3b).
When graphene sheets are rolled into nanoscale cylindrical structures, single walls (SWCNTs) or multi-walled carbon nanotubes (CNTs) with high mechanical strength, good chemical stability, high conductivity, and good capacitance retention will be formed [45,46,47]. Thus, CNT has considerable potential for electro-sorption, indicating that it is very suitable as a CDI electrode material. Meanwhile, CNT is considered an ideal candidate for CDI electrode materials. In the research by Hu et al., they prepared N-doped carbon nanotube materials containing cobalt and cobalt oxides (Co-CO3O4/N-CNT) by the pyrolysis method [48]. They uniformly encapsulated Co-CO3O4 nanoparticles in situ in the internal channels of conductive CNTs (Figure 3c). The electrode material has an excellent desalination ability of 66.91 mg/g NaCl. The interaction between cobalt and cobalt oxide is key to improving desalination capacity.
However, the electro-adsorption performance of a single carbon can no longer meet the application requirements of the rapidly developing CDI technology. In order to obtain electrode materials with better electro-sorption performance, researchers combined MOFs derived carbon, Prussian blue (PB), and blue analogues (PBAs) with other carbon materials, such as MOFs-derived carbon-graphene composites, PBA-CNTs, MOFs-derived carbon-carbon nanofiber composites, and MOFs-derived carbon–carbon nanotube composites [49,50,51,52]. Shi et al. [53] synthesized Fe-MOF and graphene at 800 °C and prepared a new type of MCDI electrode. The electrode material exhibits an excellent electro-adsorption capacity of 37.6 mg/g and high desalination stability in 1000 mg/L NaCl solution and 1.2 V voltage. The excellent performance of this material is important for two reasons. On the one hand, graphene provides a highly conductive pathway, and on the other hand, Fe-MOF provides a highly porous carbon structure.
Prussian blue (PB) and blue analogues (PBAs) belong to intercalation materials. Compared with other carbon materials, this material has the advantages of high theoretical specific capacity, safety, low toxicity, and easy synthesis [54,55]. However, this material has some disadvantages, such as poor electronic conductivity, particle aggregation, and poor cycling stability [56,57]. Therefore, the defects of PBA limit its practical application in CDI. Combining carbon materials with PBA has proven to be an effective way to improve the desalination performance of PBA [58,59]. Gong et al. [60] successfully prepared a stepwise hollow Prussian blue/CNTs composite (SHPB/CNTs) by combining PBA with CNTs (Figure 3d). When the material is used as a CDI electrode, it shows an ultra-high desalination capacity of 103.4 mg/g and excellent cycle stability.
Figure 3. (a) Manganese oxide nanowires with different tunnel crystal structures as HCDI electrodes for ion removal [38]. Copyright 2018, Elsevier. (b) Schematic diagram of the construction process of graphene-based freestanding membrane electrodes and seawater desalination [44]. Copyright 2022, Elsevier. (c) Co-Co3O4 encapsulated in nitrogen-doped carbon nanotubes for capacitive desalination [48]. Copyright 2022, Elsevier. (d) Schematic illustration of the synthesis process for SHPB [60]. Copyright 2022, Elsevier.
Figure 3. (a) Manganese oxide nanowires with different tunnel crystal structures as HCDI electrodes for ion removal [38]. Copyright 2018, Elsevier. (b) Schematic diagram of the construction process of graphene-based freestanding membrane electrodes and seawater desalination [44]. Copyright 2022, Elsevier. (c) Co-Co3O4 encapsulated in nitrogen-doped carbon nanotubes for capacitive desalination [48]. Copyright 2022, Elsevier. (d) Schematic illustration of the synthesis process for SHPB [60]. Copyright 2022, Elsevier.
Energies 16 01136 g003
In addition, in order to find high-performance materials that can remove salt ions through other mechanisms, Younes et al. [61] prepared cobalt ferric oxide (CFO) metal oxide nanoparticles (NPs) through the hydrothermal method and then prepared high-performance nanocomposites by vacuum filtration or the freeze-drying method. The electrode material exhibits an excellent electro-adsorption capacity of 64.5 mg/g and high desalination stability in 50 mg/L NaCl solution and 1.2 V voltage. The excellent performance of this material benefits from the porous and interconnected 3D structure of the 3DrGO, and it provides a larger surface area to form EDL capacitance. In addition, the added porous 3DrGO entangled with the spinel crystals (CoFe2O4) in the composite allowed for a quick ion diffusion across the interconnected open macroporous structures.

2.2. Water Softening

Hard water refers to water containing more soluble calcium-magnesium compounds. Although hard water will not directly endanger human health, it can bring many troubles to life and industry, such as reduced washing efficiency of soap and detergent and scaling of water appliances [11,62,63]. More seriously, the existence of hard water will increase the production cost of enterprises and reduce their profits. Water hardness (expressed in units of mg/L) is the total amount of divalent cations such as Mg2+ or Ca2+ in the solution [64]. The more calcium and magnesium salts dissolved in the water, the greater the water’s hardness. At the same time, the greater the hardness of the water, the easier it is for minerals such as calcium and magnesium to precipitate and eventually cause blockages in equipment such as water pipes and boilers. As the cost of equipment maintenance increases gradually, the hardness of water has become one of the criteria for judging production costs. Therefore, the softening of hard water has important practical significance. Various methods have been developed to address the problem of water hardenings, such as chemical precipitation, ion exchange, adsorption, electrodialysis, electrolysis, electrochemistry, and nanofiltration [65,66,67,68,69,70,71,72]. CDI is a chemical-free and cost-effective technology that has been increasingly applied to water softening. Table 1 summarizes some research results of hard water treatment with CDI in recent years.
Preparing advanced carbon materials such as carbon aerogels, graphene, and CNTs requires expensive precursors and complex procedures. Low costs and renewability have led to biomass materials attracting the attention of researchers. Deng et al. [76] synthesized laminated activated carbon with an ultra-high specific surface area (1943.2 m2/g) and large oxygen-containing functional groups by KOH charring and activating chestnut inner shells at high temperatures (Figure 4a). A mixed solution containing NaCl and CaCl2 is injected into the CDI device. The carbonized and activated chestnut shell was used as a CDI electrode. Under the condition of the 30 mL/min volume flow rate, 1.4 V voltage was applied to the CDI device for 20 min. It was found that the adsorption capacities of CaCl2 and NaCl were 21.0 mg/g and 17.7 mg/g, respectively.
The low selectivity and electro-adsorption properties of carbon-based electrode materials severely limit the application of CDI technology in water softening [13,80]. Pseudocapacitive electrode materials selectively interact with specific ions through Faradaic redox reactions or ion (de)intercalation, providing high selectivity for the electro-sorption of Ca2+ in water [81,82]. Xu et al. [73] used the co-precipitation method to prepare copper hexacyanoferrate (CuHCF) as a pseudo-capacitive electrode. CuHCF is used as a cathode and activated carbon as an anode. The HCDI cells without ion-exchange membranes were used to study the Ca2+ adsorption performance of CuHCF, as shown in Figure 4b. A high electrical adsorption capacity of 42.8 mg/g was obtained, which exceeded the reported carbon-based electrode. At the same time, the adsorption selectivity of CuHCF to Ca2+ is 3.05, higher than Na+ and Mg2+.
Yoon et al. [75] successfully prepared a Ca-alginate coated electrode and used it for hard water softening. The results showed that the removal rate of Ca2+ by Ca-alginate coated-CDI was 44% higher than that by conventional CDI. Further studies revealed that the calcium alginate coating showed comparable performance to conventional MCDI when used as an MCDI electrode.
Pseudocapacitor-type HCDI, which is carried out through the intrinsic pseudocapacitive effect, is now gradually being used in deionization. Xu et al. [77] prepared a manganese spinel ferrite (MFO) selective adsorption electrode by one-step solvothermal synthesis (Figure 4c). MFO electrode has a highly selective ability to adsorb hardness ions 534.6 μmol/g (CaCl2) and 980.4 μmol/g (MgCl2), mainly because of its unique crystal structure and pseudocapacitive property. It was also found that the electro-affinity of Mg2+ for Na+ was much higher than that of Ca2+ in the mixed solution, achieving a super high hardness selectivity factor of 34.76. The electrochemical response of unary and multiple electrolytes revealed a unique pseudocapacitive affinity based on the cation (de)intercalation-redox mechanism, which is the main reason for the selective removal of hardness ions by electrode materials. This work provides an in-depth study of the intrinsic mechanism of redox reactions and efficient ion selectivity, which provides a new reference for developing new Faraday materials with high selectivity for water softening.
The coexistence of Tetracycline (TC) and hardness ions in natural water are common [83]. More importantly, TC and hardness ions quickly form TC-metal compound pollutants in the environment [84]. Therefore, the simultaneous removal of TC and hardness ions has important practical significance for environmental protection. Considering this, Sun et al. [85] assembled a symmetrical CDI device using kelp-derived hierarchical porous carbon (KHPC) to remove TC and hardness ions from water simultaneously (Figure 4d). The excellent deionization performance of KHPC electrodes is mainly caused by the improved potential-induced electrical adsorption performance of KHPC, high specific surface area, and suitable hierarchical pore structure. Although CDI water treatment technology has been applied to water softening, more research still needs to be done in this area. Furthermore, most of the research is still in the laboratory stage. The scaling formed during the operation of the CDI device may deteriorate the water quality, which still needs further study. Therefore, it is necessary to develop electrode materials with low cost, no pollution, and high selectivity and to conduct in-depth research on the sustainable operation of CDI in practical applications.
Figure 4. (a) Fabrication of CSx from chestnut inner shell and Structure diagram of CS500 [76]. Copyright 2022, Elsevier. (b) Application of CuHCF electrode material in hard water treatment [73]. Copyright 2020, American Chemical Society. (c) (1) Schematic illustration of the preparation of MFO nanospheres; (2) Schematic illustration of selective ion electro-sorption of the MFO electrode in the PHCDI system [77]. Copyright 2021, American Chemical Society. (d) Simultaneous removal of TC and hardness ions from water using the KHPC electrodes with a symmetrical CDI device [85]. Copyright 2021, Elsevier.
Figure 4. (a) Fabrication of CSx from chestnut inner shell and Structure diagram of CS500 [76]. Copyright 2022, Elsevier. (b) Application of CuHCF electrode material in hard water treatment [73]. Copyright 2020, American Chemical Society. (c) (1) Schematic illustration of the preparation of MFO nanospheres; (2) Schematic illustration of selective ion electro-sorption of the MFO electrode in the PHCDI system [77]. Copyright 2021, American Chemical Society. (d) Simultaneous removal of TC and hardness ions from water using the KHPC electrodes with a symmetrical CDI device [85]. Copyright 2021, Elsevier.
Energies 16 01136 g004

2.3. Removal of Heavy Metals

With rapid economic development, the industry has been expanding in recent years. Toxic and harmful heavy metals are released into the environment with industrial wastewater. Heavy metal pollution has seriously threatened the natural environment and human health. In the past decades, various techniques have been proposed to remove heavy metals from water, including chemical precipitation, ion exchange, electrochemistry, adsorption, membrane separation, coagulation, and other methods [86,87,88,89]. However, these techniques have disadvantages, such as high energy consumption and a tendency to cause secondary pollution [90,91,92,93,94,95]. CDI technology does not use chemical reagents and does not produce waste during water treatment. The advantages of CDI technology has attracted many researchers to explore its potential application in removing heavy metals in wastewater. Table 2 summarizes some research results of CDI application in heavy metal removal in recent years.
Electrostatic interaction is the primary mechanism of CDI technology to remove heavy metals, but non-electrostatic adsorption (physical or chemical adsorption) may also play a role [96]. Tang et al. [97] used the activated carbon cloth electrode as the CDI electrode to adsorb Zn2+ under the condition of applying 1.2 V voltage to the CDI device (Figure 5a). It was found that electrostatic interactions caused 87% of the total adsorbed Zn2+, and the remaining 13% was caused by non-electrostatic adsorption. The limited double-layer (EDL) capacitance and desalination capacity of traditional carbon-based electrodes seriously affects their practical applications. Li et al. [98] prepared a self-supporting porous carbon nanofiber membrane (ZnO@N-PCNM) with high nitrogen doping and nano-Zn content by electrospinning technology using ZIF-8 as a template (Figure 5b). The prepared ZnO@N-PCNM-10 was used as a CDI electrode. At an applied voltage of 1.8 V, the material successfully removed Pb2+ (99.42%), Cu2+ (68.46%), and Cd2+ (70.36%) ions. The results indicate that the material effectively removes heavy metal ions. High N doping and Zn loading make the material obtain higher specific capacitance and better electro-sorption performance.
Generally, solutions contain many kinds of ions, so the process of removing target ions will inevitably be interfered with by other ions. Therefore, the selective removal of target ions by CDI systems remains a significant challenge in practical applications where multi-ion hybrid solutions are treated. Zhang et al. [99] used the FCDI device to remove and separate Cu2+ from a salty solution with the assistance of (electro)deposition. During migration, Cu2+ is deposited on the carbon particles, whereas most Na+ is retained in the solution. It is noteworthy that Na+ desorption is strong; therefore, the separation of the two ions is more pronounced in the short-circuited closed cycle (SCC) mode. At an initial pH of 8.0, only Cu2+ was present on the cathode. In contrast, at an initial pH of 2.5, there was more Cu than Cu2+, indicating that Cu2+ ions were effectively electrodeposited onto the cathode and formed Cu. The presence of Cu2+ in the carbon particles could be caused by weak electro- or Physico-chemical adsorption. It was found that the difference in ion distribution in the flow electrode led to the slower decay of Cu2+ removal performance than that of Na+. This shows that FCDI technology has advantages in removing Cu2+.
Many pseudocapacitive electrode materials based on Faraday redox reactions are used to efficiently and selectively remove various heavy metal ions. Xu et al. [100] prepared poly-pyrrole-coated dual-metal perovskite-type oxide as a pseudocapacitive cathode (Figure 5c). The maximum Cd removal capacity of 144.6 mg/g was achieved by asymmetry pseudocapacitive deionization. The CDI process maintained a retention rate of 93.4% after 150 h of operation. Cd is easily affected by pH value. With the decrease in pH value, the adsorption capacity of Cd decreased. The main reason for this phenomenon is that Cd competes with H+ for active sites. When the voltage value is changed, the removal capacity increases with the increase of voltage, but the energy consumption also increases. Compared with single metal oxide electrode materials, bimetallic oxides have better electrochemical performance and cycle stability and are easier to be popularized in industry. Therefore, bimetallic oxide is a potential CDI electrode material.
From the perspective of environmental chemistry, V5+ is the most toxic vanadium ion in various valence states, so V5+ is the main target for removing vanadium ions. Bao et al. [101] combined ion exchange resin with activated carbon to prepare a composite electrode, which was used as the electrode of the CDI battery to adsorb V5+. Finally, a 106.89 mg/g high electric adsorption capacity was obtained. In another work, the team prepared a composite material that combines porous carbon derived from ZIF-8 with activated carbon [102]. The electrode is used as the electrode of CDI. A high adsorption value of 77.27 mg/g was obtained at a voltage of 1.2 V.
Generally, the ion radius, ion charge, and an ion’s physicochemical affinity for the electrode will affect the selectivity of the electrode material to the ion [103,104]. It has been shown that ions with a small hydration radius, higher valence ions, and homovalent ions are more readily adsorbed by electrode materials [105,106,107]. For example, Huang et al. [105] carried out batch electro-sorption experiments using an activated carbon cloth for Cr3+, Cd2+, and Pb2+ (Figure 5d). It was found that the order of removal efficiency of these three cations was Cr3+ > Pb2+ > Cd2+. The order of the hydration radius of these three cations was Cr3+ > Pb2+ > Cd2+. Cr3+ has the highest hydration radius and valence state compared to the other two ions. Therefore, the removal efficiency of Cr3+ is better than the other two ions. Generally, CDI technology is very effective in removing ions when the ion concentration is low.
Coincidentally, CDI technology matches most heavy metal-polluted water sources and can be combined with other water treatment technologies to enhance the removal effect. Therefore, CDI technology has broad prospects in heavy metal removal.
Figure 5. (a) The process of electro-assisted adsorption of Zn2+ from aqueous solution using activated carbon cloth as electrode in batch-flow mode [97]. Copyright 2019, American Chemical Society. (b) Controlled synthesis of ZnO-modified N-doped porous carbon nanofiber membrane for removal of heavy metal ions by CDI [98]. Copyright 2022, Elsevier. (c) Schematic illustration for preparation process of FeMnO3 and FeMnO3@PPy [100]. Copyright 2022, Elsevier. (d) The schematic of CDI assembly for metal removal [105]. Copyright 2015, Elsevier.
Figure 5. (a) The process of electro-assisted adsorption of Zn2+ from aqueous solution using activated carbon cloth as electrode in batch-flow mode [97]. Copyright 2019, American Chemical Society. (b) Controlled synthesis of ZnO-modified N-doped porous carbon nanofiber membrane for removal of heavy metal ions by CDI [98]. Copyright 2022, Elsevier. (c) Schematic illustration for preparation process of FeMnO3 and FeMnO3@PPy [100]. Copyright 2022, Elsevier. (d) The schematic of CDI assembly for metal removal [105]. Copyright 2015, Elsevier.
Energies 16 01136 g005
Table 2. Removal efficiency of heavy metal ions by CDI.
Table 2. Removal efficiency of heavy metal ions by CDI.
PollutantConcentration (mg/L)ElectrodesFlow Rate (mL/min)Applied Voltage (V)Removal EfficiencyRef.
Cu2+96 Activated carbon10-94.00%[99]
Pb2+50self-supporting porous carbon nanofiber membrane101.899.42%[98]
Cu2+68.46%
Cd2+70.36%
Zn2+40Nitrogen-doped porous carbon201.033.10 mg/g[108]
As0.02–0.1Activated carbon30001–1.5>80%[109]
Cd2+20Perovskite oxide201.2144.60 mg/g[100]
Pb2+100Wood-converted Carbon51.219.52 mg/g[107]
Cr2+50 20.00 mg/g
V5+1500Composite materials based on ion-exchange resin and activated carbon30-106.89 mg/g[101]
V5+750Composite materials based on porous carbon derived from ZIF-8 and activated carbon30 1.277.27 mg/g[102]
U6+100highly porous phosphate-functionalized graphene501.2 545.70 mg/g[110]
Ni2+50Multi-walled CNTs501.5 145.73 mg/g[111]

2.4. Removal of Phosphate and Nitrate

Products such as fertilizers and pesticides used by humans are rich in N and P. Excessive N and P are released into natural water bodies through industrial effluents, agricultural runoff, and municipal wastewater systems [112] (Figure 6). Excessive enrichment of phosphate and nitrate in water will lead to eutrophication [113,114,115,116,117,118]. Algal blooms caused by water eutrophication seriously endanger the water supply system and ecosystem [119,120]. In addition, phosphate is an indispensable resource for industrial and agricultural production, but phosphate rock is a nonrenewable resource, so the world is facing a shortage of phosphorus resources [121,122]. Therefore, removing and recovering phosphate and nitrate from water is urgent. In recent years, some researchers have applied CDI technology to remove phosphate and nitrate in water, and some research progress has been made. Table 3 summarizes some research results of CDI application in phosphate and nitrate removal in recent years.
For example, Zhang et al. [123] developed a terephthalic acid intercalated carbon nanotube composite material (ZnZr-COOH/CNT) as an anode for CDI to evaluate its phosphate removal capacity (Figure 7a). For the wastewater containing 6 mg/L phosphate, the maximum adsorption capacity of 13.65 mg/g was obtained by applying 1.2 V voltage to the CDI device. The network structure of CNTs provides more space for phosphate adsorption and enhances selective phosphate adsorption. At the same time, there are two main reasons why the electrode material has a high affinity for phosphate. First, Zn and Zr in the electrode are strongly complexed with phosphate. Secondly, a hydrogen bond was formed between the hydroxyl group of phosphate and the carboxyl group of terephthalic acid.
Compared with CDI, FCDI uses flow electrodes, significantly improving desalination performance. The flow electrode provides a long-term continuous operation system for FCDI. Zhang et al. [124] used FCDI to remove and recover phosphorus from low-concentration wastewater in the SCC mode (Figure 7b). In the six cycles, the phosphorus removal rate and adsorption capacity remained stable at about 97% and 0.97 mg/g. This shows that the FCDI system has good cycle removal performance for phosphorus. Interestingly, when the pH value is 2.15–7.20, the phosphate is mainly H2PO4, and when the pH value is 5, the removal effect is the best. At pH 7.20–12.35, HPO42− is the primary phosphate. The diffusion coefficient of H2PO4 is not only larger than HPO42−, but also the hydration radius of H2PO4 is smaller than HPO42−. This shows that H2PO4 can be more easily transferred from the spacing channel to the flow electrode. The above results show that the removal efficiency of phosphorus by FCDI is highly influenced by the initial pH value of feed water, and it is easier to remove phosphorus in a weak acid environment. They also found that the FCDI system could enrich phosphorus in solution 16 times after 11 h of operation. This shows that the mobile electrode can concentrate and enrich low-concentration phosphorus-containing wastewater, significantly promoting the recovery efficiency of phosphate.
Bian et al. [125] also studied the removal of phosphorus by the FCDI system. They explored phosphate ions’ existing form and migration behavior in the FCDI system (Figure 7c). When the pH value of the solution is 5, the phosphorus removal effect is the best, and the phosphorus removal rate is up to 38.3 mg/min. Compared with pH 9, the phosphorus removal efficiency at pH 5 increased by 84~104%. The main reason for the increase in phosphorus removal efficiency is that when the pH value is 5, P changes from HPO42− to H2PO4−. The primary phosphate types in the feed solution and flow electrode will change with the change of pH value, and the reaction formula is shown in Equations (1)–(3) [126]. They also studied the energy consumption of the FCDI system for phosphorus removal. They also studied the energy consumption of the FCDI system for phosphorus removal. When the pH is reduced from 9 to 5, the energy consumption decreases by 42.9–56.1%. Based on the research results, they found that the removal rate and recovery rate of phosphate ions mainly depend on the pH value of the feed solution or electrolyte. The pH value of the solution is critical to the removal and recovery of phosphorus by FCDI. Therefore, proper pH values can improve phosphorus removal and recovery rates.
H3PO4 ⇌ H+ + H2PO4 pKa1 = 2.148
H2PO4 ⇌ H+ + HPO42− pKa2 = 7.198
HPO42− ⇌ H+ + PO43− pKa3 = 12.375
Vivianite is the general name of a group of phosphate minerals with similar structures. Vivianite is considered slow available fertilizer and a possible reagent for manufacturing lithium iron phosphate [127,128]. Thus, some researchers have attempted to recover the phosphate from the wastewater and prepare it as a vivianite with some economic value. For example, Zhang et al. [129] pioneered the construction of a coupling system of FCDI and fluidized bed crystallization to recover phosphorus from wastewater. First, the flow electrode concentrates the phosphate, and then the concentrated phosphate forms the vivianite in the fluidized bed crystallization reactor. The experimental results show that the FCDI system can concentrate 63% phosphate into the electrode chamber. The fluidized bed crystallization system can convert approximately 80% phosphate to vivianite. This study provides a new avenue for effectively recovering phosphorus from wastewater as a value-added product.
Phosphate and nitrate need to be removed from the water, and the recovered phosphate can form products for secondary utilization. However, nitrate needs to be degraded without pollution after removal to avoid environmental pollution again. Recently, some researchers have applied CDI or MCDI technology to remove nitrate from wastewater. Fateminia et al. [130] prepared a ternary composite electrode for CDI to remove nitrate. An electro-adsorption capacity of 6.01 mg/g was obtained with a nitrate removal efficiency of 60.01% under 2.0 V applied to the CDI device. This technology mainly uses the synergistic effect of four materials (activated carbon/PVDF/polyyline/ZrO2) to enhance the nitrate adsorption capacity in the feed solution. With the advantages of low cost and low energy consumption, MCDI is a promising alternative technology for water treatment. In another study, Cetinkaya [131] applied MCDI to remove nitrate from water. The method achieved a nitrate removal rate of 83.07% with an applied voltage of 0.8 V. The study also showed that the main factor affecting the environment is the materials used in the manufacturing process. Therefore, environmentally friendly materials should be used in the nitrate removal process to avoid contamination of the environment by the materials used.
Although CDI technology has been applied for the removal of nitrate, however, the treatment process will produce nitrate-concentrated waste liquid [132]. Therefore, the concentrated waste liquid needs to be further treated. Some researchers have studied the simultaneous removal and degradation of nitrate. This method makes CDI technology more economical and applicable in nitrate removal. Hu et al. [133] doped Pd nanoparticles into Pd/NiAl-layered metal oxide by in-situ hydrothermal method and successfully prepared Pd-containing film electrodes. They used the electrode in the electro-adsorption/reduction solution containing nitrate (Figure 7d). The study successfully converted NO3 in solution to non-polluting N2, reducing secondary waste generation.
A similar study was done by Rogers et al. [134]. They used catalytic capacitive deionization (CCDI) technology for the removal and degradation of nitrate from wastewater and used commonly used porous carbon as the cathode, as shown in Figure 7e. The technology has the common advantages of catalytic water treatment and CDI technology. At 1.5 V, 91% NO3 can be reduced to N2. They also analyzed the overall energy efficiency of the treatment process. CCDI may be considered an economical and applicable wastewater treatment method further to promote the development of the next-generation CDI system.
Figure 7. (a) Mechanism diagram of phosphate adsorption by ZnZr-COOH/CNT electrode [120]. Copyright 2022, Elsevier. (b) Schematic diagram of the FCDI cell operated in the SCC mode for phosphorus removal and recovery [124]. Copyright 2020, Elsevier. (c) The existing form and migration behavior of phosphate ions in FCDI at different pH values [125]. Copyright 2020, American Chemical Society. (d) The process of electro-sorption/reduction of NO3 by CDI with Pd/NiAl LMO electrode [133]. Copyright 2018, Elsevier. (e) The process of electro-sorption/reduction of NO3 by CDI with In-Pd NP electrode [134]. Copyright 2021, American Chemical Society.
Figure 7. (a) Mechanism diagram of phosphate adsorption by ZnZr-COOH/CNT electrode [120]. Copyright 2022, Elsevier. (b) Schematic diagram of the FCDI cell operated in the SCC mode for phosphorus removal and recovery [124]. Copyright 2020, Elsevier. (c) The existing form and migration behavior of phosphate ions in FCDI at different pH values [125]. Copyright 2020, American Chemical Society. (d) The process of electro-sorption/reduction of NO3 by CDI with Pd/NiAl LMO electrode [133]. Copyright 2018, Elsevier. (e) The process of electro-sorption/reduction of NO3 by CDI with In-Pd NP electrode [134]. Copyright 2021, American Chemical Society.
Energies 16 01136 g007
Table 3. Removal efficiency of phosphate and nitrate by CDI.
Table 3. Removal efficiency of phosphate and nitrate by CDI.
PollutantConcentrationCDI typeApplied Voltage (V)Removal EfficiencypHEnergy ConsumptionRef.
P6 mg/LCDI1.213.65 mg/g70.0075 kWh/g-P[123]
P2.5 mMCDI1.235 mg/g6.5-[135]
P62 mg/LFCDI-164 mg/L per cycle8.327.8 kWh/kg-P[136]
P8 mg/LFCDI-61.9%-21.8 kWh/kg-P[137]
P100 mg/LFCDI1.297%5-[124]
P500 mg/LFCDI1.238.3 mg/min50.59 kWh/kg-P and 0.043 kWh/m3-water[125]
N0.5 mMCDI1.22.4 × 10−3 mmol/m25.6248.8 kJ/mol-NO3−[138]
N50 mg/LCCDI1.591%-13.2 kWh·m−3 order−1[134]
N42 mg/LCCDI1.0-6.5-[133]
N200 mg/LMCDI1.853.3%5.32-[139]
N70 mg/LCDI2.060.01%6-[130]
N250 mg/LMCDI0.883.07%--[131]

2.5. Removal of Organic Contaminants

The coexistence of organic contaminants in wastewater will cause electrode pollution and deterioration, significantly inhibiting CDI’s practical application. The organic contaminants in the solution will form scaling on the electrode and block the pores of the electrode material, seriously reducing the CDI operation efficiency and the electro-adsorption performance of the electrode material [140]. In order to ensure the continuous operation of the CDI system in practical application, effective methods are needed to remove organic contaminants from wastewater.
Some researchers combined CDI technology with other technologies to treat wastewater containing multiple pollutants. For example, Chen et al. [141] combined CDI and electro-oxidation (CDI-EO) to develop a hybrid system for the simultaneous removal of heavy metal ions and organics (Figure 8). At the same time, they have developed two electrodes that selectively remove heavy metals and organic contaminants. The anode of the CDI-EO cell is carbon-coated graphite paper for the removal of heavy metals, and the cathode is a RuO2-IrO2-protected titanium plate for the degradation of organic contaminants. CDI-EO battery is used to treat wastewater containing Cu2+ and Acid Orange 7. At the current density of 5 mA/cm2, 80% Cu2+ and 89% Acid Orange 7 were removed, respectively. Interestingly, the degradation process of Acid Orange 7 and the removal process of Cu2+ will not affect and interfere with each other. The new technology should test the electro-adsorption performance and investigate its energy consumption because the technology with high energy consumption has no practical significance. The study achieved low energy consumption of 4.66 kWh/m3-wastewater. This proves that the CDI-EO system not only removes metal ions and organic contaminants but also has the advantage of low energy consumption, making it economically suitable for wastewater treatment.
In another study, Liang et al. [142] combined ultrafiltration with CDI. They invented a novel integrated ultrafiltration-capacitive-deionization (UCDI) process that can simultaneously remove organic contaminants and inorganic salts from wastewater. Firstly, the system uses an ultrafiltration component or electro-catalytic oxidation to remove organic contaminants. Secondly, the system uses CDI electrodes to adsorb and remove inorganic salts. The UCDI process thus achieves the simultaneous removal of organic contaminants and inorganic salts from wastewater by combining ultrafiltration and CDI. It was found that if organic contaminants, Ca2+ and Mg2+, are present in the solution simultaneously, the organic contaminants may have complexation and promote the removal of Ca2+ and Mg2+.
The electrochemical advanced oxidation process (EAOP) is a cleaning technology for removing organic contaminants. Some researchers have combined EAOP with CDI to simultaneously remove organic contaminants and inorganic salts from wastewater [143]. The electrode of the CDI battery is an activated carbon plate. To remove phenol and sodium chloride, a 2.0 V voltage battery with oxygen is applied. Under the optimal conditions, the removal rates of phenol, TOC, and salinity are 90%, 60%, and 20%, respectively. Organic contaminants in the solution can be removed by degradation through direct or indirect oxidation [144]. As expected, the phenol removal pathways in this study included both direct and indirect oxidation.
At present, there are still few studies based on CDI technology for the treatment of organic contaminants. The above research shows that CDI technology needs to combine with other technologies to form a new system to have the ability to remove organic contaminants. Even though the integrated approach based on CDI solves the problem of synergistic removal of organic contaminants and other ions in wastewater, the facilities and operation are complex in the later stage, and the maintenance cost is high.

2.6. Water Disinfection

The disinfection process is usually the final stage of water treatment. Water disinfection is an important guarantee to ensure that the water used by human beings is free from harmful bacteria and viruses. The main methods of water disinfection are chemical and physical [145]. The chemical method uses disinfectants for disinfection, such as chlorine, hydrochloric acid, ammonia, alum, ozone, and other disinfectants [146,147]. Although these disinfectants remove harmful pathogens, they can produce toxic by-products (DBP) during disinfection [146,148,149,150]. Physical methods mainly include UV radiation disinfection, but this method requires high clarity of the water, poor disinfection, and high cost and energy consumption [146]. Therefore, finding an effective, low-cost, and clean disinfection technology is essential.
In recent years, the disinfection performance of CDI systems has become increasingly prominent. Many studies have shown that CDI is an alternative disinfection technology with great potential. Usually, bacteria survive in an environment with a specific pH range, which gives them a negative electrostatic charge [151]. Therefore, bacteria are considered negative ions in an aqueous solution. Then, the bacteria move towards the anode electrode under the action of an external electric field. During the disinfection process, bacteria will be adsorbed to the anode surface by the electric field force. The bacteria will be killed when contacting the bactericidal materials on the anode surface [152]. In conclusion, CDI can be used to remove bacteria from aqueous solutions. Wang et al. [153] prepared Alkoxysilane octadecyl dimethyl-ammonium chloride (ODDMAC) functionalized graphene oxide (GO) hybrid electrodes for capacitive deionization. The technology mainly uses ODDMAC attached to the surface of GO material to kill bacteria. Based on the above killing principle, under the condition of applying 1.2 V voltage to the CDI battery, the water with a pollutant concentration of 104 CFU/mL was sterilized. The results showed that the disinfection efficiency of the hybrid electrode to E. coli, P. aeruginosa, and S. aureus was 99.65%, 98.64%, and 97.69%, respectively. Cao et al. [154] used chitosan to modify 3-D channel structured ordered mesoporous carbons (Ia3d). The altered activated carbon was used as a CDI electrode to disinfect seawater. The results showed that only 0.01% of E. coli survived after 30 min when the brine was disinfected at 1.2 V. Chitosan has a bactericidal effect. At the same time, anionic Ia3d was modified by chitosan, and the modified Ia3d became cationic. Therefore, the electrode materials synthesized in this study have conductivity and bactericidal effects. The above research proves that CDI technology has excellent bactericidal performance. The ability of CDI technology to disinfect is mainly attributed to the bactericidal effect of the composite materials. When the composite material is used as the anode of the CDI battery, the bacteria with negative static charge pass through the CDI electrode will be absorbed and killed by the anode electrode.

3. Energy Recovery and Energy Consumption Assessment

Energy recovery is a significant advantage of CDI in the field of water treatment because energy recovery can reduce energy consumption. Energy consumption is not only an essential indicator for evaluating the performance of CDI, but it is also a key indicator of superiority over other technologies [14,155]. More importantly, the low energy consumption makes CDI technology occupy a larger market in water treatment. The essence of CDI is to realize the migration of solution ions through the action of the electric field force. The migration of ions on the electrode surface and between solutions will promote the formation and disappearance of the double electric layer. This is very similar to the principle of an electric double-layer capacitor, which shows that CDI also has the characteristics of energy storage and recovery. If the CDI battery is regarded as an energy storage element, the adsorption and desorption stages can be regarded as the charging and discharging stages of the energy storage element.
More and more people have recently built energy recovery devices for CDI technology and use the devices to recover energy from water treatment processes. Chen et al. [156] successfully constructed an energy recovery system based on a four-switch buck-boost converter, which can transfer electric energy from the CDI module to the supercapacitor (Figure 9). The study’s results showed that the system achieved an excellent energy recovery rate of 49.6%. It has been found that higher NaCl concentrations and shorter distances between electrodes are more conducive to energy recovery. This is because electronic resistances and ionic resistances produce low energy dissipation. In other studies, by changing the circuit connection sequence of the CDI system, not only can 70% of the energy be effectively recovered, the recovered energy can also be used in another ion adsorption process [157].
The two most common charging modes are constant voltage (CV) and constant current (CC) operation. Chen et al. [158] found that the energy recovery rate decreased with increasing charge current, as did the voltage. The research shows that 46.6% of the energy can be recovered by connecting external loads. In addition, the energy recovery rate also showed a significant downward trend with an increasing discharge current. The energy recovery rate was generally higher in the CC mode than in the CV mode. The main reason for this phenomenon is the severe energy loss caused by ohmic and non-ohmic resistance at higher charging currents.
Zhao et al. [159] and Choi [160] also found that the energy consumption of CC mode is lower than that of CV mode. Another study found that the resistance dissipation of CC mode was smaller than that of CV mode, so the energy consumption was lower [161], although, most studies show that CC mode is more energy-saving than CV mode. However, the CV mode is still widely studied and applied in academic and commercial fields. This is mainly because the CC mode has a certain degree of operational difficulty [155]. This is because it is difficult to control the voltage level of CDI cells in CC mode [155]. More importantly, the most energy-efficient CDI operation mode may not be the best solution. Although a model can achieve energy efficiency, it does not necessarily have a high removal efficiency. For example, although CV mode has higher energy consumption than CC mode, CC mode cannot achieve the required removal efficiency in producing ultrapure water [162].
Recent issues regarding the energy consumption of MCDI have focused on comparative studies of different modes of operation, while the development of new charging modes has been neglected. Choi et al. [163] proposed an advanced staged voltage (SV) mode for the first time. The voltage applied in this mode gradually increases with the operation of the MCDI battery (Figure 10). In flexible operation, the SV mode meets the product water quality requirements and achieves the required removal efficiency and minimum energy consumption, although, the CC mode of operation shows the highest energy efficiency compared to other modes. However, the SV mode may be a promising operating strategy to achieve the required product quality.

4. Challenges and Future Perspectives

Although CDI technology has made significant progress in water treatment, there are still some key challenges to be solved in the multi-functional application of CDI. Currently, most of the CDI technology application research is only based on laboratory research on simulated water. The composition of synthetic water is simple, while that of actual water is complex. Actual water contains a variety of non-target treatment objects, such as organic, inorganic, and impurity ions, which may hurt the application of CDI and reduce the treatment performance. Therefore, future research should be conducted on the actual water environment to promote CDI technology’s commercial application further.
The sustainability of CDI is challenging because of the scaling during operation. Scaling affects the removal of CDI and increases energy consumption. In the future, the collaborative combination of CDI and other technologies should be deeply studied to minimize the impact of scaling on the CDI operation process. Only by promoting the sustainable operation of CDI can the application of expanded CDI technology be further expanded.
The primary research of CDI technology has been extensively studied, and the results have been rich enough. However, there are still few studies on the pilot scale. Therefore, CDI technology needs more pilot-scale research in seawater desalination, hard water softening, heavy metal removal, organic contaminants removal, phosphate and nitrate removal, water disinfection, energy collection, and other fields.
Current research suggests that energy recovery is feasible for CDI/MCDI systems, but more research still needs to be done. Researchers should do more work to improve energy recovery rates further and promote energy recovery commercialization. With the aggravation of global warming and resource shortage, CDI technology should make further progress in energy conservation and resource recovery.
Electrode materials should be developed for single or multiple target pollutants to achieve efficient removal. Electrode regeneration is also noteworthy as it can reduce energy consumption and economic costs.

5. Conclusions

With the increasing population of the world and the rapid development of industry, the shortage of water resources has become increasingly prominent. Based on the advantages of energy conservation, high efficiency, low cost, green, and pollution-free CDI technology, CDI has been widely used in many fields. This paper introduces the principle of CDI technology and CDI device configuration and comprehensively summarizes the latest research progress and existing problems of CDI technology in water treatment, including seawater desalination, hard water softening, heavy metal removal, organic contaminants removal, phosphate and nitrate removal, water disinfection, and energy collection. The current challenges of the application of CDI technology in water treatment are discussed, and future research directions and prospects of CDI are proposed. Recent research shows that compared with other technologies, CDI technology has a great application prospect in the field of water treatment. This work has significant reference value for promoting the development of CDI technology. CDI is a challenging but potential water treatment technology. Even after decades of development, it is still considered a new water treatment technology.

Author Contributions

For research articles with several authors, Conceptualization, C.X. and S.B.; methodology, C.X., S.B. and B.C.; software, W.D.; validation, Y.L.; formal analysis, C.X.; investigation, C.X.; resources, S.B. and Y.Z.; data curation, C.X.; writing—original draft preparation, C.X.; writing—review and editing, C.X. and S.B.; visualization, C.X.; supervision, Y.Z.; project administration, S.B.; funding acquisition, S.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China, grant number 51874222 and 52074204.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Adam, M.; Othman, M.; Kurniawan, T.; Puteh, M.; Ismail, A.; Khongnakorn, W.; Rahman, M.; Jaafar, J. Advances in adsorptive membrane technology for water treatment and resource recovery applications: A critical review. J. Environ. Chem. Eng. 2022, 10, 107633. [Google Scholar] [CrossRef]
  2. Liu, E.; Lee, L.; Ong, S.; Ng, H. Treatment of industrial brine using capacitive deionization (CDI) towards zero liquid discharge—Challenges and optimization. Water Res. 2020, 183, 116059. [Google Scholar] [CrossRef] [PubMed]
  3. Shahmirzadi, M.; Hosseini, S.; Luo, J.; Ortiz, I. Significance, evolution and recent advances in adsorption technology, materials and processes for desalination, water softening and salt removal. J. Environ. Manag. 2018, 215, 324–344. [Google Scholar] [CrossRef] [PubMed]
  4. Ahualli, S.; Iglesias, G.; Delgado, Á. Principles and theoretical models of CDI: Experimental approaches. In Interface Science and Technology; Elsevier: Amsterdam, The Netherlands, 2018; pp. 169–192. [Google Scholar]
  5. Khosravi, R.; Moussavi, G.; Ghaneian, M.; Ehrampoush, M.; Barikbin, B.; Ebrahimi, A.; Sharifzadeh, G. Chromium adsorption from aqueous solution using novel green nanocomposite: Adsorbent characterization, isotherm, kinetic and thermodynamic investigation. J. Mol. Liq. 2018, 256, 163–174. [Google Scholar] [CrossRef]
  6. Zahoran, R.; Kukovecz, A.; Toth, A.; Horvath, D.; Schuszter, G. High-speed tracking of fast chemical precipitations. Phys. Chem. Chem. Phys. 2019, 21, 11345–11350. [Google Scholar] [CrossRef] [Green Version]
  7. Zha, L.; Li, J.; Zhou, B.; Yuan, Y.; Yuan, H. Research progress on the ion exchange treatment and resource utilization of chromium(Ⅵ) electroplating wastewater. Mater. Prot. 2021, 54, 123–126. [Google Scholar]
  8. Chen, B.; Bao, S.; Zhang, Y.; Ren, L. A novel and sustainable technique to precipitate vanadium from vanadium-rich solutions via efficient ultrasound irradiation. J. Clean. Prod. 2022, 339, 130755. [Google Scholar] [CrossRef]
  9. Jarin, M.; Dou, Z.; Gao, H.; Chen, Y.; Xie, X. Salinity exchange between seawater/brackish water and domestic wastewater through electrodialysis for potable water. Front. Environ. Sci. Eng. 2023, 17, 16. [Google Scholar] [CrossRef]
  10. Xing, W.; Liang, J.; Tang, W.; He, D.; Yan, M.; Wang, X.; Luo, Y.; Tang, N.; Huang, M. Versatile applications of capacitive deionization (CDI)-based technologies. Desalination 2020, 482, 114390. [Google Scholar] [CrossRef]
  11. Choi, J.; Dorji, P.; Shon, H.; Hong, S. Applications of capacitive deionization: Desalination, softening, selective removal, and energy efficiency. Desalination 2019, 449, 118–130. [Google Scholar] [CrossRef]
  12. Huang, W.; Zhang, Y.; Bao, S.; Song, S. Desalination by capacitive deionization with carbon-based materials as electrode: A review. Surf. Rev. Lett. 2013, 20, 1330003. [Google Scholar] [CrossRef]
  13. Suss, M.; Porada, S.; Sun, X.; Biesheuvel, P.; Yoon, J.; Presser, V. Water desalination via capacitive deionization: What is it and what can we expect from it? Energy Environ. Sci. 2015, 8, 2296–2319. [Google Scholar] [CrossRef] [Green Version]
  14. Zhao, Y.; Wang, Y.; Wang, R.; Wu, Y.; Xu, S.; Wang, J. Performance comparison and energy consumption analysis of capacitive deionization and membrane capacitive deionization processes. Desalination 2013, 324, 127–133. [Google Scholar] [CrossRef]
  15. Tang, W.; Liang, J.; He, D.; Gong, J.; Tang, L.; Liu, Z.; Wang, D.; Zeng, G. Various cell architectures of capacitive deionization: Recent advances and future trends. Water Res. 2019, 150, 225–251. [Google Scholar] [CrossRef]
  16. Suss, M.; Presser, V. Water Desalination with Energy Storage Electrode Materials. Joule 2018, 2, 10–15. [Google Scholar] [CrossRef] [Green Version]
  17. Zhang, C.; He, D.; Ma, J.; Tang, W.; Waite, T. Faradaic reactions in capacitive deionization (CDI)–Problems and possibilities: A review. Water Res. 2018, 128, 314–330. [Google Scholar] [CrossRef] [PubMed]
  18. Lee, J.; Park, K.; Eum, H.; Lee, C. Desalination of a thermal power plant wastewater by membrane capacitive deionization. Desalination 2006, 196, 125–134. [Google Scholar] [CrossRef]
  19. Chen, Z.; Xu, X.; Ding, Z.; Wang, K.; Sun, X.; Lu, T.; Konarova, M.; Eguchi, M.; Shapter, J.; Pan, L.; et al. Ti3C2 MXenes-derived NaTi2(PO4)3/MXene nanohybrid for fast and efficient hybrid capacitive deionization performance. Chem. Eng. J. 2021, 407, 127148. [Google Scholar] [CrossRef]
  20. Lee, J.; Kim, S.; Kim, C.; Yoon, J. Hybrid capacitive deionization to enhance the desalination performance of capacitive techniques. Energy Environ. Sci. 2014, 7, 3683–3689. [Google Scholar] [CrossRef]
  21. Ratajczak, P.; Suss, M.; Kaasik, F.; Beguin, F. Carbon electrodes for capacitive technologies. Energy Storage Mater. 2019, 16, 126–145. [Google Scholar] [CrossRef]
  22. Jeon, S.; Park, H.; Yeo, J.; Yang, S.; Cho, C.; Han, M.; Kim, D. Desalination via a new membrane capacitive deionization process utilizing flow-electrodes. Energy Environ. Sci. 2013, 6, 1471–1475. [Google Scholar] [CrossRef]
  23. Wang, J.; Shi, Z.; Fang, J.; Chu, B.; Li, N.; Shui, L.; Wang, G.; Chen, F. The optimized flow-electrode capacitive deionization (FCDI) performance by ZIF-8 derived nanoporous carbon polyhedron. Sep. Purif. Technol. 2022, 281, 119345. [Google Scholar] [CrossRef]
  24. Epshtein, A.; Nir, O.; Monat, L.; Gendel, Y. Treatment of acidic wastewater via fluoride ions removal by SiO2 particles followed by phosphate ions recovery using flow-electrode capacitive deionization. Chem. Eng. J. 2020, 400, 125892. [Google Scholar] [CrossRef]
  25. Cho, Y.; Lee, K.; Yang, S.; Choi, J.; Park, H.; Kim, D. A novel three-dimensional desalination system utilizing honeycomb-shaped lattice structures for flow-electrode capacitive deionization. Energy Environ. Sci. 2017, 10, 1746–1750. [Google Scholar] [CrossRef]
  26. Kim, S.; Lee, J.; Kim, C.; Yoon, J. Na2FeP2O7 as a Novel Material for Hybrid Capacitive Deionization. Electrochim. Acta 2016, 203, 265–271. [Google Scholar] [CrossRef]
  27. Divyapriya, G.; Vijayakumar, K.; Nambi, I. Development of a novel graphene/Co3O4 composite for hybrid capacitive deionization system. Desalination 2019, 451, 102–110. [Google Scholar] [CrossRef]
  28. Jacob, L.; Prasanna, K.; Vengatesan, M.; Santhoshkumar, P.; Lee, C.; Mittal, V. Binary Cu/ZnO decorated graphene nanocomposites as an efficient anode for lithium ion batteries. J. Ind. Eng. Chem. 2018, 59, 108–114. [Google Scholar] [CrossRef]
  29. Blair, J.; Murphy, G. Electrochemical Demineralization of Water with Porous Electrodes of Large Surface Area; ACS Publications: Washington, DC, USA, 1960. [Google Scholar]
  30. Ryoo, M.; Kim, J.; Seo, G. Role of titania incorporated on activated carbon cloth for capacitive deionization of NaCl solution. J. Colloid Interface Sci. 2003, 264, 414–419. [Google Scholar] [CrossRef] [PubMed]
  31. Ryoo, M.; Seo, G. Improvement in capacitive deionization function of activated carbon cloth by titania modification. Water Res. 2003, 37, 1527–1534. [Google Scholar] [CrossRef]
  32. Hu, C.; Liu, F.; Lan, H.; Liu, H.; Qu, J. Preparation of a manganese dioxide/carbon fiber electrode for electrosorptive removal of copper ions from water. J. Colloid Interface Sci. 2015, 446, 359–365. [Google Scholar] [CrossRef]
  33. Chou, T.; Huang, C.; Doong, R.; Hu, C. Architectural design of hierarchically ordered porous carbons for high-rate electrochemical capacitors. J. Mater. Chem. A 2013, 1, 2886–2895. [Google Scholar] [CrossRef]
  34. Wan, C.; Jiao, Y.; Li, J. Core-shell composite of wood-derived biochar supported MnO2 nanosheets for supercapacitor applications. RSC Adv. 2016, 6, 64811–64817. [Google Scholar] [CrossRef]
  35. Athouel, L.; Arcidiacono, P.; Ramirez-Castro, C.; Crosnier, O.; Hamel, C.; Dandeville, Y.; Guillemet, P.; Scudeller, Y.; Guay, D.; Belanger, D.; et al. Investigation of cavity microelectrode technique for electrochemical study with manganese dioxides. Electrochim. Acta 2012, 86, 268–276. [Google Scholar] [CrossRef]
  36. Younes, H.; Zou, L. Asymmetric configuration of pseudocapacitive composite and rGO electrodes for enhanced capacitive deionization. Environ. Sci.-Water Res. Technol. 2020, 6, 392–403. [Google Scholar] [CrossRef]
  37. Adorna, M., Jr.; Borines, M.; Van Dien, D.; Doong, R. Coconut shell derived activated biochar-manganese dioxide nanocomposites for high performance capacitive deionization. Desalination 2020, 492, 114602. [Google Scholar] [CrossRef]
  38. Byles, B.; Cullen, D.; More, K.; Pomerantseva, E. Tunnel structured manganese oxide nanowires as redox active electrodes for hybrid capacitive deionization. Nano Energy 2018, 44, 476–488. [Google Scholar] [CrossRef]
  39. Yin, H.; Zhao, S.; Wan, J.; Tang, H.; Chang, L.; He, L.; Zhao, H.; Gao, Y.; Tang, Z. Three-dimensional graphene/metal oxide nanoparticle hybrids for high-performance capacitive deionization of saline water. Adv. Mater. 2013, 25, 6270–6276. [Google Scholar] [CrossRef]
  40. Xu, Y.; Lin, Z.; Huang, X.; Wang, Y.; Huang, Y.; Duan, X. Functionalized graphene hydrogel-based high-performance supercapacitors. Adv. Mater. 2013, 25, 5779–5784. [Google Scholar] [CrossRef]
  41. Xu, Y.; Chen, C.; Zhao, Z.; Lin, Z.; Lee, C.; Xu, X.; Wang, C.; Huang, Y.; Shakir, M.; Duan, X. Solution processable holey graphene oxide and its derived macrostructures for high-performance supercapacitors. Nano Lett. 2015, 15, 4605–4610. [Google Scholar] [CrossRef]
  42. Ji, L.; Meduri, P.; Agubra, V.; Xiao, X.; Alcoutlabi, M. Graphene-based nanocomposites for energy storage. Adv. Energy Mater. 2016, 6, 1502159. [Google Scholar] [CrossRef] [Green Version]
  43. Younes, H.; Lou, D.; Rahman, M.; Choi, D.; Hong, H.; Zou, L. Review on 2D MXene and graphene electrodes in capacitive deionization. Environ. Technol. Innov. 2022, 28, 102858. [Google Scholar] [CrossRef]
  44. Zhang, G.; Li, W.; Chen, Z.; Long, J.; Xu, C. Freestanding N-doped graphene membrane electrode with interconnected porous architecture for efficient capacitive deionization. Carbon 2022, 187, 86–96. [Google Scholar] [CrossRef]
  45. Oladunni, J.; Zain, J.; Hai, A.; Banat, F.; Bharath, G.; Alhseinat, E. A comprehensive review on recently developed carbon based nanocomposites for capacitive deionization: From theory to practice. Sep. Purif. Technol. 2018, 207, 291–320. [Google Scholar] [CrossRef]
  46. Feng, J.; Xiong, S.; Ren, L.; Wang, Y. Atomic layer deposition of TiO2 on carbon-nanotubes membrane for capacitive deionization removal of chromium from water. Chin. J. Chem. Eng. 2022, 45, 15–21. [Google Scholar] [CrossRef]
  47. Cao, Z.; Wei, B. A perspective: Carbon nanotube macro-films for energy storage. Energy Environ. Sci. 2013, 6, 3183–3201. [Google Scholar] [CrossRef] [Green Version]
  48. Hu, X.; Min, X.; Li, X.; Si, M.; Liu, L.; Zheng, J.; Yang, W.; Zhao, F. Co-Co3O4 encapsulated in nitrogen-doped carbon nanotubes for capacitive desalination: Effects of nano-confinement and cobalt speciation. J. Colloid Interface Sci. 2022, 616, 389–400. [Google Scholar] [CrossRef]
  49. Liu, Y.; Ma, J.; Lu, T.; Pan, L. Electrospun carbon nanofibers reinforced 3D porous carbon polyhedra network derived from metal-organic frameworks for capacitive deionization. Sci. Rep. 2016, 6, 32784. [Google Scholar] [CrossRef] [Green Version]
  50. Xu, X.; Wang, M.; Liu, Y.; Lu, T.; Pan, L. Metal-organic framework-engaged formation of a hierarchical hybrid with carbon nanotube inserted porous carbon polyhedra for highly efficient capacitive deionization. J. Mater. Chem. A 2016, 4, 5467–5473. [Google Scholar] [CrossRef]
  51. Gao, T.; Zhou, F.; Ma, W.; Li, H. Metal-organic-framework derived carbon polyhedron and carbon nanotube hybrids as electrode for electrochemical supercapacitor and capacitive deionization. Electrochim. Acta 2018, 263, 85–93. [Google Scholar] [CrossRef]
  52. Weng, J.; Wang, S.; Zhang, P.; Li, C.; Wang, G. A review of metal-organic framework-derived carbon electrode materials for capacitive deionization. New Carbon Mater. 2021, 36, 117–129. [Google Scholar] [CrossRef]
  53. Shi, W.; Ye, C.; Xu, X.; Liu, X.; Ding, M.; Liu, W.; Cao, X.; Shen, J.; Yang, H.; Gao, C. High-performance membrane capacitive deionization based on metal-organic framework-derived hierarchical carbon structures. ACS Omega 2018, 3, 8506–8513. [Google Scholar] [CrossRef]
  54. Zhou, A.; Cheng, W.; Wang, W.; Zhao, Q.; Xie, J.; Zhang, W.; Gao, H.; Xue, L.; Li, J. Hexacyanoferrate-type prussian blue analogs: Principles and advances toward high-performance sodium and potassium ion batteries. Adv. Energy Mater. 2021, 11, 2000943. [Google Scholar] [CrossRef]
  55. Li, Q.; Zheng, Y.; Xiao, D.; Or, T.; Gao, R.; Li, Z.; Feng, M.; Shui, L.; Zhou, G.; Wang, X.; et al. Faradaic electrodes open a new era for capacitive deionization. Adv. Sci. 2020, 7, 2002213. [Google Scholar] [CrossRef]
  56. Qian, J.; Wu, C.; Cao, Y.; Ma, Z.; Huang, Y.; Ai, X.; Yang, H. Prussian blue cathode materials for sodium-ion batteries and other ion batteries. Adv. Energy Mater. 2018, 8, 1702619. [Google Scholar] [CrossRef]
  57. Zhang, W.; Song, H.; Cheng, Y.; Liu, C.; Wang, C.; Khan, M.; Zhang, H.; Liu, J.; Yu, C.; Wang, L.; et al. Core-shell prussian blue analogs with compositional heterogeneity and open cages for oxygen evolution reaction. Adv. Sci. 2019, 6, 1801901. [Google Scholar] [CrossRef]
  58. Nai, J.; Lou, X. Hollow structures based on prussian blue and its analogs for electrochemical energy storage and conversion. Adv. Mater. 2019, 31, e1706825. [Google Scholar] [CrossRef] [PubMed]
  59. Vafakhah, S.; Guo, L.; Sriramulu, D.; Huang, S.; Saeedikhani, M.; Yang, H. Efficient sodium-ion intercalation into the freestanding prussian blue/graphene aerogel anode in a hybrid capacitive deionization system. ACS Appl. Mater. Interfaces 2019, 11, 5989–5998. [Google Scholar] [CrossRef] [PubMed]
  60. Gong, A.; Zhao, Y.; Liang, B.; Li, K. Stepwise hollow Prussian blue/carbon nanotubes composite as a novel electrode material for high-performance desalination. J. Colloid Interface Sci. 2022, 605, 432–440. [Google Scholar] [CrossRef]
  61. Younes, H.; Rahman, M.; Hong, H.; AlNahyan, M.; Ravaux, F. Capacitive deionization performance of asymmetric nanoengineered CoFe2O4 carbon nanomaterials composite. Environ. Sci. Pollut. Res. Int. 2022. [Google Scholar] [CrossRef]
  62. Gabrielli, C.; Maurin, G.; Francy-Chausson, H.; Thery, P.; Tran, T.; Tlili, M. Electrochemical water softening: Principle and application. Desalination 2006, 201, 150–163. [Google Scholar] [CrossRef]
  63. Zhi, S.; Zhang, K. Hardness removal by a novel electrochemical method. Desalination 2016, 381, 8–14. [Google Scholar] [CrossRef]
  64. Hazeltine, B.; Bull, C. Field Guide to Appropriate Technology; Elsevier: Amsterdam, The Netherlands, 2003. [Google Scholar]
  65. Mohammadian, M.; Sahraei, R.; Ghaemy, M. Synthesis and fabrication of antibacterial hydrogel beads based on modified-gum tragacanth/poly(vinyl alcohol)/Ag0 highly efficient sorbent for hard water softening. Chemosphere 2019, 225, 259–269. [Google Scholar] [CrossRef]
  66. Ma, Z.; Ren, L.; Ying, D.; Jia, J.; Shao, J. Dual-layer Janus charged nanofiltration membranes constructed by sequential electrospray polymerization for efficient water softening. Chemosphere 2023, 310, 136929. [Google Scholar] [CrossRef] [PubMed]
  67. Kianfar, F.; Kianfar, E. Synthesis of isophthalic acid/aluminum nitrate thin film nanocomposite membrane for hard water softening. J. Inorg. Organomet. Polym. Mater. 2019, 29, 2176–2185. [Google Scholar] [CrossRef]
  68. Oliveira, E.; Alves, C.; Franca, A.; Nascimento, R.; Sasaki, J.; Loiola, A. NaA zeolite synthesized over fiberglass as a strategy for optimizing hard water softening. Quim. Nova 2022, 45, 16–22. [Google Scholar]
  69. Eleni, G.; Georgios, P.; Konstantina, K.; Alexandros, B. Ca2+ removal from water by the use of Na-palygorskite for potential water softening. Water Supply 2022, 22, 156–169. [Google Scholar]
  70. Dou, W.; Zhou, Z.; Jiang, L.; Jiang, A.; Huang, R.; Tian, X.; Zhang, W.; Chen, D. Sulfate removal from wastewater using ettringite precipitation: Magnesium ion inhibition and process optimization. J. Environ. Manag. 2017, 196, 518–526. [Google Scholar] [CrossRef]
  71. Arias-Paic, M.; Cawley, K.; Byg, S.; Rosario-Ortiz, F. Enhanced DOC removal using anion and cation ion exchange resins. Water Res. 2016, 88, 981–989. [Google Scholar] [CrossRef]
  72. Jin, H.; Yu, Y.; Chen, X. Membrane-based electrochemical precipitation for water softening. J. Membr. Sci. 2020, 597, 117639. [Google Scholar] [CrossRef]
  73. Xu, Y.; Zhou, H.; Wang, G.; Zhang, Y.; Zhang, H.; Zhao, H. Selective pseudocapacitive deionization of calcium ions in copper hexacyanoferrate. Acs Appl. Mater. Interfaces 2020, 12, 41437–41445. [Google Scholar] [CrossRef]
  74. Nie, P.; Hu, B.; Shang, X.; Xie, Z.; Huang, M.; Liu, J. Highly efficient water softening by mordenite modified cathode in asymmetric capacitive deionization. Sep. Purif. Technol. 2020, 250, 117240. [Google Scholar] [CrossRef]
  75. Yoon, H.; Lee, J.; Kim, S.; Kang, J.; Kim, S.; Kim, C.; Yoon, J. Capacitive deionization with Ca-alginate coated-carbon electrode for hardness control. Desalination 2016, 392, 46–53. [Google Scholar] [CrossRef]
  76. Deng, D.; Luhasile, M.; Li, H.; Pan, Q.; Zheng, F.; Wang, Y. A novel layered activated carbon with rapid ion transport through chemical activation of chestnut inner shell for capacitive deionization. Desalination 2022, 531, 115685. [Google Scholar] [CrossRef]
  77. Xu, Y.; Xiang, S.; Zhou, H.; Wang, G.; Zhang, H.; Zhao, H. Intrinsic pseudocapacitive affinity in manganese spinel ferrite nanospheres for high-performance selective capacitive removal of Ca2+ and Mg2+. ACS Appl. Mater. Interfaces 2021, 13, 38886–38896. [Google Scholar] [CrossRef] [PubMed]
  78. Seo, S.; Jeon, H.; Lee, J.; Kim, G.; Park, D.; Nojima, H.; Lee, J.; Moon, S. Investigation on removal of hardness ions by capacitive deionization (CDI) for water softening applications. Water Res. 2010, 44, 2267–2275. [Google Scholar] [CrossRef] [PubMed]
  79. Tuan, T.; Chung, S.; Lee, J.; Lee, J. Improvement of water softening efficiency in capacitive deionization by ultra purification process of reduced graphene oxide. Curr. Appl. Phys. 2015, 15, 1397–1401. [Google Scholar] [CrossRef]
  80. Li, Y.; Chen, N.; Li, Z.; Shao, H.; Qu, L. Frontiers of carbon materials as capacitive deionization electrodes. Dalton Trans. 2020, 49, 5006–5014. [Google Scholar] [CrossRef]
  81. Wang, S.; Wang, G.; Wu, T.; Li, C.; Wang, Y.; Pan, X.; Zhan, F.; Zhang, Y.; Wang, S.; Qiu, J. Membrane-free hybrid capacitive deionization system based on redox reaction for high-efficiency NaCl removal. Environ. Sci. Technol. 2019, 53, 6292–6301. [Google Scholar] [CrossRef]
  82. Zhang, X.; Zuo, K.; Zhang, X.; Zhang, C.; Liang, P. Selective ion separation by capacitive deionization (CDI) based technologies: A state-of-the-art review. Environ. Sci.-Water Res. Technol. 2020, 6, 243–257. [Google Scholar] [CrossRef]
  83. Chen, Y.; Li, H.; Wang, Z.; Tao, T.; Hu, C. Photoproducts of tetracycline and oxytetracycline involving self-sensitized oxidation in aqueous solutions: Effects of Ca2+ and Mg2+. J. Environ. Sci. 2011, 23, 1634–1639. [Google Scholar] [CrossRef]
  84. Werner, J.; Arnold, W.; McNeill, K. Water hardness as a photochemical parameter: Tetracycline photolysis as a function of calcium concentration, magnesium concentration, and pH. Environ. Sci. Technol. 2006, 40, 7236–7241. [Google Scholar] [CrossRef] [PubMed]
  85. Sun, N.; Zhou, H.; Zhang, H.; Zhang, Y.; Zhao, H.; Wang, G. Synchronous removal of tetracycline and water hardness ions by capacitive deionization. J. Clean. Prod. 2021, 316, 128251. [Google Scholar] [CrossRef]
  86. Chai, W.; Cheun, J.; Kumar, P.; Mubashir, M.; Majeed, Z.; Banat, F.; Ho, S.; Show, P. A review on conventional and novel materials towards heavy metal adsorption in wastewater treatment application. J. Clean. Prod. 2021, 296, 126589. [Google Scholar] [CrossRef]
  87. Kim, S.; Chu, K.; Al-Hamadani, Y.; Park, C.; Jang, M.; Kim, D.; Yu, M.; Heo, J.; Yoon, Y. Removal of contaminants of emerging concern by membranes in water and wastewater: A review. Chem. Eng. J. 2018, 335, 896–914. [Google Scholar] [CrossRef]
  88. Al-Qodah, Z.; Al-Shannag, M. Heavy metal ions removal from wastewater using electrocoagulation processes: A comprehensive review. Sep. Sci. Technol. 2017, 52, 2649–2676. [Google Scholar] [CrossRef]
  89. Li, P.; Tao, H. Cell surface engineering of microorganisms towards adsorption of heavy metals. Crit. Rev. Microbiol. 2015, 41, 140–149. [Google Scholar] [CrossRef]
  90. Ding, W.; Bao, S.; Zhang, Y.; Xiao, J. Efficient selective extraction of scandium from red mud. Miner. Process. Extr. Metall. Rev. 2022. [Google Scholar] [CrossRef]
  91. Ding, W.; Bao, S.; Zhang, Y.; Xiao, J. Mechanism and kinetics study on ultrasound assisted leaching of gallium and zinc from corundum flue dust. Miner. Eng. 2022, 183, 107624. [Google Scholar] [CrossRef]
  92. Zhang, L.; Wu, Y.; Qu, X.; Li, Z.; Ni, J. Mechanism of combination membrane and electro-winning process on treatment and remediation of Cu2+ polluted water body. J. Environ. Sci. 2009, 21, 764–769. [Google Scholar] [CrossRef]
  93. Chen, B.; Bao, S.; Zhang, Y. Synergetic strengthening mechanism of ultrasound combined with calcium fluoride towards vanadium extraction from low-grade vanadium-bearing shale. Int. J. Min. Sci. Technol. 2021, 31, 1095–1106. [Google Scholar] [CrossRef]
  94. Crini, G.; Lichtfouse, E. Advantages and disadvantages of techniques used for wastewater treatment. Environ. Chem. Lett. 2019, 17, 145–155. [Google Scholar] [CrossRef]
  95. Bobade, V.; Eshtiagi, N. Heavy metals removal from wastewater by adsorption process: A review. In Proceedings of the Asia Pacific Confederation of Chemical Engineering Congress, Melbourne, Australia, 27 September–1 October 2015. [Google Scholar]
  96. Chen, R.; Sheehan, T.; Ng, J.; Brucks, M.; Su, X. Capacitive deionization and electrosorption for heavy metal removal. Environ. Sci. Water Res. Technol. 2020, 6, 258–282. [Google Scholar] [CrossRef]
  97. Tang, W.; Wang, X.; Zeng, G.; Liang, J.; Li, X.; Xing, W.; He, D.; Tang, L.; Liu, Z. Electro-assisted adsorption of Zn(II) on activated carbon cloth in batch-flow mode: Experimental and theoretical investigations. Environ. Sci. Technol. 2019, 53, 2670–2678. [Google Scholar] [CrossRef]
  98. Li, Y.; Xu, R.; Qiao, L.; Li, Y.; Wang, D.; Li, D.; Liang, X.; Xu, G.; Gao, M.; Gong, H.; et al. Controlled synthesis of ZnO modified N-doped porous carbon nanofiber membrane for highly efficient removal of heavy metal ions by capacitive deionization. Microporous Mesoporous Mater. 2022, 338, 111889. [Google Scholar] [CrossRef]
  99. Zhang, X.; Yang, F.; Ma, J.; Liang, P. Effective removal and selective capture of copper from salty solution in flow electrode capacitive deionization. Environ. Sci. Water Res. Technol. 2020, 6, 341–350. [Google Scholar] [CrossRef]
  100. Xu, Y.; Xiang, S.; Zhang, X.; Zhou, H.; Zhang, H. High-performance pseudocapacitive removal of cadmium via synergistic valence conversion in perovskite-type FeMnO3. J. Hazard. Mater. 2022, 439, 129575. [Google Scholar] [CrossRef] [PubMed]
  101. Bao, S.; Chen, Q.; Zhang, Y.; Tian, X. Optimization of preparation conditions of composite electrodes for selective adsorption of vanadium in CDI by response surface methodology. Chem. Eng. Res. Des. 2021, 168, 37–45. [Google Scholar] [CrossRef]
  102. Wan, X.; Bao, S.; Zhang, Y. ZIF-8-derived porous carbon: Application in capacitive deionization for vanadium (V) adsorption. J. Appl. Electrochem. 2022, 52, 639–651. [Google Scholar] [CrossRef]
  103. Gabelich, C.; Tran, T.; Suffet, I. Electrosorption of inorganic salts from aqueous solution using carbon aerogels. Environ. Sci. Technol. 2002, 36, 3010–3019. [Google Scholar] [CrossRef]
  104. Gao, Y.; Pan, L.; Li, H.; Zhang, Y.; Zhang, Z.; Chen, Y.; Sun, Z. Electrosorption behavior of cations with carbon nanotubes and carbon nanofibres composite film electrodes. Thin Solid Film. 2009, 517, 1616–1619. [Google Scholar] [CrossRef]
  105. Huang, Z.; Lu, L.; Cai, Z.; Ren, Z. Individual and competitive removal of heavy metals using capacitive deionization. J. Hazard. Mater. 2016, 302, 323–331. [Google Scholar] [CrossRef] [PubMed]
  106. Li, H.; Zou, L.; Pan, L.; Sun, Z. Using graphene nano-flakes as electrodes to remove ferric ions by capacitive deionization. Sep. Purif. Technol. 2010, 75, 8–14. [Google Scholar] [CrossRef]
  107. He, R.; Neupane, M.; Zia, A.; Huang, X.; Bowers, C.; Wang, M.; Lu, J.; Yang, Y.; Dong, P. Binder-free wood converted carbon for enhanced water desalination performance. Adv. Funct. Mater. 2022, 32, 2208040. [Google Scholar] [CrossRef]
  108. Xu, Y.; Gao, Y.; Xiang, S.; Zhou, J.; Liu, F.; Li, Z.; Zhou, H. Selective pseudocapacitive separation of zinc ions via silk cocoon derived N-doped porous carbon. Desalination 2023, 546, 116220. [Google Scholar] [CrossRef]
  109. Zhang, W.; Mossad, M.; Yazdi, J.; Zou, L. A statistical experimental investigation on arsenic removal using capacitive deionization. Desalination Water Treat. 2016, 57, 3254–3260. [Google Scholar] [CrossRef]
  110. Liao, Y.; Wang, M.; Chen, D. Electrosorption of uranium(VI) by highly porous phosphate-functionalized graphene hydrogel. Appl. Surf. Sci. 2019, 484, 83–96. [Google Scholar] [CrossRef]
  111. Iftekhar, S.; Farooq, M.; Sillanpaa, M.; Asif, M.; Habib, R. Removal of Ni(II) using multi-walled carbon nanotubes electrodes: Relation between operating parameters and capacitive deionization performance. Arab. J. Sci. Eng. 2017, 42, 235–240. [Google Scholar] [CrossRef]
  112. Zhang, M.; Song, G.; Gelardi, D.; Huang, L.; Khan, E.; Masek, O.; Parikh, S.; Ok, Y. Evaluating biochar and its modifications for the removal of ammonium, nitrate, and phosphate in water. Water Res. 2020, 186, 116303. [Google Scholar] [CrossRef]
  113. Velusamy, K.; Periyasamy, S.; Kumar, P.; Vo, D.-V.; Sindhu, J.; Sneka, D.; Subhashini, B. Advanced techniques to remove phosphates and nitrates from waters: A review. Environ. Chem. Lett. 2021, 19, 3165–3180. [Google Scholar] [CrossRef]
  114. Karthikeyan, P.; Elanchezhiyan, S.; Preethi, J.; Meenakshi, S.; Park, C. Mechanistic performance of polyaniline-substituted hexagonal boron nitride composite as a highly efficient adsorbent for the removal of phosphate, nitrate, and hexavalent chromium ions from an aqueous environment. Appl. Surf. Sci. 2020, 511, 145543. [Google Scholar] [CrossRef]
  115. Nodeh, H.; Sereshti, H.; Afsharian, E.; Nouri, N. Enhanced removal of phosphate and nitrate ions from aqueous media using nanosized lanthanum hydrous doped on magnetic graphene nanocomposite. J. Environ. Manag. 2017, 197, 265–274. [Google Scholar] [CrossRef] [PubMed]
  116. Priya, E.; Kumar, S.; Verma, C.; Sarkar, S.; Maji, P. A comprehensive review on technological advances of adsorption for removing nitrate and phosphate from waste water. J. Water Process Eng. 2022, 49, 103159. [Google Scholar] [CrossRef]
  117. Berkessa, Y.; Mereta, S.; Feyisa, F. Simultaneous removal of nitrate and phosphate from wastewater using solid waste from factory. Appl. Water Sci. 2019, 9, 28. [Google Scholar] [CrossRef]
  118. Cai, T.; Park, S.; Li, Y. Nutrient recovery from wastewater streams by microalgae: Status and prospects. Renew. Sustain. Energy Rev. 2013, 19, 360–369. [Google Scholar] [CrossRef]
  119. Huang, H.; Xiao, X.; Yan, B.; Yang, L. Ammonium removal from aqueous solutions by using natural Chinese (Chende) zeolite as adsorbent. J. Hazard. Mater. 2010, 175, 247–252. [Google Scholar] [CrossRef]
  120. Gizaw, A.; Zewge, F.; Kumar, A.; Mekonnen, A.; Tesfaye, M. A comprehensive review on nitrate and phosphate removal and recovery from aqueous solutions by adsorption. AQUA-Water Infrastruct. Ecosyst. Soc. 2021, 70, 921–947. [Google Scholar] [CrossRef]
  121. Chen, H.; Zhang, H.; Tian, J.; Shi, J.; Linhardt, R.; Ye, T.; Chen, S. Recovery of high value-added nutrients from fruit and vegetable industrial wastewater. Compr. Rev. Food Sci. Food Saf. 2019, 18, 1388–1402. [Google Scholar] [CrossRef] [Green Version]
  122. Meena, V. Role of Rhizospheric Microbes in Soil; Springer: Berlin/Heidelberg, Germany, 2018. [Google Scholar]
  123. Zhang, H.; Wang, Q.; Zhang, J.; Chen, G.; Wang, Z.; Wu, Z. Development of novel ZnZr-COOH/CNT composite electrode for selectively removing phosphate by capacitive deionization. Chem. Eng. J. 2022, 439, 135527. [Google Scholar] [CrossRef]
  124. Zhang, J.; Tang, L.; Tang, W.; Zhong, Y.; Luo, K.; Duan, M.; Xing, W.; Liang, J. Removal and recovery of phosphorus from low-strength wastewaters by flow-electrode capacitive deionization. Sep. Purif. Technol. 2020, 237, 116322. [Google Scholar] [CrossRef]
  125. Bian, Y.; Chen, X.; Ren, Z. pH dependence of phosphorus speciation and transport in flow-electrode capacitive deionization. Environ. Sci. Technol. 2020, 54, 9116–9123. [Google Scholar] [CrossRef]
  126. Benjamin, M.M. Water Chemistry; Waveland Press: Long Grove, IL, USA, 2014. [Google Scholar]
  127. Tian, J.; Cheng, X.; Deng, S.; Liu, J.; Qu, B.; Dang, Y.; Holmes, D.; Waite, T. Inducing in situ crystallization of vivianite in a UCT-MBR system for enhanced removal and possible recovery of phosphorus from sewage. Environ. Sci. Technol. 2019, 53, 9045–9053. [Google Scholar] [CrossRef] [PubMed]
  128. Wu, Y.; Luo, J.; Zhang, Q.; Aleem, M.; Fang, F.; Xue, Z.; Cao, J. Potentials and challenges of phosphorus recovery as vivianite from wastewater: A review. Chemosphere 2019, 226, 246–258. [Google Scholar] [CrossRef] [PubMed]
  129. Zhang, C.; Cheng, X.; Wang, M.; Ma, J.; Collins, R.; Kinsela, A.; Zhang, Y.; Waite, T. Phosphate recovery as vivianite using a flow-electrode capacitive desalination (FCDI) and fluidized bed crystallization (FBC) coupled system. Water Res. 2021, 194, 116939. [Google Scholar] [CrossRef]
  130. Fateminia, R.; Rowshanzamir, S.; Mehri, F. Synergistically enhanced nitrate removal by capacitive deionization with activated carbon/PVDF/polyaniline/ZrO2 composite electrode. Sep. Purif. Technol. 2021, 274, 119108. [Google Scholar] [CrossRef]
  131. Cetinkaya, A. Life cycle assessment of environmental effects and nitrate removal for membrane capacitive deionization technology. Environ. Monit. Assess. 2020, 192, 543. [Google Scholar] [CrossRef]
  132. Garcia-Quismondo, E.; Santos, C.; Soria, J.; Palma, J.; Anderson, M. New operational modes to increase energy efficiency in capacitive deionization systems. Environ. Sci. Technol. 2016, 50, 6053–6060. [Google Scholar] [CrossRef] [PubMed]
  133. Hu, C.; Dong, J.; Wang, T.; Liu, R.; Liu, H.; Qu, J. Nitrate electro-sorption/reduction in capacitive deionization using a novel Pd/NiAl-layered metal oxide film electrode. Chem. Eng. J. 2018, 335, 475–482. [Google Scholar] [CrossRef]
  134. Rogers, T.; Guo, S.; Arrazolo, L.; Garcia-Segura, S.; Wong, M.; Verduzco, R. Catalytic capacitive deionization for adsorption and reduction of aqueous nitrate. ACS EST Water 2021, 1, 2233–2241. [Google Scholar] [CrossRef]
  135. Gao, F.; Li, X.; Shi, W.; Wang, Z. Highly selective recovery of phosphorus from wastewater via capacitive deionization enabled by ferrocene-polyaniline-functionalized carbon nanotube electrodes. ACS Appl. Mater. Interfaces 2022, 14, 31962–31972. [Google Scholar] [CrossRef] [PubMed]
  136. Xu, L.; Yu, C.; Tian, S.; Mao, Y.; Zong, Y.; Zhang, X.; Zhang, B.; Zhang, C.; Wu, D. Selective recovery of phosphorus from synthetic urine using flow-electrode capacitive deionization (FCDI)-based technology. ACS EST Water 2020, 1, 175–184. [Google Scholar] [CrossRef]
  137. Zhang, C.; Wang, M.; Xiao, W.; Ma, J.; Sun, J.; Mo, H.; Waite, T. Phosphate selective recovery by magnetic iron oxide impregnated carbon flow-electrode capacitive deionization (FCDI). Water Res. 2021, 189, 116653. [Google Scholar] [CrossRef] [PubMed]
  138. Chen, J.; Zuo, K.; Li, Y.; Huang, X.; Hu, J.; Yang, Y.; Wang, W.; Chen, L.; Jain, A.; Verduzco, R.; et al. Eggshell membrane derived nitrogen rich porous carbon for selective electrosorption of nitrate from water. Water Res. 2022, 216, 118351. [Google Scholar] [CrossRef] [PubMed]
  139. Chen, L.; He, F.; Li, F. Denitrification enhancement by electro-adsorption/reduction in capacitive deionization (CDI) and membrane capacitive deionization (MCDI) with copper electrode. Chemosphere 2022, 291, 132732. [Google Scholar] [CrossRef] [PubMed]
  140. Minhas, M.; Jande, Y.; Kim, W. Hybrid reverse osmosis-capacitive deionization versus two-stage reverse osmosis: A comparative analysis. Chem. Eng. Technol. 2014, 37, 1137–1145. [Google Scholar] [CrossRef]
  141. Chen, W.; He, X.; Jiang, Z.; Li, B.; Li, X.; Lin, L. A capacitive deionization and electro-oxidation hybrid system for simultaneous removal of heavy metals and organics from wastewater. Chem. Eng. J. 2023, 451, 139071. [Google Scholar] [CrossRef]
  142. Liang, S.; Lia, M.; Cao, J.; Zuo, K.; Bian, Y.; Xiao, K.; Huang, X. Integrated ultrafiltration capacitive-deionization (UCDI) for enhanced antifouling performance and synchronous removal of organic matter and salts. Sep. Purif. Technol. 2019, 226, 146–153. [Google Scholar] [CrossRef]
  143. Duan, F.; Li, Y.; Cao, H.; Wang, Y.; Crittenden, J.; Zhang, Y. Activated carbon electrodes: Electrochemical oxidation coupled with desalination for wastewater treatment. Chemosphere 2015, 125, 205–211. [Google Scholar] [CrossRef]
  144. Martinez-Huitle, C.; Ferro, S. Electrochemical oxidation of organic pollutants for the wastewater treatment: Direct and indirect processes. Chem. Soc. Rev. 2006, 35, 1324–1340. [Google Scholar] [CrossRef]
  145. Sun, X.; Chen, M.; Wei, D.; Du, Y. Research progress of disinfection and disinfection by-products in China. J. Environ. Sci. 2019, 81, 52–67. [Google Scholar] [CrossRef]
  146. Krasner, S.W. The formation and control of emerging disinfection by-products of health concern. Philos. Trans. R. Soc. Math. Phys. Eng. Sci. 2009, 367, 4077–4095. [Google Scholar] [CrossRef] [Green Version]
  147. Richardson, S.; Kimura, S. Emerging environmental contaminants: Challenges facing our next generation and potential engineering solutions. Environ. Technol. Innov. 2017, 8, 40–56. [Google Scholar] [CrossRef]
  148. Yang, Y.; Komaki, Y.; Kimura, S.; Hu, H.; Wagner, E.; Marinas, B.; Plewa, M. Toxic impact of bromide and iodide on drinking water disinfected with chlorine or chloramines. Environ. Sci. Technol. 2014, 48, 12362–12369. [Google Scholar] [CrossRef] [PubMed]
  149. Richardson, S.; Plewa, M.; Wagner, E.; Schoeny, R.; DeMarini, D. Occurrence, genotoxicity, and carcinogenicity of regulated and emerging disinfection by-products in drinking water: A review and roadmap for research. Mutat. Res. -Rev. Mutat. Res. 2007, 636, 178–242. [Google Scholar] [CrossRef] [PubMed]
  150. Li, C.; Wang, D.; Xu, X.; Wang, Z. Formation of known and unknown disinfection by-products from natural organic matter fractions during chlorination, chloramination, and ozonation. Sci. Total Environ. 2017, 587, 177–184. [Google Scholar] [CrossRef]
  151. Wang, Y.; El-Deen, A.; Li, P.; Oh, B.; Guo, Z.; Khin, M.; Vikhe, Y.; Wang, J.; Hu, R.; Boom, R.; et al. High-performance capacitive deionization disinfection of water with graphene oxide-graft-quaternized chitosan nanohybrid electrode coating. ACS Nano 2015, 9, 10142–10157. [Google Scholar] [CrossRef] [Green Version]
  152. Laxman, K.; Myint, M.; Al Abri, M.; Sathe, P.; Dobretsov, S.; Dutta, J. Desalination and disinfection of inland brackish ground water in a capacitive deionization cell using nanoporous activated carbon cloth electrodes. Desalination 2015, 362, 126–132. [Google Scholar] [CrossRef]
  153. Wang, G.; Wang, J.; Wang, D.; Yuan, Y.; Wang, L.; Zhao, Z.; Qiu, J. Contact-active antimicrobial graphene oxide electrodes by covalent grafting of alkoxysilane octadecyldimethyl (3-trimethoxysilylpropyl)-ammonium chloride (ODDMAC) for capacitive deionization disinfection. Mater. Lett. 2020, 272, 127876. [Google Scholar] [CrossRef]
  154. Cao, C.; Wu, X.; Zheng, Y.; Chen, Y. Three-dimensional cubic ordered mesoporous carbon with chitosan for capacitive deionization disinfection of water. Environ. Sci. Pollut. Res. 2020, 27, 15001–15010. [Google Scholar] [CrossRef]
  155. Kang, J.; Kim, T.; Jo, K.; Yoon, J. Comparison of salt adsorption capacity and energy consumption between constant current and constant voltage operation in capacitive deionization. Desalination 2014, 352, 52–57. [Google Scholar] [CrossRef]
  156. Chen, Y.; Chen, J.; Lin, C.; Hou, C. Integrating a supercapacitor with capacitive deionization for direct energy recovery from the desalination of brackish water. Appl. Energy 2019, 252, 113417. [Google Scholar] [CrossRef]
  157. Andres, G.; Yoshihara, Y. A capacitive deionization system with high energy recovery and effective re-use. Energy 2016, 103, 605–617. [Google Scholar] [CrossRef]
  158. Chen, L.; Yin, X.; Zhu, L.; Qiu, Y. Energy recovery and electrode regeneration under different charge/discharge conditions in membrane capacitive deionization. Desalination 2018, 439, 93–101. [Google Scholar] [CrossRef]
  159. Zhao, R.; Satpradit, O.; Rijnaarts, H.; Biesheuvel, P.; van der Wal, A. Optimization of salt adsorption rate in membrane capacitive deionization. Water Res. 2013, 47, 1941–1952. [Google Scholar] [CrossRef]
  160. Choi, J.H. Comparison of constant voltage (CV) and constant current (CC) operation in the membrane capacitive deionisation process. Desalination Water Treat. 2015, 56, 921–928. [Google Scholar] [CrossRef]
  161. Qu, Y.; Campbell, P.; Gu, L.; Knipe, J.; Dzenitis, E.; Santiago, J.; Stadermann, M. Energy consumption analysis of constant voltage and constant current operations in capacitive deionization. Desalination 2016, 400, 18–24. [Google Scholar] [CrossRef] [Green Version]
  162. Lee, H.; Jin, Y.; Hong, S. Recent transitions in ultrapure water (UPW) technology: Rising role of reverse osmosis (RO). Desalination 2016, 399, 185–197. [Google Scholar] [CrossRef]
  163. Choi, J.; Kim, J.; Hong, S. Staged voltage mode in membrane capacitive deionization: Comparison with constant voltage and constant current modes. Desalination 2020, 479, 114327. [Google Scholar] [CrossRef]
Figure 2. Architecture diagrams and removal mechanisms of (a) CDI, (b) MCDI, (c) FCDI, and (d) HCDI.
Figure 2. Architecture diagrams and removal mechanisms of (a) CDI, (b) MCDI, (c) FCDI, and (d) HCDI.
Energies 16 01136 g002
Figure 6. Source and control of nutrients (N and P) in the water environment [112]. Copyright 2020, Elsevier.
Figure 6. Source and control of nutrients (N and P) in the water environment [112]. Copyright 2020, Elsevier.
Energies 16 01136 g006
Figure 8. Removal process of Cu2+ and Acid Orange 7 by CDI-EO system [141]. Copyright 2022, Elsevier.
Figure 8. Removal process of Cu2+ and Acid Orange 7 by CDI-EO system [141]. Copyright 2022, Elsevier.
Energies 16 01136 g008
Figure 9. Schematic diagram of energy recovery system and charging based on our-switch buck-boost converter [156]. Copyright 2019, Elsevier.
Figure 9. Schematic diagram of energy recovery system and charging based on our-switch buck-boost converter [156]. Copyright 2019, Elsevier.
Energies 16 01136 g009
Figure 10. Schematic diagram of SV mode in MCDI [163]. Copyright 2020, Elsevier.
Figure 10. Schematic diagram of SV mode in MCDI [163]. Copyright 2020, Elsevier.
Energies 16 01136 g010
Table 1. Removal efficiency of hardness ions by CDI.
Table 1. Removal efficiency of hardness ions by CDI.
Hardness Ion/ConcentrationApplied Voltage (V)Flow Rate (mL/min)Removal EfficiencyCDI Type/ElectrodeRef.
100 mg/L CaCl21.42542.8 mg/gHCDI with copper hexacyanoferrate as cathode electrode[73]
10 mM CaCl21.210248 μmol/gCDI with porous mordenite modified activated carbon electrode[74]
10 mM CaCl21.2214.2 mg/gCa-alginate coated on carbon electrode[75]
0.95 g/L CaCl21.42521.0 mg/gCDI of KOH carbonized and activated carbon electrode[76]
150 mg/L CaCl21.220534.6 μmol/gPHCI with manganese spinel ferrite as cathode electrode[77]
75 mg/L MgSO4980.4 μmol/g
350 mg/L CaCO31.51680%CDI with activated carbon cloth electrode[78]
35 mg/L CaCO32103.5 mg/gCDI with purified reduced graphene oxide electrodes[79]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Bao, S.; Xin, C.; Zhang, Y.; Chen, B.; Ding, W.; Luo, Y. Application of Capacitive Deionization in Water Treatment and Energy Recovery: A Review. Energies 2023, 16, 1136. https://doi.org/10.3390/en16031136

AMA Style

Bao S, Xin C, Zhang Y, Chen B, Ding W, Luo Y. Application of Capacitive Deionization in Water Treatment and Energy Recovery: A Review. Energies. 2023; 16(3):1136. https://doi.org/10.3390/en16031136

Chicago/Turabian Style

Bao, Shenxu, Chunfu Xin, Yimin Zhang, Bo Chen, Wei Ding, and Yongpeng Luo. 2023. "Application of Capacitive Deionization in Water Treatment and Energy Recovery: A Review" Energies 16, no. 3: 1136. https://doi.org/10.3390/en16031136

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop