Next Article in Journal
Fast Identification Method of Mine Water Source Based on Laser-Induced Fluorescence Technology and Optimized LSTM
Next Article in Special Issue
Odorous Substances in Urban Drainage Pipelines and the Removal Technology: A Review
Previous Article in Journal
Characteristics of Debris Flow Activities at Different Scales after the Disturbance of Strong Earthquakes—A Case Study of the Wenchuan Earthquake-Affected Area
Previous Article in Special Issue
A Review of Recent Advances in Detection and Treatment Technology for Perfluorinated Compounds
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review of Research Progress in the Preparation and Application of Ferrate(VI)

1
School of Environment and Architecture, University of Shanghai for Science and Technology, Shanghai 200093, China
2
Yk Pao School, Shanghai 201600, China
3
State Key Laboratory of Pollution Control and Resources Reuse, College of Environmental Science and Engineering, Tongji University, Shanghai 200092, China
*
Author to whom correspondence should be addressed.
Water 2023, 15(4), 699; https://doi.org/10.3390/w15040699
Submission received: 1 January 2023 / Revised: 20 January 2023 / Accepted: 29 January 2023 / Published: 10 February 2023

Abstract

:
Ferrate(VI) is a green and efficient water treatment agent for drinking and wastewater. It is widely used in water treatment because it has multi-functional uses such as oxidation, algae removal, disinfection, and adsorption flocculation. It does not cause secondary pollution to the environment. This paper compares ferrate(VI) with other water treatment agents and discusses three methods of preparing ferrate(VI). The removal, adsorption, and control of organic matter, algae, disinfection by-products, and heavy metal ions in water when ferrate(VI) was used as an oxidant, disinfectant, and coagulant were summarized. Ferrate(VI) has some advantages in removing toxic, harmful, and difficult-to-degrade substances from water. Due to the disadvantages of ferrate(VI) such as oxidation selectivity and instability, it is necessary to develop the hyphenated techniques of ferrate(VI). In this review, three hyphenated techniques of ferrate(VI) are summarized: ferrate(VI)–photocatalytic synergistic coupling, ferrate(VI)–PAA synergistic coupling, and ferrate(VI)–PMS synergistic coupling.

1. Introduction

With the rapid development of urbanization and industrialization, the demand for commercial products has also increased dramatically. However, some unknown toxic pollutants and their usage cause environmental issues yearly [1,2]. For example, various micropollutants have been detected in surface water and certain new pollutants (such as toxic hormone chemicals, skin care products, and toxins) [3,4]. Although the levels of these pollutants in water are low, they can have a negative impact on human health, ecosystems, and biodiversity [5,6].
Recently, the most common methods for treating pollutants in the water environment are physical adsorption [7,8], biodegradation [9,10], and advanced oxidation [5,11]. The physical adsorption technique is mainly used to remove pollutants from the water due to the excellent adsorption ability of the adsorbent material, which may enrich the pollutants in the adsorbent. There are two organic and carbon-based adsorbents [12], mostly used in treating aqueous samples [13]. Although the adsorbed material has the advantages of low cost and good adsorption performance, the removal of pollutants by physical adsorption is based mainly on enrichment rather than degradation. Therefore, physical adsorption requires secondary treatment if it is not completely removed in treating pollutants. It is easy to cause recontamination problems, which hinders the wide application of adsorption materials. Compared to physical adsorption, biological treatment can completely remove contaminants. The biological treatment method is mainly based on the ability of microorganisms to use water pollutants as organic carbon sources, and realize the decomposition and transformation of pollutants through the metabolism of the microorganisms themselves. However, the effect of biological treatment is easily affected by environmental factors, the environmental requirements are high, the treatment cycle is long, and the treatment of antibiotic-contaminated water bodies is prone to the production of resistant bacteria and genes; additionally, there is a risk of biosafety [14]. Compared with the first two, advanced oxidation methods produce hydroxyl radicals (·OH) with strong oxidizing capacity characteristics. Under certain reaction conditions, the pollutants are directly mineralized or improved by oxidation, so that macromolecular refractory organic matter is oxidized into low-toxicity or non-toxic small molecules, which has good application prospects [15,16]. However, the advanced oxidation technique has various disadvantages in the application of pollution removals, such as complex operation, high cost, and difficult operating parameters. Additionally, the treatment impact is easily influenced by several variables, and the effect is unstable [17].
Compared to the advanced oxidation method, the green chemical ferrate(VI) has several advantages in treating pollutants from wastewater [18,19,20]. Ferrate(VI) is a hexavalent iron salt with a high redox potential that exhibits high oxidation and bactericidal disinfection capabilities throughout the treatment process [21] and can also adsorb the particles and cause flocculation [22]; it does not cause secondary pollution and has been recognized as a green multi-functional water treatment agent.
This review summarizes the properties and applications of ferrate(VI), including the preparation method of ferrate(VI) and the removal, control, and adsorption of organic matter, algae, disinfection by-products, and heavy metal ions in water when ferrate(VI) is used as an oxidant, disinfectant, and coagulant. At the same time, the combination of ferrate(VI) and other substances is described, which can provide a reference for the further application of ferrate(VI).

2. Properties of Ferrate(VI)

2.1. Structure of Ferrate(VI)

The iron valence state in ferrate(VI) is positive hexavalent, which has strong oxidation. Among them, the UV-visible spectrum of potassium ferrate(VI) has a characteristic peak absorption at 510 nm and 800 nm. Its Infrared Spectroscopy (IR) also has a strong absorption peak at 800 cm−1 and a shoulder peak at 778 cm−1, which is the characteristic peak of Fe–O bond telescopic vibration in potassium ferrite, so infrared spectroscopy can be used to qualitatively and quantitatively determine potassium ferrate(VI). Figure 1 shows the slightly distorted tetrahedral structure of FeO42− in solid ferrate(VI), determined using X-ray photoelectron spectroscopy (XPS) and infrared spectroscopy (IR) [23,24].

2.2. Stability of Ferrate(VI)

The stability of ferrate(VI) is one of the problems limiting its large-scale preparation and application. Because ferrate(VI) is poorly stable and difficult to store, most of the current applied research on ferrate(VI) is limited to the laboratory. The storage environment requirements for ferrate(VI) are more stringent and require low temperature and dry storage. Once removed from the dry environment, ferrate(VI) is highly susceptible to the influence of water molecules. Its stability will rapidly decrease and oxidative decomposition will occur, with the following reaction Equation (1).
4 F e O 4 2 - + 10 H 2 O 4 F e ( O H ) 3 + 3 O 2 + O H -
The main factors affecting the stability of ferrate(VI) include the temperature of the solution, pH, and the coexistence of ions in the solution. The temperature has an important influence on the stability of ferrate(VI); the higher the temperature, the faster the decomposition rate of ferrate(VI) [25], so it is necessary to maintain a relatively low temperature in the process of electrolysis and oxidation preparation of ferrate(VI). Ferrate(VI) exists in four main forms, including H3FeO4+, H2FeO4, HFeO4, and FeO42−. Under acidic conditions, H3FeO4+ and H2FeO4 are the main forms, while HFeO4 and FeO42− are the main forms under neutral or alkaline conditions. FeO42− is extremely unstable and easy to decompose under an acidic environment, while FeO42− has strong stability when pH is controlled at 9~10. The main reaction equations for FeO42− under acid–base conditions are shown in Equations (2) and (3) below.
A c i d i c : F e O 4 2 - + 8 H + + 3 e - F e 3 + + 4 H 2 O
A l k a l i n e : F e O 4 2 - + 4 H 2 O + 3 e - F e ( O H ) 3 + 5 O H -
In addition to the above influencing factors, the components in the ferrate(VI) solution can also affect the rate of its decomposition. When the ferrate(VI) solution contains phosphate, it can hinder the decomposition rate of the ferrate(VI). At the same time, light also has a particular influence on the stability of ferrate(VI) [23]; ultraviolet light irradiation will accelerate the decomposition rate of ferrate(VI), so attention should be paid to avoiding light during storage.

2.3. Oxidation of Ferrate(VI)

Ferrate(VI) is strongly oxidizing and its redox potential under acidic conditions is 2.20 V, which is higher than some common oxidants such as ozone (O3), hydrogen peroxide (H2O2), chlorine (Cl2), chlorine dioxide (ClO2), and potassium permanganate (KMnO4) [26], and a comparison of redox potentials are shown in Table 1.
It was found that the strong oxidizing property of ferrate(VI) can effectively remove NH3, CN, and SCN, as well as carboxylic acid, phenol, amino acid, and organic nitrogen compounds from water [27]. Sailo et al. [28] used ferrate(VI) to remove Bisphenol A, and when the concentration ratio of ferrate(VI) to pollutants was increased to 3:1, the removal rate of Bisphenol A reached 100%, and the apparent rate constant was 8.35 × 102 M−1min−1. Additionally, with the increase in the amount of ferrate(VI) dosage, the removal effect will also increase. At the same time, the strong oxidation of ferrate(VI) can destroy the cell wall of bacteria and algae cells, and play a role in sterilization and algae. Zhou et al. [29] found that ferrate(VI) pre-oxidation can remove algae cells by destroying the protective organic layer. Therefore, ferrate(VI) is also an effective bactericidal disinfectant and algaecide. At the same time, when used as a disinfectant, ferrate(VI) is not like disinfection by-products (DBPs) that are harmful to the human body and produced during chlorination treatment, such as trihalomethanes (THMs); on the contrary, ferrate(VI) can effectively control the formation of chlorination disinfection by-products [30].

3. Preparation of Ferrate(VI)

Based on the performance and application advantages of ferrate(VI), the efficient preparation of ferrate(VI) is also essential. At present, the preparation methods of ferrate(VI) mainly include the dry oxidation method [31], the wet oxidation method [32], and the electrochemical method [33]. Among them, the electrochemical method has a simple operation process, simple preparation, and high yield [31].

3.1. Dry Oxidation

The dry oxidation method, also known as the solid-phase melting method, involves the formation of ferrate(VI) by combining iron filings or iron compounds with caustic alkali and peroxides (such as Na2O2) in a high-temperature environment. E. Martinez-Tamayo et al. [34] used Na2O2 and FeSO4 to prepare sodium ferrate(VI) and then mixed it with a saturated potassium hydroxide solution and purified it to obtain solid potassium ferrate(VI). Talaiekhozani et al. [31] improved the reaction conditions. They found that, in the absence of O2 and CO2, Na2O2 reacts with iron oxide at a temperature of 370 °C to form sodium ferrate(VI), as shown in Equation (4). The reaction under these conditions can reduce the possibility of explosion. L. Ninane et al. [35] reduced the cost of the preparation process to meet the needs in the actual water. Moreover, ferrous sulfate, potassium hydroxide, and calcium hypochlorite (or chlorine) are used as raw materials and the reaction time is shortened. The main raw material, ferrous sulfate, is a by-product of the titanium dioxide production process, so the cost of preparation can be reduced. The dry oxidation method can generate large quantities of high ferrate(VI), still, it is carried out under high temperature and high-pressure conditions, and most of the reaction processes are exothermic, which increases the risk of explosion. The reaction also has high requirements for the alkaline corrosion resistance of the equipment, resulting in high preparation costs and difficulty in achieving industrial production [36].
3 N a 2 O 2 + F e 2 O 3 370   ° C 2 N a 2 F e O 4 + N a 2 O

3.2. Wet Oxidation

Wet oxidation is also known as hypochlorite oxidation. It is a process in which trivalent iron ions are oxidized to ferrate(VI) using hypochlorite in a concentrated alkali solution. To prepare a high concentration of sodium ferrate(VI) solution, a saturated KOH solution is poured into the solution, which precipitates potassium ferrate(VI) due to its lower solubility compared to sodium ferrate(VI). The reaction equation is shown in Equations (5)–(7) [37,38].
F e 3 + + 3 O H - F e O H 3
2 F e O H 3 + 3 N a C l O + 4 N a O H 2 N a 2 F e O 4 + 3 N a C l + 5 H 2 O
N a 2 F e O 4 + 2 K O H K 2 F e O 4 + 2 N a O H
As early as 1950, the preparation of ferrate(VI) using the wet method was reported [39], and the purity also reached 96.9%, but the yield of ferrate(VI) was less than 15%. Wet oxidation is one of the commonly used methods for the preparation of ferrate(VI) because it uses easily available raw materials and has a simple process. However, the wet oxidation method also has disadvantages such as: the need for large reactors for the reaction, high consumption of alkali, high preparation cost and low yield of ferrate(VI) [33], which makes it challenging to use on large-scale applications. To improve the product and reduce the cost of the preparation of ferrate(VI) by wet oxidation, further research is necessary. Tiwari et al. [40] used strong alkali solutions of potassium hydroxide to replace sodium hydroxide, simplifying the wet oxidation process and improving the yield and purity to 70% and 99%, respectively. Similarly, Tong et al. [32] designed a novel reactor that can control the time and temperature of the reaction, and the concentration of ferrate(VI) generated by the new device, which is 7.6 × 10−2 M compared to conventional reaction units, is twice as much as the old unit. The wet oxidation method requires the passage of chlorine gas into the reactor to prepare a saturated hypochlorite solution, which presents specific safety hazards in the preparation process. To address this problem, some scholars have turned their attention to the use of persulfate ions instead of hypochlorite ions to oxidize Fe(III) to ferrate(VI) using the free radicals generated by persulfate activation [41]. Although the purity of ferrate(VI) prepared by the wet oxidation method reaches 96%, there are still problems of low yield and high preparation cost, so a green and efficient preparation method is urgently needed.

3.3. Electrochemical Method

Compared with the preparation of dry and wet oxidation methods, electrochemical methods have attracted much attention because of their simple operation, simple process, and high product purity. The principle of this method is that under the solution condition of strong alkali, the anode material (pure iron, cast iron, and wrought iron) is consumed, causing the anode to undergo oxidation in the superpassivation potential region and dissolve to form ferrate(VI). The main reaction formula in the electrolysis process is shown in Equations (8)–(10) [42,43].
A n o d e r e a c t i o n s : F e + 8 O H - F e O 4 2 - + 4 H 2 O + 6 e -
C a t h o d e r e a c t i o n : 2 H 2 O + 2 e - H 2 + 2 O H -
G e n e r a l r e a c t i o n : F e + 2 O H - + 2 H 2 O F e O 4 2 - + 3 H 2
During the electrolytic preparation of ferrate(VI), the anode loses electrons and generates ferrate(VI). Therefore, the anode material’s composition can affect ferrate preparation (VI). Talaiekhozani et al. [31] found that by increasing the iron anode material’s carbon content, the prepared ferrate’s concentration and electrolytic efficiency (VI) improved. By using iron rods with 0.9% carbon content as anodes, they achieved an electrolysis efficiency of 70%, which was six times higher than iron rods, with only 0.08% carbon content. By increasing the carbon content of the iron anode, small holes in the electrode are formed, which promotes iron precipitation. In addition to the anode material’s effect on the efficiency of ferrate(VI) generation, the anode morphology also affects the concentration of ferrate(VI) generated [44]. Ding et al. [45] used a porous magnetic iron electrode as the anode, and the results showed that the electrolysis efficiency was as high as 52.3% under the same conditions. This porous magnetic iron electrode not only has a high carbon content, but also contains a large number of small iron electrode pores, which are more conducive to the precipitation of iron and help to slow down the passivation of the electrode. This promotes the generation of high ferrate(VI). In addition, when the anode material contains some non-metallic elements, the preparation effect can be improved. Diaz et al. [46] found that the presence of silicon in cast iron slowed down the anodic passivation process when studying the effect of different electrode materials on the formation of the passivation layer during the generation of generating ferrate(VI). The concentration of lye is also very important for the production of ferrate(VI) [47]. Sibel et al. [48] explored the effect of lye concentration on electrolytic preparation and experimented with different concentrations of NaOH, and the ferrate(VI) concentration increased with the increase of lye concentration. However, the optimal concentration was 16 mol/L, and the resulting ferrate(VI) concentration was 1.98 mM. Although the higher concentration of lye can increase the concentration of ferrate(VI), at the same time, the increase in lye concentration will increase the viscosity of the electrolyte, causing the electrons to overcome the resistance of the electrolyte, which increases energy consumption, reduces the transfer efficiency, and inhibits the formation of ferrate(VI), thereby increasing the preparation cost.
In the process of electrolysis, it is necessary to reduce costs and energy consumption. Yang et al. [49] used a gas diffusion electrode (GDE) for cathodic electrolysis to prepare ferrate(VI), which can increase the cathode potential and reduce the tank pressure, thereby reducing the energy consumption in production, which is 31.96–56.67% compared to the traditional electrolysis method. However, as the electrolysis time increases, a passivation layer is formed on the surface of the anode, which hinders the formation of ferrate(VI) [50]. The use of periodic intermittent reverse current electrolysis instead of conventional continuous electrolysis with a DC power supply can slow down passivation. Pi et al. [51] used periodic reverse current to prepare sodium ferrate(VI), which increased the rate of ferrate(VI) synthesis and the concentration of the prepared sodium ferrate(VI) by 75.0%.

4. Application of Ferrate(VI) in Water Treatment

As a water treatment agent, ferrate(VI) is widely used, and Figure 2 shows a schematic diagram of the treatment of wastewater by electrochemically prepared high ferrate [18]. Ferrate(VI) is a hexavalent iron salt with strong oxidation, and redox potential up to 2.20 V under acidic conditions, and its oxidation performance is higher than other common strong oxidants (such as ozone, chlorine, and potassium permanganate). The redox product of ferrate(VI) is a non-toxic and harmless ferric iron salt that can be used as a flocculant to improve the removal of pollutants [52], and does not produce toxic by-products compared to chlorine and ozone treatment [53], so it also has good potential for use in pollutant degradation and removal. Ferrate(VI) is often used as an oxidizing agent to remove pharmaceuticals and personal care products, phenolic compounds, and algae.

4.1. Ferrate(VI) as an Oxidant

4.1.1. PPCPs

Pharmaceuticals and personal care products (PPCPs) are an emerging class of pollutants that include drugs such as antibiotics, β-blockers, antiepileptic drugs, and daily care products such as skincare, cosmetics, hairspray, and hair dyes [54]. Although the concentration of PPCPs in the environment is low, there are significant safety risks. When PPCPs enter the human body through the food chain, they can cause harm to human health, such as carcinogenic, teratogenic and mutagenic [55]. Ferrate(VI) is commonly used to remove PPCPs from water. It has been shown to have a good removal effect on contaminants such as theophylline [56], metoprolol [57], carbamazepine (CBZ) [58], and sulfachloropyridazine (SCP) [32] from water.
SCP belongs to the sulfonamide class of antibiotics (SA), and ferrate(VI) has also been effective in removing sulfonamide antibiotics [32]. SA is composed of heterocyclic N aromatic rings linked by aniline and sulfa groups. Acosta-Rangel et al. [59] found that SA can be completely removed within 5 min when the molar ratio of ferrate(VI) to SA is 6:1 and the pH is 3. Similarly, Pi et al. [51] used the ferrate(VI) to remove SCP, and at a molar ratio of 15:1, the removal rate of SCP was as high as 90% under neutral conditions. The degradation of SCP by ferrate(VI) is divided into two pathways: first, the SO2 in the sulfonamide group is converted to SO42−, and the intermediate products of ferrate(VI), Fe(IV), and Fe(V) can also convert SO23− to SO42−; second, the -NH2 in the aniline group is oxidized to -NHOH and -NO, and finally to -NO2 [60]. Additionally, the oxidation of sulfonamide groups in SCP by ferrate(VI) can reduce the toxicity of SCP, and the degradation schematic is shown in Figure 3.
Ferrate(VI) has a high removal rate for electron-rich pollutants in PPCPs, such as phenol, olefins, amines, and aniline moieties. It has been found that roxithromycin and oleandomycin in macrolide antibiotics contain amine groups, which are easily oxidized by ferrate(VI) compared to other antibiotics. Therefore, ferrate(VI) has a better removal rate for antibiotics containing amine structures [61]. Similarly, tetracyclines, chlortetracycline, and oxytetracycline have phenol and olefin groups and are therefore easily removed by ferrate(VI) oxidation [62].
However, the removal of pollutants by ferrate(VI) is susceptible to the pH. Ma et al. [63] found that, in the degradation of tetracycline hydrochloride (TCH) using ferrate(VI), the pH of the reaction solution had a large effect on the removal of TCH, and lowering the pH of the reaction solution could improve the removal rate of TCH with an optimal pH of 9–10 and a molar ratio of ferrate(VI) to TCH of 1:10, resulting in the removal rate of TCH being 98.57%. Similarly, when ferrate(VI) is used to degrade PPCPs, the removal rate can be improved by reducing the pH from 12 to 7 [28]. This is because the redox potential of ferrate(VI) gradually increases as the pH decreases, and therefore the reactivity also increases as the pH decreases. However, the pH value also affects the stability of ferrate(VI); ferrate(VI) is unstable in an acidic environment and is prone to decomposition. The higher the pH value, the more stable ferrate(VI) is. At pH 9~10, FeO42− is highly stable and this is the best condition for removing TCH. Similarly, when chemical oxidation is used to degrade pollutants, active species make a significant contribution to the removal of pollutants. It has been found that when ferrate(VI) degrades pollutants, intermediate iron species [Fe(IV) and Fe(V)] are involved in the reaction [64,65,66]. The activity of intermediate iron species is several orders of magnitude higher than that of ferrate(VI), which facilitates the degradation and removal of organic pollutants.

4.1.2. Phenolic Compounds

Phenolic compounds are widely used in the chemical industry, but they can be harmful if they remain in the environment for an extended period. Ferrate(VI) has a good removal effect on phenolic compounds and can effectively degrade phenolic compounds in a short time [67]. When using ferrate(VI) to degrade phenolic pollutants, the dosage of ferrate(VI) is an important influencing factor. Zheng et al. [20] found that a molar ratio of ferrate(VI) to 4-Tert-Butylphenol (4-tBP) greater than 10 can achieve complete degradation of 4-tBP within 5 min. Similarly, Li et al. [68] found that when the content of ferrate(VI) increased, the removal rate of Bisphenol A (BPA) also increased, and BPA could be completely degraded within 5 min under the optimal molar ratio of ferrate(VI):BPA = 5:1. Similarly, Widhiastuti et al. [69] found that when the molar ratio of ferrate(VI):BPA in a composite reverse osmosis concentrate is 50, the degradation rate of BPA exceeds 90%. In addition to the dosage, pH is also an important factor that influences the effectiveness of ferrate(VI) in removing phenolic compounds. Delaude et al. [43] found that when ferrate(VI) alone is used to remove phenol from water, the effect is not significant. However, when sodium hypochlorite reagent is added to the reaction solution, the removal rate of phenol can be significantly improved. This is because the sodium hypochlorite solution is weakly alkaline, the ferrate(VI) is more stable in the alkaline environment, and the rate of self-decomposition is inhibited, allowing most of the ferrate(VI) to be used to oxidize phenol. Similarly, Sun et al. [70] once again proved that ferrate(VI) is the most effective method for the removal of BPA under alkaline conditions; in addition, the study also found that Fe(OH)3 is also conducive to the removal of BPA. Fe(OH)3 is an effective coagulant, thus demonstrating that ferrate(VI) is effective at removing phenolic compounds.
Most phenolic compounds contain benzene rings and hydroxyl groups, and ferrate(VI) tends to react with these substances, so rapid degradation of phenolic compounds can be achieved. Ferrate(VI) can gradually break down phenolic compounds into low-molecular compounds, reducing the antibacterial properties of the decomposition products. As a result, the toxicity of phenolic compounds after oxidation of ferrate(VI) can be reduced [71].

4.1.3. Algae

Reservoir eutrophication and algal bloom outbreaks in drinking water sources have become significant threats to the current water ecological environment [72,73]. The overgrowth of algae can easily cause a series of problems such as foul odor and deterioration in water quality. At the same time, algae itself is also a precursor of DBPs, so removing algae is of great benefit for the control of DBPs.
Numerous studies have shown that ferrate(VI) is more effective at removing algae than other methods. For example, Ma et al. [74] compared the effectiveness of potassium ferrate(VI) (K2FeO4), polyferric chloride (PFC), and potassium permanganate (KMnO4) in removing algae in water. The results showed that the removal effect of K2FeO4 on algae was better than that of PFC and KMnO4, and the removal rate was as high as 92.4%. Similarly, Emília et al. [73] used ferrate(VI) to remove chlorella, and when 8.34 micromole of ferrate(VI) was added, the removal rate of chlorella was 89%. The mechanism by which ferrate(VI) removes algae is mainly through oxidation, which destroys the surface structure of algae cells, this causes the extracellular sheath to crack and the intracellular substances to discharge, which leads to the death of the algae [75]. Additionally, the reduction product Fe(OH)3 of ferrate(VI) has a flocculation effect, which can further adsorb and remove algal cells. Therefore, based on the advantages of ferrate(VI) algae removal, ferrate(VI) pre-oxidation is usually used to remove algae [76]. He et al. [77] found that after the oxidation treatment with ferrate(VI), not only was the structure of the algae cells destroyed, but the particle size and zeta potential of the algae also increased by 20% and 55%, respectively, which promotes the aggregation of algae and improves algae removal efficiency. Increasing the dosage of ferrate(VI) or the reaction time can also improve the algae removal rate. Dong et al. [78] used ferrate(VI) to remove cyanobacteria, and the dosage of ferrate(VI) was increased or the oxidation time was prolonged; consequently, the removal of algal cells and algal organic matter (AOM) by ferrate(VI) increased. Ferrate(VI) can effectively inhibit algae growth, enhance the flocculation effect of water bodies, and increase the purification capacity of a water body. At the same time, ferrate(VI) also reduces the concentration of DBPs produced during subsequent chlorination. An amount of 32% of TCM, 33% BDCM, 43% TCP, and 82% TCNM were controlled by Fe(VI) (0.03 mM) pre-oxidation, respectively [77].

4.1.4. Reduction Products of Ferrate(VI)

The reduction product Fe(III) or iron hydroxide of ferrate(VI) has a strong coagulating effect that can adsorb particles in water and reduce turbidity. Therefore, the reduced product of ferrate(VI) can also be used as a coagulant. It has the following advantages when used as a coagulant: (1) It produces low amounts of sludge and is low cost. In removing turbidity, potassium ferrate(VI) produces much less sludge than other coagulants (FeSO4-7H2O), iron nitrate [Fe(NO3)3], iron chloride (FeCl3), and aluminum sulfate [Al2(SO4)3]. (2) It can rapidly reduce colloidal stability. The ferrate(VI) can destabilize the colloid within 1 min, while it takes 30 min to achieve the same effect using iron and ferrous salts [79].
Based on the strong oxidizing and reducing capabilities of ferrate(VI), it can be used in the pre-oxidation, conventional, and deep treatment stages of water treatment processes. In addition to removing trace pollutants from water, it can also save on the use of flocculants and reduce costs. Ferrate(VI) is often used to remove sewage that is rich in heavy metals. Tiwari et al. [80] found that, at a lower dosage, the use of ferrate(VI) was able to reduce many metallic elements as well as toxic heavy metals in treated water. Similarly, Lan et al. [81] added ferrate(VI) to wastewater containing arsenic and antimony, and the removal of antimony increased with a longer contact time. This is because the addition of ferrate(VI) decomposes itself over time, and the decomposition product, Fe(III), can effectively adsorb and remove arsenic and antimony from the water. At the same time, pH is also one of the factors affecting the removal of heavy metals. Pruce et al. [82] found that when the pH rises from 5 to 10, the removal rate of Cd (II) is increased; when the pH is 10, Cd (II) is completely adsorbed and removed. Similarly, M.R et al. [83] found that at high pH values, heavy metals condense, promoting the adsorption of heavy metals by ferrate(VI). This is because, under alkaline conditions, Fe(III), a product of ferrate(VI) decomposition, can combine with OH ions in solution to form Fe(OH)3 with coagulability. The amount of ferrate(VI) dosing is proportional to the removal rate of heavy metals [84]. Prucek et al. [85] increased the ratio of ferrate(VI) to antimony to 10:1, and after some time, the removal rate of antimony reached 80%. In addition to heavy metals, ferrate(VI) can remove radionuclides from water as a coagulant [62,86,87]. Petrov et al. [88] removed α and β radionuclides from wastewater using potassium ferrate(VI). There are also indications that when used as a scavenger, ferrate(VI) can remove almost all radionuclides except cesium [31]. With its superior performance, ferrate(VI) can remove colloidal particles and adsorb heavy metal ions in water without causing secondary pollution, making it a green and efficient water treatment agent.
Table 2 summarizes the oxidative removal conditions and removal rates of various organic pollutants by ferrate(VI).

4.2. Ferrate(VI) as a Disinfectant

4.2.1. Water Disinfection

Disinfection in the water treatment system is essential for removing bacteria and viruses. During the treatment process, disinfectants are added to inhibit the growth of microorganisms. Ferrate(VI) can be used as a strong oxidant and disinfectant for water sterilization and disinfection. Compared with traditional chlorine disinfection, ferrate(VI) effectively removes microorganisms, and reduces the precursors of disinfection by-products. The schematic diagram in Figure 4 [100] illustrates the disinfection of microorganisms such as bacteria by ferrate(VI). Furthermore, ferrate(VI) kills microorganisms by oxidizing their nucleic acids and destroying their protein structures: on the one hand, ferrate(VI) kills microorganisms by oxidizing the nucleic acids of bacteria and viruses and destroying their protein structures; on the other hand, the iron hydroxide colloids obtained by ferrate(VI) reduction can adsorb bacteria and viruses and cause microorganisms to aggregate and precipitate.
Ferrate(VI) has advantages over other disinfection when used as a disinfectant, and the process is safer. Table 3 summarizes the advantages and disadvantages of ferrate(VI) and other disinfectants.
When ferrate(VI) is used to disinfect wastewater, not only can it remove organic matter from the wastewater, but it can also sterilize it. Ferrate(VI) mainly kills microorganisms in two ways: 1. By using strong oxidation, it oxidizes the cell wall, DNA, and other important microbial structures, thereby inactivating them; 2. The ferrate(VI) reaction product is ferric iron, which has a strong coagulating effect under alkaline conditions, and can remove colloidal substances such as microorganisms. Talaiekhozani et al. [101] conducted research using ferrate(VI) as a disinfectant and showed that a ferrate(VI) concentration controlled at 0–50 ppm can kill all microorganisms in water. Ferrate(VI) can destroy the surface structure of bacteria with its strong oxidation, and can rapidly inactivate Escherichia coli, total coliform, Bacillus subtilis, etc. [31]. When using ferrate(VI) for disinfection, pH is an important influencing factor. When the pH was 8.2, 6 mg/L of ferrate(VI) eliminated 99% of E. coli from the water. However, when the pH decreased to 7, the disinfection effect of ferrate(VI) increased, eliminating 99.9% of E. coli flora and 97% of bacteria in the water [31]. This is because ferrate(VI) becomes more active at lower pH levels. This shows that ferrate(VI) is highly effective at removing microorganisms such as bacteria from the water. Li et al. [102] compared ferrate(VI)-treated water with chlorinated treated water to study the bacterial communities in each. In the chlorinated samples, the number of viable bacteria increased with the increase in chlorine dosage, while in the ferrate(VI)-treated samples, the trend decreased with the increasing ferrate(VI) dose. This suggests that ferrate(VI) is more effective at removing microorganisms such as bacteria more significantly than chlorine, and can reduce the formation of DBPs.

4.2.2. Control Application of Disinfection By-Products

Chlorine is a commonly used disinfectant in wastewater treatment plants, due to its low cost and high disinfection efficiency. However, during the disinfection process, it reacts with natural organic matter (NOM) in water to form volatile trihalomethanes (THMs) and refractory halogenacetic acids (HAAs) [103,104], as well as haloacetonitriles (HANs) and haloaldehydes (HKs). It has been shown that the use of ferrate(VI) (0.03 mM) in pre-oxidation can reduce the amounts of TCM and BDCM by 32% and 33%, respectively [77]. The mechanism by which ferrate(VI) can inhibit the formation of chlorine disinfection by-products is through the transfer of amines, alcohols, and amino acids from NOM to carbonyl compounds, which are then further oxidized to acids that are highly resistant to chlorine, thereby hindering the production of THMs.
When ferrate(VI) is used as a disinfectant, it can effectively control the formation of disinfection by-products. Ferrate(VI) removes total organic carbon (TOC) better than chlorine and produces relatively small amounts of THMs. Li et al. [105] studied the effect of ferrate(VI) on the formation of DBPs in sewage plants. The contents of trichloromethane (TCM), dichlorobromomethane (DCBM), dibromochloromethane (DBCM), and tribromomethane (TBM) in water samples decreased with the addition of ferrate(VI) concentration, and ferrate(VI) pre-treatment could effectively reduce the amount of THMs generated during chlorination. The control mechanism of ferrate(VI) on DBPs is mainly the oxidation of its precursors to achieve removal. The control mechanism of ferrate(VI) on disinfection by-products is summarized in Table 4. The effect of ferrate(VI) on the control of DBP production is mainly influenced by the amount and pH of the ferrate(VI) dosing. Similarly, Li et al. [105] found that the levels of TCM, DCBM, DBCM, and TBM decreased with increasing ferrate(VI) concentration when ferrate(VI) was added to the water samples. The pH of the solution also affected the production of DBPs. Dong et al. [78] found in their experiments that the yield of DBPs in Chlorella solution gradually decreased when the pH was increased from 7 to 9. However, Ding et al. [45] found that the best removal of dichloroacetonitrile (DCAN) by ferrate(VI) was achieved at pH 6.5. The control of DBPs by ferrate(VI) varies under different pH conditions. Under acidic conditions, the pre-oxidation of ferrate(VI) was effective in reducing the generation of THMs; on the contrary, for HANs, a lesser amount was generated under alkaline conditions. This is related to the fact that the precursors of HANs, tryptophan or kynurenine, are more active under acidic conditions [106]. Therefore, when using ferrate(VI) to control DBPs, the pH should be determined according to the properties of the DBPs.

5. Hyphenated Techniques of Ferrate(VI)

Although ferrate(VI) is an effective tool for removing refractory pollutants in water, its application still has some challenges. Ferrate(VI) has strong selectivity for removal, and different pollutants have different removal rates [111]. Ferrate(VI) reacts quickly with organic pollutants containing an electron-rich organic fraction first, and reacts slowly with other organic substances (e.g., ibuprofen and meprobamate) [112,113]. Ferrate(VI) is added to help the solution decompose rapidly. Therefore, to overcome these problems, the hyphenated technique of ferrate(VI) has been developed. Shao et al. [27] observed that when compared to ferrate(VI) treatment alone, CaSO3 can promote the conversion of ferrate(VI) to phases Fe(IV) and Fe(V) and increase the removal rate of organic pollutants by 6.1 times. At present, the hyphenated techniques of ferrate(VI) include ferrate(VI) and photocatalysis, which uses light energy to convert ferrate(VI) into Fe(IV) and Fe(V) with a stronger oxidation capacity [114]. The removal rate of pollutants by the combination of ferrate(VI) and peroxymonosulfate (PMS) is significantly higher than that of ferrate(VI) or PMS alone. Ferrate(VI) is combined with PAA, and PAA will also promote the decomposition of ferrate(VI) into Fe(IV) and Fe(V) with strong oxidizing ability, thereby improving the removal rate of organic pollutants. Table 5 summarizes the hyphenated techniques for ferrate(VI).

5.1. Ferrate(VI)–Photocatalysis

To further overcome the selectivity of ferrate(VI) for the scavengers, the removal of organics can be enhanced by ferrate(VI) activation or catalysis [115]. It was found that ferrate(VI) can act as an electron acceptor in the UV/TiO2 system, and the decomposition of ferrate(VI) to Fe(IV), Fe(V), and Fe(III), shown in Equations (11)–(13) [116]. In addition, the researchers also found that under UV irradiation, it is accompanied by the production of superoxide free radicals, as shown in Figure 5 [116]. The combined system can effectively degrade organic matter in water, and photocatalysis uses sunlight as the light source, which can reduce costs and will not produce secondary pollutants. The principle of ferrate(VI) and photocatalysis is mainly based on the catalyst TiO2 under ultraviolet irradiation, which causes electron transition on the valence band to form electron-hole pairs, while electrons are negatively charged, have reduction, holes and undergo oxidation, ferrate(VI) can be used as an electron trap, and can effectively prevent electrons and holes from combining, while electron-hole pairs can reduce ferrate(VI) to Fe(IV) and Fe(V) with stronger oxidation capacity. Tian et al. [114] also proved that Fe(IV) and Fe(V) are stronger than ferrate(VI), about 1000 times more, so the degradation of the target pollutant by ferrate(VI)–photocatalysis is more thorough.
H F e O 4 - + e c b - H F e O 4 2 -
H F e O 4 2 - + e c b - H F e O 4 3 -
H F e O 4 3 - + 3 H 2 O + e c b - F e ( O H ) 3 + 4 O H -
For example, Wu et al. [116] found a method that can effectively degrade 2,4-DCP in water; the activation of ferrate(VI) by ultraviolet light can significantly increase the degradation rate of 2,4-DCP by ferrate(VI), which is 6.9 times and 9.2 times higher than that of ultraviolet irradiation alone and ferrate(VI) oxidation alone, respectively. Similarly, Yuan et al. [117] studied the degradation of dimethyl phthalate (DMP) by K2FeO4 and TiO2 photocatalysis, and found that the ferrate(VI)-TiO2-UV system had the best degradation effect on DMP, which was significantly higher than the degradation rate under potassium ferrate(VI) and UV irradiation alone. Similarly, Heng et al. [91] also used ferrate(VI)/ultraviolet light to synergistically degrade p-nitrophenol in water, and the removal rate of ferrate(VI)/ultraviolet light system was significantly higher than that of ferrate(VI) alone, and the CODCr degradation rate of p-nitrophenol reached 85.71%. It can be seen from the data that ferrate(VI) and photocatalytic technology can significantly improve the removal rate of organic pollutants. In addition, ferrate(VI)–photocatalytic technology effectively removes algae and other substances in water. Xia et al. [118] used ferrate(VI)–photocatalysis to remove algal toxins from water. When the dosage of ferrate(VI) was 10 mg/L, the removal rate of algal toxin by the combination of ferrate(VI) and photocatalytic technology increased from 63% to 100%, which could quickly remove the algal toxin. Combined with several studies, it can be seen that the effect of ferrate(VI) and photocatalytic technology is significantly higher than that of ferrate(VI) alone, and has a good application prospect.

5.2. Ferrate(VI)–PAA

Peracetic acid (PAA CH3COOOH) is a volatile and easily decomposable chemical with strong oxidation properties, and the decomposition products CH3COOH, CH3OH, and O2 are harmless. In addition, PAA, like ferrate(VI), can be used as a disinfectant, and the DBPs produced are less harmful than halogenated disinfection by-products [16,119]. It is an effective method to activate ferrate(VI) by transferring single or double electrons to reducing agents (SO32− and S2O32−) [120] and solid materials (carbonaceous materials and silica gels) [121,122] to improve the removal rate of pollutants. PAA combined with ferrate(VI) is a new area of research. Kim et al. [123] used Fe(II) to activate PAA, promoting the transformation of ferrate(VI) into higher oxidizing capacity Fe(IV) and Fe(V) that can effectively remove pollutants from water.
Wang et al. [124] found that when ferrate(VI) is combined with PAA, the removal effect of ferrate(VI) on micropollutants is improved by promoting electron transfer efficiency. At pH 7.3 and 8, the addition of PAA to the ferrate(VI) system has a better effect on the removal of CBZ than ferrate(VI) alone. Similarly, Manoli et al. [125] found that when ferrate(VI) (200 μM) and PAA (100 μM) combined at pH 9, they could remove 80–90% of PPCPs in 1 min. PAA can enhance the removal rate of ferrate(VI) from organic matter by decomposing into Fe(IV) and Fe(V) through electron transfer, which greatly improves the oxidation performance of Fe(IV) and Fe(V) compared to ferrate(VI) [27], and at the same time, PAA will also decompose to produce strong oxidizing organic radicals (CH3C(O)O), making the degradation of organic matter more thorough and more rapid. Manoli [125] also demonstrated that the main cause of the degradation of organic matter in the ferrate(VI)–PAA system is Fe(V)/Fe(IV). When ferrate(VI) is used in combination with PAA, most of Fe(IV) and Fe(V) in the system are used for organic degradation, which is in contrast to using ferrate(VI) alone, as shown in Figure 6. The ferrate(VI)–PAA oxidation system has great potential for the degradation of pollutants in water, and also reduces the cost of removing pollutants, making it a promising application.

5.3. Ferrate(VI)–PMS

As ferrate(VI) is in an aqueous solution, it is extremely unstable and will decompose rapidly under acidic and neutral conditions. Therefore, a solution is needed to increase the oxidation time of ferrate(VI) in order to improve the efficiency of the reaction [126]. Shao [127] and Yang [128] et al. combined ferrate(VI) with various chemicals to improve the removal of contaminants. For example, the application of ferrate(VI)/sulfite can degrade enrofloxacin and phenol has a better degradation effect than ferrate(VI) alone. However, Wu et al. [113,114] found that the ferrate(VI)/PMS process was more effective than that of atrazine in the degradation of atrazine, and hydroxyl radicals and sulfate radicals were detected in the process of ferrate(VI)/PMS for the degradation of atrazine. It has been reported that ferrate(VI) can be used in combination with PMS, since PMS is also an oxidant, which can act as an electron acceptor or donor to improve degradation efficiency against contaminants [129]. The self-decomposition of ferrate(VI) produces Fe3+ and Fe2+, which can activate PMS and form sulfate radicals ( S O 4 - ) [130].
However, ferrate(VI)/PMS systems are affected by various factors when degrading contaminants. The pH of the initial solution has a certain effect on the degradation effect. He [131] et al. found that the removal rates of the ferrate(VI)/PMS system at pH 4, 5, 6, 7, and 8 were 74.7%, 81.9%, 68.1%, 53.2%, and 44.8%, respectively. This is because, between pH 4 and 7, the main type of ferrate(VI) is HFeO4, which is higher in activity but easily decomposes [132]. When the pH is higher than 7, FeO42− becomes the main species, although the stability is improved, the activity is reduced, the activation of PMS is weakened, and the removal rate of diclofenac(DCF) is reduced. Excess H+ will form hydrogen bonds with the main substance of PMS (HSO5) in an acidic environment, inhibiting the synergistic effect of ferrate(VI) and PMS and reducing the removal rate of DCF. In addition, the dose of ferrate(VI) and PMS also affects the removal rate. When Feng [133] et al. used ferrate(VI) and PMS to degrade flumequine (FLU), the solubility of ferrate(VI) remained unchanged, and the molar ratio of PMS:FLU was increased from 10 to 200, and the removal rate of FLU was also increased from 21% to 61%. Since both substances are oxidants, have an oxidizing effect, and have a synergistic effect, the removal rate will be increased by increasing the dose. This is because reactive sulfur species (RSS) (e.g., S O 4 - ) are generated during the reaction. Reactive oxygen radicals (ROS) (OH·) and high-valent iron (FeIV and FeV), iron oxide/iron hydroxide, a reduction product of ferrate(VI), also accelerate the formation of S O 4 - in PMS [134]. Núria et al. [135] activated PMS in sunlight to enhance the disinfection and decontamination capacity of ferrate(VI) under neutral conditions. When using ferrate(VI)/PMS/solar to degrade sulfamethoxazole (SMX), within one hour, SMX degradation was nearly 80%, and the effect was much higher than the removal rate of ferrate(VI) alone or ferrate(VI)/PMS. The combination of ferrate(VI) and suitable agents can not only enhance the stability of ferrate(VI), but also improve the removal rate of ferrate(VI) to target pollutants, which has broad research and development prospects.
Table 5. Removal of pollutants by ferrate(VI)-combined technology.
Table 5. Removal of pollutants by ferrate(VI)-combined technology.
Combined TechnologypHPollutantsRemoval Rate/%Active SpeciesReferences
Ferrate(VI)–UV62,4-DCP75.8Fe(IV), Fe(V) and·O2[116]
6BPA79.2Fe(IV), Fe(V) and·O2
2Formaldehyde100Fe(IV), Fe(V) and·O2[136]
9DMP40Fe(IV), Fe(V) and·O2[117]
711P-nitrophenol62.6576.31Fe(IV), Fe(V) and·O2[91]
Ferrate(VI)–PAA8CBZ100Fe(IV), Fe(V) and CH3C(O)O·[124]
6–9CBZ100Fe(IV), Fe(V) and CH3C(O)O·[125]
Ferrate(VI)–aluminum saltCODMnUV2544060[29]
DOCUV254Protein35.933.771.8Fe(IV), Fe(V)[137]
Microcystic aeruginosa59.3Fe(IV), Fe(V)
Thallium92[138]
Ferrate(VI)–PMS6Atrazine100Fe(IV), Fe(V),·OH and SO4·[113]
5SMX80[139]

6. Conclusions

Ferrate(VI) is a highly efficient water treatment agent that combines the functions of oxidation, disinfection, and adsorption flocculation. The application prospect of ferrate(VI) is also extensive. According to this paper, the following conclusions are drawn: (1) Ferrate(VI) is very widely used in the field of water treatment. Due to its strong oxidizing properties, it can remove most organic pollutants and algae from the water. As a disinfectant, it can quickly inactivate bacteria and viruses, and has a control effect on the disinfection by-products produced during chlorine disinfection, which is a non-toxic and harmless disinfectant. As a coagulant, it can adsorb most of the particles in water and has a certain adsorption effect on heavy metals, reducing water turbidity. (2) The disadvantage of ferrate(VI) is that it is unstable, so how to obtain high-purity ferrate(VI) in the process of preparation is one of the following research hotspots. Among the existing methods for the preparation of ferrate(VI), electrolysis is the most promising green method; at the same time, the influencing factors in the preparation process are solution temperature, pH, current density, alkali concentration, etc. The influence of these factors on the results from continuous experiments must be controlled. Ferrate(VI) is also very harsh on preservation conditions, which is unfavorable for its storage and transportation. Developing the in situ preparation process and improving its preservation requirements will, to some extent, promote its application. (3) Ferrate(VI) has strong oxidizing properties and is effective at pollutant removal. However, ferrate(VI) is highly selective of the removables and the removal rate varies for different pollutants. Combining ferrate(VI) with other technologies can improve the removal rate of target pollutants. The hyphenated techniques of ferrate(VI) coupling mainly include ferrate(VI)–PMS, ferrate(VI)–PAA, ferrate(VI)–photocatalytic coupling, etc. In recent years, ferrate(VI)-coupling technology has also been the hotspot and application direction of research on ferrate(VI). Therefore, the combination of ferrate(VI) and other water treatment agents is a new research direction to promote the application of ferrate(VI).

Author Contributions

Writing—original draft preparation, J.Y. and S.H.; writing—review and editing, J.Y., K.Z., Q.Z. and S.; visualization, J.Y., C.W., and S.Y.; and supervision, W.P. and Y.Z. project administration, W.P. and Y.Z.; funding acquisition, Y.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science Foundation of Shanghai (Grant No. 20ZR1438200) and the National Key Research and Development Project of China (Grant No. 2021YFC3200805).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bhavya, G.; Belorkar, S.A.; Mythili, R.; Geetha, N.; Shetty, H.S.; Udikeri, S.S.; Jogaiah, S. Remediation of Emerging Environmental Pollutants: A Review Based on Advances in the Uses of Eco-Friendly Biofabricated Nanomaterials. Chemosphere 2021, 275, 129975. [Google Scholar] [CrossRef] [PubMed]
  2. Satapute, P.; Paidi, M.K.; Kurjogi, M.; Jogaiah, S. Physiological adaptation and spectral annotation of Arsenic and Cadmium heavy metal-resistant and susceptible strain Pseudomonas taiwanensis. Environ. Pollut. 2019, 251, 555–563. [Google Scholar] [CrossRef]
  3. Gago-Ferrero, P.; Gros, M.; Ahrens, L.; Wiberg, K. Impact of on-site, small and large scale wastewater treatment facilities on levels and fate of pharmaceuticals, personal care products, artificial sweeteners, pesticides, and perfluoroalkyl substances in recipient waters. Sci. Total. Environ. 2017, 601–602, 1289–1297. [Google Scholar] [CrossRef] [PubMed]
  4. Klatte, S.; Schaefer, H.C.; Hempel, M. Pharmaceuticals in the Environment—A Short Review on Options to Minimize the Exposure of Humans, Animals and Ecosystems. Sustain. Chem. Pharm. 2016, 5, 61–66. [Google Scholar] [CrossRef]
  5. Yang, Z.; Su, R.; Luo, S.; Spinney, R.; Cai, M.; Xiao, R.; Wei, Z. Comparison of the reactivity of ibuprofen with sulfate and hydroxyl radicals: An experimental and theoretical study. Sci. Total. Environ. 2017, 590-591, 751–760. [Google Scholar] [CrossRef]
  6. Wang, Y.; Gao, Z.; Shang, Y.; Qi, Z.; Zhao, W.; Peng, Y. Chemical structure-based predictive model for the oxidation of trace organic contaminants by sulfate radical. Water Res. 2017, 116, 106–115. [Google Scholar]
  7. Wang, Y.; Gao, Z.; Shang, Y.; Qi, Z.; Zhao, W.; Peng, Y. Proportional Modulation of Zinc-Based Mof/Carbon Nanotube Hybrids for Simultaneous Removal of Phosphate and Emerging Organic Contaminants with High Efficiency. Chem. Eng. J. 2020, 417, 128063. [Google Scholar] [CrossRef]
  8. Liu, J.; Li, B.; Wang, G.; Qin, L.; Ma, X.; Hu, Y.; Cheng, H. Facile synthesis of flower-like CoFe2O4 particles for efficient sorption of aromatic organoarsenicals from aqueous solution. J. Colloid Interface Sci. 2020, 568, 63–75. [Google Scholar] [CrossRef]
  9. Ghattas, A.K.; Fischer, F.; Wick, A.; Ternes, T.A. Anaerobic Biodegradation of (Emerging) Organic Contaminants in the Aquatic Environment. Water Res. 2017, 116, 268–295. [Google Scholar] [CrossRef]
  10. Annamalai, J.; Vasudevan, N. Enhanced biodegradation of an endocrine disrupting micro-pollutant: Di (2-ethylhexyl) phthalate using biogenic self-assembled monolayer of silver nanoparticles. Sci. Total. Environ. 2020, 719, 137115. [Google Scholar] [CrossRef] [PubMed]
  11. Xie, X.; Zhao, W.; Hu, Y.; Xu, X.; Cheng, H. Permanganate oxidation and ferric ion precipitation (KMnO4-Fe(III)) process for treating phenylarsenic compounds. Chem. Eng. J. 2018, 357, 600–610. [Google Scholar] [CrossRef]
  12. Kennedy, A.M.; Reinert, A.M.; Knappe, D.R.; Ferrer, I.; Summers, R. Full- and pilot-scale GAC adsorption of organic micropollutants. Water Res. 2015, 68, 238–248. [Google Scholar] [CrossRef]
  13. Aris, A.Z.; Hir, Z.A.M.; Razak, M.R. Metal-organic frameworks (MOFs) for the adsorptive removal of selected endocrine disrupting compounds (EDCs) from aqueous solution: A review. Appl. Mater. Today 2020, 21, 100796. [Google Scholar] [CrossRef]
  14. Mir-Tutusaus, J.A.; Baccar, R.; Caminal, G.; Sarrà, M. Can White-Rot Fungi Be a Real Wastewater Treatment Alternative for Organic Micropollutants Removal? A Review. Water Res. 2018, 138, 137–151. [Google Scholar] [CrossRef]
  15. Usman, M.; Ho, Y.-S. A bibliometric study of the Fenton oxidation for soil and water remediation. J. Environ. Manag. 2020, 270, 110886. [Google Scholar] [CrossRef]
  16. Shi, C.; Wang, Y.; Zhang, K.; Lichtfouse, E.; Li, C.; Zhang, Y. Fe-biochar as a safe and efficient catalyst to activate peracetic acid for the removal of the acid orange dye from water. Chemosphere 2022, 307, 135686. [Google Scholar] [CrossRef] [PubMed]
  17. Dar, A.A.; Wang, X.; Wang, S.; Ge, J.; Shad, A.; Ai, F.; Wang, Z. Ozonation of Pentabromophenol in Aqueous Basic Medium: Kinetics, Pathways, Mechanism, Dimerization and Toxicity Assessment. Chemosphere 2019, 220, 546–555. [Google Scholar] [CrossRef] [PubMed]
  18. Dar, A.A.; Pan, B.; Qin, J.; Zhu, Q.; Lichtfouse, E.; Usman, M.; Wang, C. Sustainable Ferrate Oxidation: Reaction Chemistry, Mechanisms and Removal of Pollutants in Wastewater. Environ. Pollut. 2021, 290, 117957. [Google Scholar] [CrossRef]
  19. Yang, T.; Wang, L.; Liu, Y.-L.; Zhang, W.; Cheng, H.-J.; Liu, M.-C.; Ma, J. Ferrate oxidation of bisphenol F and removal of oxidation products with ferrate resulted particles. Chem. Eng. J. 2019, 383, 123167. [Google Scholar] [CrossRef]
  20. Zheng, Q.; Wu, N.; Qu, R.; Albasher, G.; Cao, W.; Li, B.; Alsultan, N.; Wang, Z. Kinetics and reaction pathways for the transformation of 4-tert-butylphenol by ferrate(VI). J. Hazard. Mater. 2020, 401, 123405. [Google Scholar] [CrossRef]
  21. Liu, J.; Mulenos, M.R.; Hockaday, W.C.; Sayes, C.M.; Sharma, V.K. Ferrate(VI) pretreatment of water containing natural organic matter, bromide, and iodide: A potential strategy to control soluble lead release from PbO2(s). Chemosphere 2020, 263, 128035. [Google Scholar] [CrossRef] [PubMed]
  22. Jiang, J. Research progress in the use of ferrate(VI) for the environmental remediation. J. Hazard. Mater. 2007, 146, 617–623. [Google Scholar] [CrossRef] [PubMed]
  23. Alsheyab, M.; Jiang, J.-Q.; Stanford, C. Electrochemical generation of ferrate (VI): Determination of optimum conditions. Desalination 2010, 254, 175–178. [Google Scholar] [CrossRef]
  24. Wood, R.H. The Heat, Free Energy and Entropy of the Ferrate(VI) Ion. J. Am. Chem. Soc. 1958, 80, 2038–2041. [Google Scholar] [CrossRef]
  25. You, K.; Fu, J.X.; Tang, Y.L.; Zhao, Y.H.; Lan, Y.Y. Study on Stability of Ferrate (VI) under the Effect of the Buffering Agent and the Strong Oxidizer. Adv. Mater. Res. 2010, 113–116, 2319–2322. [Google Scholar] [CrossRef]
  26. Han, Q.; Wang, H.; Dong, W.; Liu, T.; Yin, Y. WITHDRAWN: Suppression of bromate formation in ozonation process by using ferrate(VI): Batch study. Chem. Eng. J. 2014, 236, 110–120. [Google Scholar] [CrossRef]
  27. Shao, B.; Dong, H.; Sun, B.; Guan, X. Role of Ferrate(IV) and Ferrate(V) in Activating Ferrate(VI) by Calcium Sulfite for Enhanced Oxidation of Organic Contaminants. Environ. Sci. Technol. 2018, 53, 894–902. [Google Scholar] [CrossRef]
  28. Sailo, L.; Tiwari, D.; Lee, S.-M. Degradation of some micro-pollutants from aqueous solutions using ferrate (VI): Physico-chemical studies. Sep. Sci. Technol. 2017, 52, 2756–2766. [Google Scholar] [CrossRef]
  29. Zhou, J.; Zhao, Z.; Liu, J.; Peng, W.; Peng, X.; Han, Y.; Xiao, P. Removal of Microcystis aeruginosa and control of algal organic matters by potassium ferrate(VI) pre-oxidation enhanced Fe(II) coagulation. Korean J. Chem. Eng. 2019, 36, 1587–1594. [Google Scholar] [CrossRef]
  30. Huang, X.; Deng, Y.; Liu, S.; Song, Y.; Li, N.; Zhou, J. Formation of Bromate During Ferrate(Vi) Oxidation of Bromide in Water. Chemosphere 2016, 155, 528–533. [Google Scholar] [CrossRef]
  31. Talaiekhozani, A.; Talaei, M.R.; Rezania, S. An Overview on Production and Application of Ferrate (Vi) for Chemical Oxidation, Coagulation and Disinfection of Water and Wastewater. J. Environ. Chem. Eng. 2017, 5, 1828–1842. [Google Scholar] [CrossRef]
  32. Tong, R.; Zhang, P.; Yang, Y.; Zhang, R.; Sun, X.; Ma, X.; Sharma, V.K. On line continuous chemical synthesis of ferrate(VI): Enhanced yield and removal of pollutants. J. Environ. Chem. Eng. 2021, 9, 106512. [Google Scholar] [CrossRef]
  33. Munyengabe, A.; Zvinowanda, C. Production, Characterization and Application of Ferrate(VI) in Water and Wastewater Treatments. Braz. J. Anal. Chem. 2019, 6, 40–57. [Google Scholar] [CrossRef]
  34. Martinez-Tamayo, E.; Beltran-Porter, A.; Beltran-Porter, D. Iron Compounds in High Oxidation States: Ii. Reaction between Na2o2 and Feso4. Thermochim. Acta 1986, 97, 243–255. [Google Scholar] [CrossRef]
  35. Ninane, L.; Kanari, N.; Criado, C.; Jeannot, C.; Evrard, O.; Neveux, N. New Processes for Alkali Ferrate Synthesis. In Ferrates; American Chemical Society: Washington, DC, USA, 2008; pp. 102–111. [Google Scholar]
  36. Kanari, N.; Ostrosi, E.; Ninane, L.; Neveux, N.; Evrard, O. Synthesizing alkali ferrates using a waste as a raw material. Jom 2005, 57, 39–42. [Google Scholar] [CrossRef]
  37. Sharma, V.K. Potassium ferrate(VI): An environmentally friendly oxidant. Adv. Environ. Res. 2002, 6, 143–156. [Google Scholar] [CrossRef]
  38. Song, Y.; Men, B.; Wang, D.; Ma, J. On-line batch production of ferrate with an chemical method and its potential application for greywater recycling with Al(III) salt. J. Environ. Sci. 2017, 52, 1–7. [Google Scholar] [CrossRef]
  39. Yang, W.; Vollertsen, J.; Hvitved-Jacobsen, T. Anoxic sulfide oxidation in wastewater of sewer networks. Water Sci. Technol. 2005, 52, 191–199. [Google Scholar] [CrossRef]
  40. Tiwari, D.; Kim, H.U.; Choi, B.J.; Lee, S.M.; Kwon, O.H.; Choi, K.M.; Yang, J.K. Ferrate(Vi): A Green Chemical for the Oxidation of Cyanide in Aqueous/Waste Solutions. J. Environ. Sci. Health Part A-Toxic/Hazard. Subst. Environ. Eng. 2007, 42, 803–810. [Google Scholar] [CrossRef]
  41. Lee, J.; Von Gunten, U.; Kim, J.H. Persulfate-Based Advanced Oxidation: Critical Assessment of Opportunities and Roadblocks. Environ. Sci. Technol. 2020, 54, 3064–3081. [Google Scholar] [CrossRef]
  42. Yu, X.; Licht, S. Advances in electrochemical Fe(VI) synthesis and analysis. J. Appl. Electrochem. 2008, 38, 731–742. [Google Scholar] [CrossRef]
  43. Delaude, L.; Laszlo, P. A Novel Oxidizing Reagent Based on Potassium Ferrate(Vi)(1). J. Org. Chem. 1996, 61, 6360–6370. [Google Scholar] [CrossRef] [PubMed]
  44. Zhang, H.; Zheng, L.; Li, Z.; Pi, K.; Deng, Y. One-step Ferrate(VI) treatment as a core process for alternative drinking water treatment. Chemosphere 2019, 242, 125134. [Google Scholar] [CrossRef] [PubMed]
  45. Ding, Z.; Yang, C.; Wu, Q. The Electrochemical Generation of Ferrate at Porous Magnetite Electrode. Electrochim. Acta 2004, 49, 3155–3159. [Google Scholar] [CrossRef]
  46. Diaz, M.; Doederer, K.; Keller, J.; Cataldo, M.; Donose, B.-C.; Ali, Y.; Ledezma, P. Towards in situ electro-generation of ferrate for drinking water treatment: A comparison of three low-cost sacrificial iron electrodes. J. Electroanal. Chem. 2021, 880, 114897. [Google Scholar] [CrossRef]
  47. He, T.; Zhou, B.; Chen, H.; Yuan, R. Degradation of organic chemicals in aqueous system through ferrate-based processes: A review. J. Environ. Chem. Eng. 2022, 10, 108706. [Google Scholar] [CrossRef]
  48. Barışçı, S. Electrochemical Ferrate(VI) Synthesis with Cast Iron Shavings and Its Potential Application for Household Graywater Recycling. J. Environ. Eng. 2018, 144, 04018073. [Google Scholar] [CrossRef]
  49. Hu, Y.; Jensen, J.O.; Pan, C.; Cleemann, L.N.; Shypunov, I.; Li, Q. Immunity of the Fe-N-C catalysts to electrolyte adsorption: Phosphate but not perchloric anions. Appl. Catal. B: Environ. 2018, 234, 357–364. [Google Scholar] [CrossRef]
  50. Cataldo-Hernández, M.A.; Govindarajan, R.; Bonakdarpour, A.; Mohseni, M.; Wilkinson, D.P. Electrosynthesis of ferrate in a batch reactor at neutral conditions for drinking water applications. Can. J. Chem. Eng. 2018, 96, 1648–1655. [Google Scholar] [CrossRef]
  51. Pi, R.; Liu, H.; Sun, X.; Zhang, R.; Zhang, J.; Sharma, V.K. Strategy of periodic reverse current electrolysis to synthesize Ferrate(VI): Enhanced yield and removal of sulfachloropyridazine. Sep. Purif. Technol. 2021, 263, 118420. [Google Scholar] [CrossRef]
  52. Goodwill, J.E.; Jiang, Y.; Reckhow, D.A.; Gikonyo, J.; Tobiason, J.E. Characterization of Particles from Ferrate Preoxidation. Environ. Sci. Technol. 2015, 49, 4955–4962. [Google Scholar] [CrossRef]
  53. Jiang, Y.; Goodwill, J.E.; Tobiason, J.E.; Reckhow, D.A. Bromide oxidation by ferrate(VI): The formation of active bromine and bromate. Water Res. 2016, 96, 188–197. [Google Scholar] [CrossRef]
  54. Yang, Y.; Ok, Y.S.; Kim, K.-H.; Kwon, E.E.; Tsang, Y.F. Occurrences and removal of pharmaceuticals and personal care products (PPCPs) in drinking water and water/sewage treatment plants: A review. Sci. Total. Environ. 2017, 596–597, 303–320. [Google Scholar] [CrossRef] [PubMed]
  55. Banjac, Z.; Ginebreda, A.; Kuzmanovic, M.; Marcé, R.; Nadal, M.; Riera, J.M.; Barceló, D. Emission factor estimation of ca. 160 emerging organic microcontaminants by inverse modeling in a Mediterranean river basin (Llobregat, NE Spain). Sci. Total. Environ. 2015, 520, 241–252. [Google Scholar] [CrossRef]
  56. Sun, S.; Jiang, J.; Pang, S.; Ma, J.; Xue, M.; Li, J.; Liu, Y.; Yuan, Y. Oxidation of theophylline by Ferrate (VI) and formation of disinfection byproducts during subsequent chlorination. Sep. Purif. Technol. 2018, 201, 283–290. [Google Scholar] [CrossRef]
  57. Mackuľak, T.; Grabic, R.; Špalková, V.; Belišová, N.; Škulcová, A.; Slavík, O.; Horký, P.; Gál, M.; Filip, J.; Híveš, J.; et al. Hospital Wastewaters Treatment: Fenton Reaction Vs. Bdde Vs. Ferrate(Vi). Environ. Sci. Pollut. Res. 2019, 26, 31812–31821. [Google Scholar] [CrossRef]
  58. Luo, C.; Feng, M.; Sharma, V.K.; Huang, C.-H. Oxidation of Pharmaceuticals by Ferrate(VI) in Hydrolyzed Urine: Effects of Major Inorganic Constituents. Environ. Sci. Technol. 2019, 53, 5272–5281. [Google Scholar] [CrossRef] [PubMed]
  59. Acosta-Rangel, A.; Sánchez-Polo, M.; Rozalen, M.; Rivera-Utrilla, J.; Polo, A.M.S.; Berber-Mendoza, M.S.; López-Ramón, M.V. Oxidation of Sulfonamides by Ferrate(Vi): Reaction Kinetics, Transformation Byproducts and Toxicity Assesment. J. Environ. Manag. 2020, 255, 109927. [Google Scholar] [CrossRef]
  60. Sun, X.; Feng, M.; Dong, S.; Qi, Y.; Sun, L.; Nesnas, N.; Sharma, V.K. Removal of Sulfachloropyridazine by Ferrate(Vi): Kinetics, Reaction Pathways, Biodegradation, and Toxicity Evaluation. Chem. Eng. J. 2019, 372, 742–751. [Google Scholar] [CrossRef]
  61. Yang, B.; Ying, G.G.; Zhao, J.L.; Liu, S.; Zhou, L.J.; Chen, F. Removal of Selected Endocrine Disrupting Chemicals (Edcs) and Pharmaceuticals and Personal Care Products (Ppcps) During Ferrate(Vi) Treatment of Secondary Wastewater Effluents. Water Res. 2012, 46, 2194–2204. [Google Scholar] [CrossRef] [PubMed]
  62. Kovalakova, P.; Cizmas, L.; Feng, M.; McDonald, T.J.; Marsalek, B.; Sharma, V.K. Oxidation of Antibiotics by Ferrate(Vi) in Water: Evaluation of Their Removal Efficiency and Toxicity Changes. Chemosphere 2021, 277, 130365. [Google Scholar] [CrossRef] [PubMed]
  63. Ma, Y.; Gao, N.; Li, C. Degradation and Pathway of Tetracycline Hydrochloride in Aqueous Solution by Potassium Ferrate. Environ. Eng. Sci. 2012, 29, 357–362. [Google Scholar] [CrossRef] [PubMed]
  64. Dong, H.; Li, Y.; Wang, S.; Liu, W.; Zhou, G.; Xie, Y.; Guan, X. Both Fe(Iv) and Radicals Are Active Oxidants in the Fe(Ii)/Peroxydisulfate Process. Environ. Sci. Technol. Lett. 2020, 7, 219–224. [Google Scholar] [CrossRef]
  65. Li, X.; Liu, X.; Lin, C.; Zhou, Z.; He, M.; Ouyang, W. Catalytic Oxidation of Contaminants by Fe-0 Activated Peroxymonosulfate Process: Fe(Iv) Involvement, Degradation Intermediates and Toxicity Evaluation. Chem. Eng. J. 2020, 382. [Google Scholar] [CrossRef]
  66. Wang, Z.; Jiang, J.; Pang, S.; Zhou, Y.; Guan, C.; Gao, Y.; Li, J.; Yang, Y.; Qiu, W.; Jiang, C. Is Sulfate Radical Really Generated from Peroxydisulfate Activated by Iron(Ii) for Environmental Decontamination? Environ. Sci. Technol. 2018, 52, 11276–11284. [Google Scholar] [CrossRef] [PubMed]
  67. Deng, J.; Wu, H.; Wang, S.; Liu, Y.; Wang, H. Removal of Sulfapyridine by Ferrate(Vi): Efficiency, Influencing Factors and Oxidation Pathway. Environ. Technol. 2019, 40, 1585–1591. [Google Scholar] [CrossRef]
  68. Li, C.; Li, X.; Graham, N.; Gao, N. The aqueous degradation of bisphenol A and steroid estrogens by ferrate. Water Res. 2008, 42, 109–120. [Google Scholar] [CrossRef]
  69. Widhiastuti, F.; Fan, L.; Paz-Ferreiro, J.; Chiang, K. Oxidative Degradation of Bisphenol a in Municipal Wastewater Reverse Osmosis Concentrate (Roc) Using Ferrate(Vi)/Hydrogen Peroxide. Process Saf. Environ. Prot. 2022, 163, 58–67. [Google Scholar] [CrossRef]
  70. Sun, X.; Zhang, Q.; Liang, H.; Ying, L.; Xiangxu, M.; Sharma, V.K. Ferrate(Vi) as a Greener Oxidant: Electrochemical Generation and Treatment of Phenol. J. Hazard. Mater. 2016, 319, 130–136. [Google Scholar] [CrossRef]
  71. Yang, T.; Wang, L.; Liu, Y.; Huang, Z.; He, H.; Wang, X.; Jiang, J.; Gao, D.; Ma, J. Comparative Study on Ferrate Oxidation of Bps and Bpaf: Kinetics, Reaction Mechanism, and the Improvement on Their Biodegradability. Water Res. 2019, 148, 115–125. [Google Scholar] [CrossRef]
  72. Bullerjahn, G.S.; McKay, R.M.; Davis, T.W.; Baker, D.B.; Boyer, G.L.; D’Anglada, L.V.; Doucette, G.J.; Ho, J.C.; Irwin, E.G.; Kling, C.L.; et al. Global Solutions to Regional Problems: Collecting Global Expertise to Address the Problem of Harmful Cyanobacterial Blooms. A Lake Erie Case Study. Harmful Algae 2016, 54, 223–238. [Google Scholar] [CrossRef] [Green Version]
  73. Kubiňáková, E.; Híveš, J.; Gál, M.; Fašková, A. Effect of Ferrate on Green Algae Removal. Environ. Sci. Pollut. Res. 2017, 24, 21894–21901. [Google Scholar] [CrossRef] [PubMed]
  74. Ma, X.-Y.; Zhang, Z.-H.; Li, Q.-S.; Hu, S.-F.; Wang, H.-Y.; Li, J.; Chen, J.-M. Simultaneous removal of algae and its odor metabolites in raw water by potassium ferrate. Desalination Water Treat. 2013, 52, 357–364. [Google Scholar] [CrossRef]
  75. Addison, E.L.; Gerlach, K.T.; Spellman, C.D.; Santilli, G.; Fairbrother, A.R.; Shepard, Z.; Dudle, J.D.; Goodwill, J.E. Physicochemical implications of cyanobacteria oxidation with Fe(VI). Chemosphere 2021, 266, 128956. [Google Scholar] [CrossRef]
  76. Ma, J.; Liu, W. Effectiveness and mechanism of potassium ferrate(VI) preoxidation for algae removal by coagulation. Water Res. 2001, 36, 871–878. [Google Scholar] [CrossRef]
  77. He, H.-Y.; Qiu, W.; Liu, Y.-L.; Yu, H.-R.; Wang, L.; Ma, J. Effect of ferrate pre-oxidation on algae-laden water ultrafiltration: Attenuating membrane fouling and decreasing formation potential of disinfection byproducts. Water Res. 2021, 190, 116690. [Google Scholar] [CrossRef]
  78. Dong, F.; Liu, J.; Li, C.; Lin, Q.; Zhang, T.; Zhang, K.; Sharma, V.K. Ferrate(VI) pre-treatment and subsequent chlorination of blue-green algae: Quantification of disinfection byproducts. Environ. Int. 2019, 133, 105195. [Google Scholar] [CrossRef] [PubMed]
  79. Aslani, H.; Nabizadeh, R.; Nasseri, S.; Mesdaghinia, A.; Alimohammadi, M.; Mahvi, A.H.; Rastkari, N.; Nazmara, S. Application of response surface methodology for modeling and optimization of trichloroacetic acid and turbidity removal using potassium ferrate(VI). Desalination Water Treat. 2016, 57, 25317–25328. [Google Scholar] [CrossRef]
  80. Tiwari, D.; Sailo, L.; Pachuau, L. Remediation of aquatic environment contaminated with the iminodiacetic acid metal complexes using ferrate(VI). Sep. Purif. Technol. 2014, 132, 77–83. [Google Scholar] [CrossRef]
  81. Lan, B.; Wang, Y.; Wang, X.; Zhou, X.; Kang, Y.; Li, L. Aqueous arsenic (As) and antimony (Sb) removal by potassium ferrate. Chem. Eng. J. 2016, 292, 389–397. [Google Scholar] [CrossRef]
  82. Prucek, R.; Jan, K.; Tucek, J. Ferrate(Vi)-Lnduced Arsenite and Arsenate Removal by in Situ Structural Incorporation into Magnetic Iron(Iii) Oxide Nanoparticles. Environ. Sci. Technol. EST 2013, 47, 3283–3292. [Google Scholar] [CrossRef]
  83. Yu, M.; Chang, Y.; Tiwari, D.; Pachuau, L.; Lee, S.; Yang, J. Treatment of wastewater contaminated with Cd(II)–NTA using Fe(VI). Desalination Water Treat. 2012, 50, 43–50. [Google Scholar] [CrossRef]
  84. Yang, J.; Tiwari, D.; Yu, M.; Pachuau, L.; Kim, W.; Lee, S. Application of Fe(VI) in the treatment of Zn(II)-NTA complexes in aqueous solutions. Environ. Technol. 2010, 31, 791–798. [Google Scholar] [CrossRef]
  85. Prucek, R.; Solar, M.; Panacek, A.; Tucek, J.; Filip, J.; Zboril, R. Study of Antimony Removal Efficiency Using Ferrate (Vi). In Proceedings of the 10th Anniversary International Conference on Nanomaterials—Research & Application (Nanocon 2018 (R)), Brno, Czech Republic, 17–19 October 2018; 2019; pp. 214–219. [Google Scholar]
  86. Kolařík, J.; Prucek, R.; Tuček, J.; Filip, J.; Sharma, V.K.; Zbořil, R. Impact of inorganic ions and natural organic matter on arsenates removal by ferrate(VI): Understanding a complex effect of phosphates ions. Water Res. 2018, 141, 357–365. [Google Scholar] [CrossRef]
  87. Huang, Z.-S.; Wang, L.; Liu, Y.-L.; Zhang, H.-Y.; Zhao, X.-N.; Bai, Y.; Ma, J. Ferrate self-decomposition in water is also a self-activation process: Role of Fe(V) species and enhancement with Fe(III) in methyl phenyl sulfoxide oxidation by excess ferrate. Water Res. 2021, 197, 117094. [Google Scholar] [CrossRef] [PubMed]
  88. Petrov, V.G.; Perfiliev, Y.D.; Dedushenko, S.K.; Kuchinskaya, T.S.; Kalmykov, S.N. Radionuclide Removal from Aqueous Solutions Using Potassium Ferrate(Vi). J. Radioanal. Nucl. Chem. 2016, 310, 347–352. [Google Scholar] [CrossRef]
  89. Sun, X.; Zu, K.; Liang, H.; Sun, L.; Zhang, L.; Wang, C.; Sharma, V.K. Electrochemical synthesis of ferrate(VI) using sponge iron anode and oxidative transformations of antibiotic and pesticide. J. Hazard. Mater. 2017, 344, 1155–1164. [Google Scholar] [CrossRef] [PubMed]
  90. Wang, K.; Shu, J.; Wang, S.; Hong, T.; Xu, X.; Wang, H. Efficient electrochemical generation of ferrate(VI) by iron coil anode imposed with square alternating current and treatment of antibiotics. J. Hazard. Mater. 2020, 384, 121458. [Google Scholar] [CrossRef] [PubMed]
  91. Hu, S.H.; Wang, J.Q.; Zhu, C.Z.; Li, Y.X. Study on Potassium Ferrate Oxidative Degradation to Solution of P-Nitrophenol. In Proceedings of the 2010 International Conference on Mechanic Automation and Control Engineering, Wuhan, China, 26–28 June 2010; pp. 1–4. [Google Scholar]
  92. Homolková, M.; Hrabák, P.; Kolář, M.; Černík, M. Degradability of chlorophenols using ferrate(VI) in contaminated groundwater. Environ. Sci. Pollut. Res. 2015, 23, 1408–1413. [Google Scholar] [CrossRef]
  93. Han, Q.; Wang, H.; Dong, W.; Liu, T.; Yin, Y.; Fan, H. Degradation of Bisphenol a by Ferrate(Vi) Oxidation: Kinetics, Products and Toxicity Assessment. Chem. Eng. J. 2015, 262, 34–40. [Google Scholar] [CrossRef]
  94. Manoli, K.; Maffettone, R.; Sharma, V.K.; Santoro, D.; Ray, A.K.; Passalacqua, K.D.; Carnahan, K.E.; Wobus, C.E.; Sarathy, S. Inactivation of Murine Norovirus and Fecal Coliforms by Ferrate(VI) in Secondary Effluent Wastewater. Environ. Sci. Technol. 2020, 54, 1878–1888. [Google Scholar] [CrossRef]
  95. Poduska, R.A.; Anderson, B.D. Successful Storage Lagoon Odor Control. J. Water Pollut. Control Fed. 1981, 53, 299–310. [Google Scholar]
  96. Wang, Y.-Y.; Jin, Z.-S.; Song, H.-L.; Song, H.-L. Oxidation of benzyl alcohol by K2FeO4 to benzaldehyde over zeolites. Russ. J. Phys. Chem. A 2016, 90, 1931–1935. [Google Scholar] [CrossRef]
  97. Fatimah, S.; Widyawati, S.; Jannah, N.; Mei, A. Ferrate As an Advance Oxidator Materials (AOM) for Removing the Congo Red Dyes: Study of Optimization and Kinetics Degradation. J. Physics: Conf. Ser. 2021, 1858, 012060. [Google Scholar] [CrossRef]
  98. Lei, B.; Zhou, G.; Cheng, T.; Du, J. Synthesis of Potassium Ferrate by Chemical Dry Oxidation and Its Properties in Degradation of Methyl Orange. Asian J. Chem. 2013, 25, 27–31. [Google Scholar] [CrossRef]
  99. Dong, F.; Lin, Q.; Li, C.; He, G.; Deng, Y. Impacts of pre-oxidation on the formation of disinfection byproducts from algal organic matter in subsequent chlor(am)ination: A review. Sci. Total. Environ. 2021, 754, 141955. [Google Scholar] [CrossRef]
  100. Sharma, V.K.; Zboril, R.; Varma, R.S. Ferrates: Greener Oxidants with Multimodal Action in Water Treatment Technologies. Accounts Chem. Res. 2015, 48, 182–191. [Google Scholar] [CrossRef]
  101. Talaiekhozani, A.; Fulazzaky, M.A.; Ponraj, M.; Majid, M.Z.A. Removal of formaldehyde from polluted air in a biotrickling filter reactor. Desalination Water Treat. 2014, 52, 3663–3671. [Google Scholar] [CrossRef]
  102. Li, C.; Dong, F.; Feng, L.; Zhao, J.; Zhang, T.; Cizmas, L.; Sharma, V.K. Bacterial community structure and microorganism inactivation following water treatment with ferrate(VI) or chlorine. Environ. Chem. Lett. 2017, 15, 525–530. [Google Scholar] [CrossRef]
  103. Seidel, C.J.; Samson, C.C.; Bartrand, T.; Ergul, A.; Summers, R.S. Disinfection Byproduct Occurrence at Large Water Systems after Stage 2 Dbpr. J. Am. Water Work. Assoc. 2017, 109, 17–30. [Google Scholar] [CrossRef]
  104. Jiang, Y.; Goodwill, J.E.; Tobiason, J.E.; Reckhow, D.A. Comparison of Ferrate and Ozone Pre-Oxidation on Disinfection Byproduct Formation from Chlorination and Chioramination. Water Res. 2019, 156, 110–124. [Google Scholar] [CrossRef]
  105. Li, C.; Luo, F.; Dong, F.; Zhao, J.; Zhang, T.; He, G.; Cizmas, L.; Sharma, V.K. Chlorine decay and trihalomethane formation following ferrate(VI) preoxidation and chlorination of drinking water. Chemosphere 2017, 187, 413–420. [Google Scholar] [CrossRef] [PubMed]
  106. Yang, X.; Gan, W.; Zhang, X.; Huang, H.; Sharma, V.K. Effect of pH on the formation of disinfection byproducts in ferrate(VI) pre-oxidation and subsequent chlorination. Sep. Purif. Technol. 2015, 156, 980–986. [Google Scholar] [CrossRef]
  107. Barışçı, S. The disinfection and natural organic matter removal performance of electro-synthesized ferrate (VI). J. Water Process Eng. 2017, 20, 84–89. [Google Scholar] [CrossRef]
  108. Farré, M.J.; Day, S.; Neale, P.A.; Stalter, D.; Tang, J.Y.; Escher, B.I. Bioanalytical and chemical assessment of the disinfection by-product formation potential: Role of organic matter. Water Res. 2013, 47, 5409–5421. [Google Scholar] [CrossRef] [PubMed]
  109. Muellner, M.G.; Wagner, E.D.; McCalla, K.; Richardson, S.D.; Woo, Y.T.; Plewa, M.J. Haloacetonitriles Vs. Regulated Haloacetic Acids: Are Nitrogen-Containing Dbps More Toxic? Environ. Sci. Technol. 2007, 41, 645–651. [Google Scholar] [CrossRef]
  110. Zhang, M.-S.; Xu, B.; Wang, Z.; Zhang, T.-Y.; Gao, N.-Y. Formation of iodinated trihalomethanes after ferrate pre-oxidation during chlorination and chloramination of iodide-containing water. J. Taiwan Inst. Chem. Eng. 2016, 60, 453–459. [Google Scholar] [CrossRef]
  111. Li, C.; Li, X.; Graham, N. A study of the preparation and reactivity of potassium ferrate. Chemosphere 2005, 61, 537–543. [Google Scholar] [CrossRef]
  112. Feng, M.; Sharma, V.K. Enhanced oxidation of antibiotics by ferrate(VI)-sulfur(IV) system: Elucidating multi-oxidant mechanism. Chem. Eng. J. 2018, 341, 137–145. [Google Scholar] [CrossRef]
  113. Wu, S.; Li, H.; Li, X.; He, H.; Yang, C. Performances and mechanisms of efficient degradation of atrazine using peroxymonosulfate and ferrate as oxidants. Chem. Eng. J. 2018, 353, 533–541. [Google Scholar] [CrossRef]
  114. Tian, S.-Q.; Wang, L.; Liu, Y.-L.; Ma, J. Degradation of organic pollutants by ferrate/biochar: Enhanced formation of strong intermediate oxidative iron species. Water Res. 2020, 183, 116054. [Google Scholar] [CrossRef] [PubMed]
  115. Sun, M.; Huang, W.; Cheng, H.; Ma, J.; Kong, Y.; Komarneni, S. Degradation of Dye in Wastewater by Homogeneous Fe(Vi)/Nahso3 System. Chemosphere 2019, 228, 595–601. [Google Scholar] [CrossRef]
  116. Wu, S.; Liu, H.; Lin, Y.; Yang, C.; Lou, W.; Sun, J.; Du, C.; Zhang, D.; Nie, L.; Yin, K.; et al. Insights into mechanisms of UV/ferrate oxidation for degradation of phenolic pollutants: Role of superoxide radicals. Chemosphere 2020, 244, 125490. [Google Scholar] [CrossRef] [PubMed]
  117. Yuan, B.L.; Li, X.Z.; Graham, N. Aqueous Oxidation of Dimethyl Phthalate in a Fe(Vi)-Tio2-Uv Reaction System. Water Res. 2008, 42, 413–420. [Google Scholar] [CrossRef]
  118. Xia, T.; Qiu, Y.; Li, Y.; Wang, S.; Zhu, L.; Bi, J.; Zhou, J. Ceramsite sand and lime modified lake bottom sediments coupled with potassium ferrate pre-oxidation as emergency control of Microcystis aeruginosa blooms. J. Water Process Eng. 2021, 44, 102412. [Google Scholar] [CrossRef]
  119. Dunkin, N.; Weng, S.; Coulter, C.G.; Jacangelo, J.G.; Schwab, K.J. Reduction of Human Norovirus GI, GII, and Surrogates by Peracetic Acid and Monochloramine in Municipal Secondary Wastewater Effluent. Environ. Sci. Technol. 2017, 51, 11918–11927. [Google Scholar] [CrossRef] [PubMed]
  120. Feng, M.; Jinadatha, C.; McDonald, T.J.; Sharma, V.K. Accelerated Oxidation of Organic Contaminants by Ferrate(Vi): The Overlooked Role of Reducing Additives. Environ. Sci. Technol. 2018, 52, 11319–11327. [Google Scholar] [CrossRef]
  121. Pan, B.; Feng, M.; McDonald, T.J.; Manoli, K.; Wang, C.; Huang, C.-H.; Sharma, V.K. Enhanced ferrate(VI) oxidation of micropollutants in water by carbonaceous materials: Elucidating surface functionality. Chem. Eng. J. 2020, 398, 125607. [Google Scholar] [CrossRef]
  122. Sun, S.; Jiang, J.; Qiu, L.; Pang, S.; Li, J.; Liu, C.; Wang, L.; Xue, M.; Ma, J. Activation of Ferrate by Carbon Nanotube for Enhanced Degradation of Bromophenols: Kinetics, Products, and Involvement of Fe(V)/Fe(Iv). Water Res. 2019, 156, 1–8. [Google Scholar] [CrossRef]
  123. Kim, J.; Zhang, T.; Liu, W.; Du, P.; Dobson, J.T.; Huang, C.-H. Advanced Oxidation Process with Peracetic Acid and Fe(II) for Contaminant Degradation. Environ. Sci. Technol. 2019, 53, 13312–13322. [Google Scholar] [CrossRef]
  124. Wang, J.; Kim, J.; Ashley, D.C.; Sharma, V.K.; Huang, C.H. Peracetic Acid Enhances Micropollutant Degradation by Ferrate(Vi) through Promotion of Electron Transfer Efficiency. Environ. Sci. Technol. 2022, 56, 11683–11693. [Google Scholar] [CrossRef]
  125. Manoli, K.; Li, R.; Kim, J.; Feng, M.; Huang, C.-H.; Sharma, V.K. Ferrate(VI)-peracetic acid oxidation process: Rapid degradation of pharmaceuticals in water. Chem. Eng. J. 2022, 429, 132384. [Google Scholar] [CrossRef]
  126. Li, C.; Lin, H.; Armutlulu, A.; Xie, R.; Zhang, Y.; Meng, X. Hydroxylamine-assisted catalytic degradation of ciprofloxacin in ferrate/persulfate system. Chem. Eng. J. 2018, 360, 612–620. [Google Scholar] [CrossRef]
  127. Shao, B.; Dong, H.; Feng, L.; Qiao, J.; Guan, X. Influence of [Sulfite]/[Fe(Vi)] Molar Ratio on the Active Oxidants Generation in Fe(Vi)/Sulfite Process. J. Hazard. Mater. 2020, 384, 121303. [Google Scholar] [CrossRef]
  128. Yang, T.; Wu, S.; Mai, J.; Chen, L.; Huang, C.; Zeng, G.; Wu, Y.; Zhu, M.; Huang, Y.; Mo, Z.; et al. Activation of ferrate(VI) by sulfite for effectively degrading iodinated contrast media and synchronously controlling I-DBPs formation. Chem. Eng. J. 2022, 442, 136011. [Google Scholar] [CrossRef]
  129. He, C.; Zhang, X.; Lv, P.; Sui, H.; Li, X.; He, L. Efficient remediation of o-dichlorobenzene-contaminated soil using peroxomonosulfate-ferrate-FeS hybrid oxidation system. J. Taiwan Inst. Chem. Eng. 2021, 119, 23–32. [Google Scholar] [CrossRef]
  130. Mohammadi, S.; Moussavi, G.; Shekoohiyan, S.; Marín, M.L.; Boscá, F.; Giannakis, S. A continuous-flow catalytic process with natural hematite-alginate beads for effective water decontamination and disinfection: Peroxymonosulfate activation leading to dominant sulfate radical and minor non-radical pathways. Chem. Eng. J. 2021, 411, 127738. [Google Scholar] [CrossRef]
  131. He, H.; Zhao, J. The Efficient Degradation of Diclofenac by Ferrate and Peroxymonosulfate: Performances, Mechanisms, and Toxicity Assessment. Environ. Sci. Pollut. Res. 2022, 30, 11959–11977. [Google Scholar] [CrossRef] [PubMed]
  132. Fei, Y.; Liu, Z.; Meng, L.; Liu, G.; Kong, D.; Pan, X.; Zhu, F.; Lu, J.; Chen, J. Experimental and theoretical study on Fe(VI) oxidative degradation of dichlorophen in water: Kinetics and reaction mechanisms. Environ. Pollut. 2022, 306, 119394. [Google Scholar] [CrossRef]
  133. Feng, M.; Cizmas, L.; Wang, Z.; Sharma, V.K. Synergistic Effect of Aqueous Removal of Fluoroquinolones by a Combined Use of Peroxymonosulfate and Ferrate(Vi). Chemosphere 2017, 177, 144–148. [Google Scholar] [CrossRef] [PubMed]
  134. Kralchevska, R.P.; Prucek, R.; Kolařík, J.; Tuček, J.; Machala, L.; Filip, J.; Sharma, V.K.; Zbořil, R. Remarkable efficiency of phosphate removal: Ferrate(VI)-induced in situ sorption on core-shell nanoparticles. Water Res. 2016, 103, 83–91. [Google Scholar] [CrossRef] [PubMed]
  135. López-Vinent, N.; Cruz-Alcalde, A.; Moussavi, G.; Gonzalez, I.D.C.; Lehmann, A.H.; Giménez, J.; Giannakis, S. Improving ferrate disinfection and decontamination performance at neutral pH by activating peroxymonosulfate under solar light. Chem. Eng. J. 2022, 450, 137904. [Google Scholar] [CrossRef]
  136. Talaiekhozani, A.; Salari, M.; Talaei, M.R.; Bagheri, M.; Eskandari, Z. Formaldehyde Removal from Wastewater and Air by Using Uv, Ferrate(Vi) and Uv/Ferrate(Vi). J. Environ. Manag. 2016, 184, 204–209. [Google Scholar] [CrossRef]
  137. Chen, Z.; Yang, B.; Wen, Q.; Tang, Y. Application of potassium ferrate combined with poly-aluminum chloride for mitigating ultrafiltration (UF) membrane fouling in secondary effluent: Comparison of oxidant dosing strategies. Chemosphere 2021, 274, 129862. [Google Scholar] [CrossRef] [PubMed]
  138. Liu, Y.; Zhang, J.; Huang, H.; Huang, Z.; Xu, C.; Guo, G.; He, H.; Ma, J. Treatment of Trace Thallium in Contaminated Source Waters by Ferrate Pre-Oxidation and Poly Aluminium Chloride Coagulation. Sep. Purif. Technol. 2019, 227, 115663. [Google Scholar] [CrossRef]
  139. Zhang, X.; Zhu, X.; Li, H.; Wang, C.; Zhang, T. Combination of Peroxymonosulfate and Fe(Vi) for Enhanced Degradation of Sulfamethoxazole: The Overlooked Roles of High-Valent Iron Species. Chem. Eng. J. 2023, 453, 139742. [Google Scholar] [CrossRef]
Figure 1. FeO42− simulated structure diagram.
Figure 1. FeO42− simulated structure diagram.
Water 15 00699 g001
Figure 2. Electrochemical preparation of ferrate(VI) treatment of wastewater.
Figure 2. Electrochemical preparation of ferrate(VI) treatment of wastewater.
Water 15 00699 g002
Figure 3. The degradation of SCP by ferrate(VI).
Figure 3. The degradation of SCP by ferrate(VI).
Water 15 00699 g003
Figure 4. Schematic diagram of ferrate(VI) disinfection.
Figure 4. Schematic diagram of ferrate(VI) disinfection.
Water 15 00699 g004
Figure 5. Free radicals produced by UV/ferrate(VI) systems.
Figure 5. Free radicals produced by UV/ferrate(VI) systems.
Water 15 00699 g005
Figure 6. Difference between ferrate(VI)–PAA system and ferrate(VI) alone.
Figure 6. Difference between ferrate(VI)–PAA system and ferrate(VI) alone.
Water 15 00699 g006
Table 1. Redox potentials of commonly used oxidants/disinfectants.
Table 1. Redox potentials of commonly used oxidants/disinfectants.
Oxidants/DisinfectantsReaction TypeE0/V
Cl2 C l 2 ( g ) + 2 e 2 C l 1.358
ClO2 C l O 2 ( aq ) + e C l O 2 0.854
O3 O 3 + 2 H + + 2 e O 4 + H 2 O 2.076
KMnO4 M n O 4 + 4 H + + 3 e M n O 2 + 2 H 2 O 1.679
M n O 4 + 8 H + + 5 e M n 2 + + 4 H 2 O 1.507
Ferrate(VI) F e O 4 2 + 8 H + + 5 e F e 3 + + 4 H 2 O 2.20
F e O 4 2 + 4 H 2 O + 3 e F e ( O H ) 3 + 5 O H 0.76
Table 2. Ferrate(VI) removal capacity of common pollutants in water.
Table 2. Ferrate(VI) removal capacity of common pollutants in water.
Type of PollutantName of PollutantpHMolar Ratio of Ferrate(VI) to PollutantsTime/minRemoval Rate/%References
MedicineSMZ:5–920:1100[89]
Mycin4–95:180[90]
SDZ720:175[59]
Sulfonamide36:115100
Cotinine6.590
Sulfachloropyridazine710:14090[51]
Atrazine725:1100[32]
Phenolic compounds4-tBP10:15100[20]
p-Nitrophenol53:15080[91]
Pentachlorophenol9[92]
BPA86099[69]
Tetrabromobisphenol A74:1399.06[93]
Bisphenol F7.5–103.33:1594[94]
AlcoholsEthanol615:13080[95]
Chain hydrocarbon derivativesIminodiacetic acid8–1090[80]
Organophosphorus pesticidesClorpyrifos7100:15100[47]
AldehydesBenzaldehyde96[96]
Dye wastewaterRhodamine B78100[97]
Congo Red9.25:12098.2
Methyl Orange66099.2[98]
AlgaeAeruginosa suspension77:1570.25[99]
Chlorella sp.8–93:146.2[78]
P. limnetica8–93:158.1
Table 3. Advantages and disadvantages of different disinfectants.
Table 3. Advantages and disadvantages of different disinfectants.
Disinfectant NameAdvantagesDisadvantages
ChlorineLow cost
Strong disinfection ability
Raw materials are readily available
Long-term preservation
Produces toxic DBPs
Dechlorination is required
Harm to health
OzoneThe by-product is oxygen, which is non-toxic and harmless
Strong disinfection capacity and fast rate
Remove odors
High cost
A small amount of virus is produced Harmful to the human body
Corrosive equipment
Ferrate(VI)Strong disinfection ability
No disinfection by-products
The reaction product has a coagulation effect
The process is simple, the footprint is small, and the cost is low
Ferrate(VI) yields are low
Ferrate(VI) has poor stability and is difficult to store for a long time
Table 4. Control mechanism of ferrate(VI) on DBPs.
Table 4. Control mechanism of ferrate(VI) on DBPs.
Disinfection MethodType of DBPsName of DBPsPrecursorsMechanism of ActionReferences
Chlorine disinfectionTHMsChloroformNOM, AOM, Phenols, Benzene, Amino acidsThe action of ferrate(VI) completely removes hydrophobic organic molecules and partially removes hydrophilic organic molecules, and reduces the formation potential of THMs, thus controlling the amount of THMs produced[107]
Bromodichloromethane
Bromomimetic
HAAsChloroacetic acidFerrate(VI) reduces the amount of DBPs produced by chlorine disinfection by oxidizing organic matter in the water to acids that are highly tolerant to chlorine[108]
Dichloroacetic acid
HANsChloroacetonitrileFerrate(VI) oxidizes organic matter in water to acids that are highly resistant to chlorine, thereby reducing the amount of DBPs produced by chlorine disinfection[109]
Bromoacetonitrile
Ozone disinfectionBromateSodium bromateBromide (Br)Ferrate(VI) has no reaction properties against Br under medium-alkaline conditions, so controlling the pH of the reaction is a crucial condition for ferrate(VI) control of bromate[30]
Chloramine disinfectionIodine-substituted DBPsDichloroiodomethaneIodide (I)High doses of ferrate(VI) oxidize iodide directly to iodate, which does not react with chlorine, thereby reducing the amount of iodosterilization by-products[110]
Dibromoiodomethane
Iodoform
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

Yu, J.; Sumita; Zhang, K.; Zhu, Q.; Wu, C.; Huang, S.; Zhang, Y.; Yao, S.; Pang, W. A Review of Research Progress in the Preparation and Application of Ferrate(VI). Water 2023, 15, 699. https://doi.org/10.3390/w15040699

AMA Style

Yu J, Sumita, Zhang K, Zhu Q, Wu C, Huang S, Zhang Y, Yao S, Pang W. A Review of Research Progress in the Preparation and Application of Ferrate(VI). Water. 2023; 15(4):699. https://doi.org/10.3390/w15040699

Chicago/Turabian Style

Yu, Jianping, Sumita, Kai Zhang, Qijia Zhu, Chengzhang Wu, Saikai Huang, Yunshu Zhang, Sijing Yao, and Weihai Pang. 2023. "A Review of Research Progress in the Preparation and Application of Ferrate(VI)" Water 15, no. 4: 699. https://doi.org/10.3390/w15040699

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