NC State
BioResources
Guo, W., and Huang, J. (2023). "Environmental applications of immobilized and bio-resourced redox mediators: A Review," BioResources 18(1), 2327-2350.

Abstract

Redox mediators (RMs), also known as electron shuttles, have been widely reported to promote both biotic and abiotic reductions of oxidized pollutants in water, soil, biogeochemical cycles, and wastewater treatment systems. However, the continuous addition of dissolved RMs is unaffordable and the potential environmental risks remain unknown because most applied RMs are synthetic chemicals. Immobilization technology enables RMs to be attached on non-dissolved supports, avoiding wash-out from the treatment systems. This realizes the reuse of RMs in scaled-up engineering applications and the in-situ remediation. Moreover, renewable natural biomass and their derivatives, such as biochar, have also aroused increased interest because they provide an economical and feasible way to solve the shortcomings of applying soluble RMs. This review presents different RM immobilization methods, which include entrapment, adsorption, and surface modification, as well as the use of bio-resourced RMs. The immobilization procedures and reaction mechanisms of the immobilized RMs and bio-resourced RMs in environmental applications are critically compared and summarized.


Download PDF

Full Article

Environmental Applications of Immobilized and Bio-Resourced Redox Mediators: A Review

Wenrui Guo,a and Jingang Huang b,c,*

Redox mediators (RMs), also known as electron shuttles, have been widely reported to promote both biotic and abiotic reductions of oxidized pollutants in water, soil, biogeochemical cycles, and wastewater treatment systems. However, the continuous addition of dissolved RMs is unaffordable and the potential environmental risks remain unknown because most applied RMs are synthetic chemicals. Immobilization technology enables RMs to be attached on non-dissolved supports, avoiding wash-out from the treatment systems. This realizes the reuse of RMs in scaled-up engineering applications and the in-situ remediation. Moreover, renewable natural biomass and their derivatives, such as biochar, have also aroused increased interest because they provide an economical and feasible way to solve the shortcomings of applying soluble RMs. This review presents different RM immobilization methods, which include entrapment, adsorption, and surface modification, as well as the use of bio-resourced RMs. The immobilization procedures and reaction mechanisms of the immobilized RMs and bio-resourced RMs in environmental applications are critically compared and summarized.

DOI: 10.15376/biores.18.1.Guo

Keywords: Immobilization; Redox mediator; Entrapment; Surface modification; Biomass

Contact information: a: PowerChina Huadong Engineering Corporation Limited, Hangzhou 311122, PR China; b: College of Materials and Environmental Science, Hangzhou Dianzi University, Hangzhou 310018, PR China; c: The Belt and Road Information Research Institute, Hangzhou Dianzi University, Hangzhou 310018, PR China; *Corresponding author: hjg@hdu.edu.cn

INTRODUCTION

In recent decades, the global environment has been challenged by serious pollution caused by synthetic chemicals, which are often recalcitrant, biorefractory, and carcinogenic. These contaminants include but are not limited to nitroaromatic compounds (NACs), azo dyes, nitrate, perchlorate, sulfate, halogenated compounds, polyhalogenated pollutants, hexavalent chromium (Cr(VI)), persistent organic pollutants, and so on. Because of the electron-withdrawing features of special groups associated with these oxidized pollutants, aerobic conditions usually fail to treat them (van der Zee and Cervantes 2009; Zhang et al. 2020). However, they can be moderately reduced under anaerobic conditions when electron donors are available, whereupon they can be readily removed aerobically. Thus, combined anaerobic-aerobic processes are typically applied to treat oxidized pollutants or concerned substances. In such approaches, the anaerobic reduction is the rate-limiting step and the bottleneck part of the process.

Redox mediators (RMs) are vital important compounds capable of accelerating redox reactions by shuttling electrons between their reduced and oxidized forms, as well as of lowering the activation energy for redox reactions. Within this catalytic process, the electrons from the primary electron donors would be quickly transferred to the final electron acceptors (Olivo-Alanis et al. 2018; Li et al. 2022a). They have been widely applied in the oxidation of H2, alcohols, biomass, cellulose, and so on (Anson and Stahl 2020; Chen et al. 2022). To promote the reductive transformation of oxidized pollutants, the use of RMs has also been extensively reported for both biotic and abiotic reduction of oxidized pollutants in water, soil, biogeochemical cycles, and wastewater treatment systems (Rau et al. 2002; O’Loughlin 2008; Uchimiya and Stone 2009; Song et al. 2021). Currently, the most frequently applied RMs consist of flavin-based compounds and quinone-based compounds. Flavin-based compounds are those such as flavin adenine dinucleotide, flavin mononucleotide, and riboflavin; quinone-based compounds include, among others, anthraquinone-2,6-disulfonate (AQDS), anthraquinone-2-sulfonate (AQS), juglone, lawsone, and natural organic matters (NOM ), such as humic acids (HA). Besides, 2,2,6,6-tetramethylpiperidine-1-oxyl (TEMPO) is another commercially available RM by the TEMPO/TEMPOH redox cycle (Zhan et al. 2019; Sun et al. 2022; Zhang et al. 2021). The chemical structure of three typical RMs and the conversion between their oxidized and reduced forms are shown in Fig. 1 (Buckel and Thauer 2018; Zhang et al. 2021).

Fig. 1. The chemical structure and conversion cycle of flavin-based (a), quinone-based (b) and TEMPO (c) RMs

The RMs have been extensively reported to accelerate biological/chemical reduction of oxidized pollutants in various bioreactors and further prevent the toxic inhibition of them to the biological processes (van der Zee et al. 2001; Liu et al. 2009; Costa et al. 2010; Saratale et al. 2011; Mani et al. 2018). In these processes, RMs could speed up the electron transferring from the electron donors – carbohydrate, protein, volatile fatty acids, alcohol, hydrogen, (biogenic)sulfide, zero valent iron and ferrous iron, nicotinamide adenine dinucleotide phosphate hydrate, and electrical power – to the acceptors of oxidized contaminants, increasing the reduction rates by several folds or even dozens. Mixed cultures of microorganisms (such as activated sludge) as well as pure strains have been reported to contribute to the above mediated processes (Borch et al. 2005; dos Santos et al. 2005; Bhushan et al. 2006; Kwon and Finneran 2006, 2008; Bibi et al. 2019; Liu et al. 2021; Ren et al. 2022). Moreover, TEMPO-mediated decolorization, cellulose oxidation, photocatalysis and electroreduction have also been reported in the literature, implying the electron shuttling role of TEMPO/TEMPOH couple (Gharehkhani et al. 2019; Wen et al. 2019; Lukyanov et al. 2022).

In practical applications, the continuous addition of soluble RMs has been unaffordable because of their washout from the treatment systems. Moreover, most applied RMs are synthetic chemicals, for which the potential environmental and health risks are still unknown. Thus, urgent ways are needed for avoiding or mitigating the loss of soluble RMs (Zhou et al. 2012; Zhang et al. 2020). These methods include decreasing the soluble RM dosage, immobilizing RMs on special materials, and/or utilizing economical and naturally available alternatives. The RM immobilization enables soluble RMs to be attached to non-dissolved and inert carriers that fulfill the electron shuttling role in the redox process. This technology has aroused the most interest because it provides an economical and feasible way to solve the shortcomings of applying soluble RMs in continuously operating reactors. The abundance of target pollutants followed the order: azo compounds (52%) > polyhalogenated pollutants (15%) > NACs (13%) > nitrate and/or nitrite (12%) > Fe (4%) and Cr(VI) (3%) (Dai et al. 2016). Furthermore, some natural biomass or their derivatives, such as activated carbon, biochar, and natural humic substance (HS) are also capable of shuttling electrons during the redox reactions, providing cost-saving alternatives in the future. However, the direct use of natural biomass as RM would bring problems such as the release of unwanted organic compounds, which might increase secondary pollution.

The immobilization of soluble RMs is a complex process, depending on the physical-chemical features of both RMs and solid carriers. Thus, the selection of immobilization methods for specific chemical RMs has always been challenging. To guide the future applications of RMs in the environmental field, it is necessary to summarize different RM immobilization methods, focusing on immobilization procedures, and further prospects; in addition, natural and bio-resourced RMs are also proposed because of their economical and eco-friendly advantages. This review paper attempts to (1) summarize different RM immobilization methods (procedures and mechanisms) such as entrapment, adsorption, and surface modification, (2) investigate the catalytic effect and bottlenecks of various immobilized RMs in environmental applications, and (3) select bio-resourced RMs to speed up the bioredox processes for pollutants the removal.

IMMOBILIZATION METHODS AND APPLICATIONS

Entrapment-type Immobilization

In biological systems, entrapment has been traditionally considered for the physical retention or fixation of microorganisms and/or enzymes by polymeric carriers such as alginate, nanomaterials, silica gel, polyacrylamide, gelatin, polyion complex, and so on (Moyo et al. 2012; Romero-Soto et al. 2021). This method keeps the suspended cells and/or enzymes trapped in a porous solid matrix and thus condenses them, preventing their wash-out or loss from environmental systems or developing enzyme electrodes/biosensors for bioelectrochemical systems (Sakurada et al. 2017). Therefore, immobilization using polymeric materials as carriers of bioactive components can improve the efficiency and stability of microorganisms compared to that with free cells and/or enzymes, thus providing a promising technology to mitigate the pollutants in the environment (Fernández-Fernández et al. 2013; Somu et al. 2022).

The RMs, like microorganisms and enzymes, also can be non-covalently entrapped in different porous polymeric substances. The reported entrapmental immobilization methods of RMs and their applications in the reductive biotransformation of oxidized pollutants and bioelectrocatalysis capability are listed in Table 1. It indicates that the entrapmental immobilization is quite simple and the operating conditions are relatively moderate, meaning that high temperature and/or high pressure are not required during the redox reaction. Calcium alginate (CA), polyvinyl alcohol (PVA), and agar gel are the popularly used polymer supports for RMs entrapment because of their relatively easy procedures and low costs (Guo et al. 2007). The RMs entrapped on CA/PVA beads have been used in the bio-decolorization and denitrification processes. In these RM-assisted systems, the adsorption of pollutants by CA and PVA polymer beads only accounted for 0.2% to 2.0% of the total removal, indicating that the main removal pathways were via bio-reduction. Under the electron-shuttling assistance of polymer-trapped RMs, the removal of azo dyes and nitrate would be promoted to some extent with yeast extract, peptone, acetate, and glucose as electron donors (Liu et al. 2012b). Additionally, silica-entrapped TEMPO is an efficient heterogeneous catalyst and is more environmental-friendly and less tedious during immobilization, attracting attentions on the selective oxidation of cellulose, glucoside, and alcohols (Palmisano et al. 2006; Chen et al. 2022).

Table 1. Entrapment-type Immobilization of RMs and Their Applications

Although presents entrapment has the advantages in immobilizing RMs, it would wrap the selected RM groups into the the polymer matrix. This means that the substrates, including electron donors and acceptors, must pass through the porous matrix to reach active RM groups and react with them, thereby limiting the electron-transferal during the bio-reduction process. Previous studies indicated that the RMs immobilized by entrapment are only capable of increasing the bio-decolorization/bio-denitrification rates up to two-fold compared to rates in the control systems (Guo et al. 2007, 2010; Liu et al. 2012b). In contrast, the RM polymer matrix might be disrupted because of the weakening mechanical strength of CA/PVA/Sol-gel beads after the long operation, resulting in the blockage of the matrix pores. To maintain the effective entrapment of soluble RMs on a more durable polymeric matrix than CA and PVA, hydrophobic RMs, such as anthraquinone, AQS, and 1,5-dichloroanthraquinone are preferred. Feng et al. (2017) developed a “foldable” but strengthened RM, i.e., AQSA-Na (anthraquinone-2-sulfonic acid sodium salt)-doped PVA-H2SO4 robust gel film electrolyte, and suggested a great contribution of loaded RMs to the outstanding electron-transferring capability. This advance enabled the fabrication of smarter bio-electrochemical systems for the continuous removal of oxidized pollutants.

For the future application of RM entrapment as a technique to enhance pollutants removal, the mechanical strength of the polymer matrix and the diffusion rates of soluble substrates must be improved. To solve these problems, the co-entrapment of RMs and microorganism/enzymes seems promising, as this approach provides a better electron/mass transferring condition between microbial cells and entrapped RMs, thereby stably accelerating the reductive transformation of oxidized pollutants. Su et al. (2009) reported that the co-entrapment of RMs and microorganisms in a polymer matrix achieved better decolorization performance than was achieved with mono-immobilization of only RMs or microorganisms. Repeated decolorization batch experiments using co-entrapment systems showed that the RMs were reusable after several operating batch cycles (Su et al. 2009; Sharma et al. 2016).

Adsorptive Immobilization

Adsorption is another frequently employed RM immobilization method (Dai et al. 2016). After adsorption of soluble RMs into the surface of solid adsorbents, the electron-shuttling groups can be concentrated, enabling their reutilization during the redox reaction of oxidized pollutants. The adsorptive types by which soluble RMs are incorporated into special adsorbents include physical and chemical pathways. Physical adsorption is governed by the van der Waals intermolecular force, which weakly binds the soluble RMs with the adsorbent’s surface and the adsorption was even reversible (Wang et al. 2010). Thus, physically adsorped RMs can be easily detached from their carrier. For a wastewater treatment system as an example, a fluctuating influent can easily wash off physically immobilized RMs from the adsorbents. However, chemically adsorbed RMs are co-joined by chemical bonds, such as covalent bonds and hydrogen bonds, which are much stronger and more stable than physical force. Consequently, chemical-adsorptive RM immobilization has been widely used, depending on the linked functional groups (hydrophilicity/hydrophobicity) and/or charges (positive/negative) between adsorbents and RMs (Cervantes et al. 2010). As indicated in recent reports, ceramsite-based products, anion exchange resin, cellulose acetate, activated carbon (AC), polyurethane foam, porous silica beads, and metal-oxides nanoparticles are typically used as the potential adsorbents to immobilize different soluble RMs (Table 2).

These above immobilization procedures are relatively easy, but in some cases they are costly due to the usage of some advanced adsorptive materials such as nanoparticles or metal-organic frameworks (Li et al. 2017; Lou et al. 2023). The adsorption performance of an adsorbent to immobilize RMs is usually determined by adsorption isotherms modeling, the result of which provides proper guidance to maximize the electron-transferring capacity. Functional groups of chemical-adsorptive immobilized RMs are tightly attached to the adsorbent’s surface. This leads to much better mass/electron transferal capability as compared with entrapmental immobilized RMs, which was inside the polymer matrix (Dai et al. 2016). Thus, the pollutant removal rates achieved using adsorptive immobilized RMs (maximally 10.4-fold) are much higher than that achieved using entrapped immobilization RMs (maximally 2.1-fold). Furthermore, benefitted by immobilized RMs, an anaerobic consortium can even utilize toxic substrates (such as phenol) as the sole energy source (electron donor) (Martínez et al. 2013).

Surface modification could potentially improve the adsorption capacity, enabling RMs to be immobilized as designed. Yuan et al. (2012) produced organic–inorganic hybrid materials using adsorption/covalence coupling methods. With OH-ceramsite and NH2-ceramsite as a base, the sulfonic acid group associated in AQS can covalently bind with the -OH and -NH2 bonds to form AQS-ceramsite. The grafted AQS on ceramsite has been shown to effectively catalyze the bio-decolorization process in a high salt environment. Furthermore, Fe3O4-quinone/TEMPO nanocomposites have also been identified as a promising RM complex because they can be easily and quickly separated from the mixed liquor with the assistance of a magnetic field; once separated, they can then be reused. Zhang and Hu (2017) introduced nano-scale Fe3O4 as an adsorbent to immobilize quinone-based RMs, achieving efficient and stable bio-reduction of NACs during multiple use cycles. Gao et al. (2018) used amino-functionalized magnetic Fe3O4 nanoparticles as a support to develop a co-immobilize TEMPO/laccase complex to remove acid fuchsin, and 50% residual activity was retained after eight cycles of operation.

Table 2. Adsorptive Immobilization of RMs and Their Applications

Solid Carbon Materials and their Modification

Solid carbon-based materials can be derived from carbon-enriched substances including coal, petcoke, resin, and agricultural/forestry residues. The commercially available carbon materials include powdered activated carbon, activated carbon felt, carbon paper, carbon nanotubes, graphene oxide (GO), graphite electrodes, and so on. These carbon materials were associated with functional groups of carboxyl, quinone, carbonyl, lactone, hydroxyl, and carboxylic anhydride in their surface, enhancing the reductive biotransformation of various oxidized pollutants by improving electron shuttling ability (van der Zee et al. 2003; Pereira et al. 2014; Colunga et al. 2015; Amezquita-Garcia et al. 2016). However, the density of these active groups of solid carbon materials is at a relatively low level, limiting the electron-shuttling rates during the pollutant removal processes. Thus, carbon materials can slightly enhance pollutant removal (maximally 3.7-fold over the control) in chemical and/or biological decolorization, denitrification, anammox, and NAC reduction processes. Although GO and reduced graphene oxide (rGO), with proper size distribution, unique spatial structure, and better redox potential, lead to a higher redox conversion of oxidized pollutant (up to 10-fold), they are costly for full-scale applications (Yin et al. 2015; Colunga et al. 2015; Li et al. 2016b).

In consideration of the above drawbacks, researchers have attempted to selectively modify the surface chemistry of carbon materials to promote the richness of desired groups and thereby develop the electron-transferal capacity for contaminants removal (Li et al. 2016a). Thermal (heat), chemical (acidic and alkaline/basic treatments), electrochemical, and biological (bio-adsorption) modification methods are typically used to change the surface characteristics of carbon materials, depending on their group features (acidic, basic, and/or neutral) (Yin et al. 2007). Research studies regarding different modification methods to fix RMs are briefly summarized in Table 3. For example, Pereira et al. (2010) compared chemical oxidation (with HNO3 and O2) and thermal treatments (in an H2/N2 atmosphere) in modifying the surface chemistry of commercial AC, suggesting that chemical modification introduced more quinone groups than did thermal treatments. In this case, the first-order reduction rate constants of azo dyes increased by 9.0-fold compared with the control treatment. Further, a two-step procedure consisting of amination and the Buchwald–Hartwig reaction was also developed to immobilize quinone-based RMs, efficiently mediating the reductive removal of azo dyes, nitrate, and NACs (Zhang et al. 2014b; Xu et al. 2015).

Table 3. Immobilization of RMs by Surface Modification and Their Applications

In (bio)electrochemical systems, electrodeposition, adsorption, and crossing-linking have been successfully applied to prepare RM-modified electrodes for use in MFC and microbial electrolysis cell (MEC) systems; which in turn benefit the electron-transferal during redox reactions, enhancing the bioelectricity generation, pollutant removal, and energy storage (Table 3). In addition, electropolymerization has been used to increase the required density of electron shuttling groups on the surface of carbon materials or functional composites. This method is usually a coating procedure to enable a conducting polymer to be fixed or deposited into a conducting material from a solution, which was previously used in the enzyme immobilization (such as encapsulating glucose oxidase electrochemical biosensors) by choosing a defined electrical potential and current (Chiorcea-Paquim et al. 2008). When applying the electropolymerization to immobilize RMs on an electrode, attempts have been made to coat polymer materials, such as polypyrrole and pyrrole, on a carbon electrode by doping designed RMs. This polymer/RMs composite film has a vast surface area and pseudo-capacitance, which has aroused great interest in its application in electricity-assisted or electricity-produced systems (Lang et al. 2010; Anwer et al. 2021). Regarding the interest in RMs to enhance pollutant removal, this composite film can also promote the functional group density and prevent the wash-out of RMs from systems. Compared to the entrapped AQDS/PVA particles, AQDS/electropolymerization-modified anode in an MFC system shows better performance in both electricity generation and pollutants removal (Martinez et al. 2017).

Natural and Bio-resourced Redox Mediator

The previously mentioned RM immobilization technology has enabled the separation of RM retention time from hydraulic retention time in environmental applications. However, the complex procedures of these immobilization processes, the time-consuming lab work to select suitable process conditions and unaffordable carriers always limited their wide applications (Rocha-Martín et al. 2021; Wang et al. 2021). To overcome these raised issues, data-driven approaches such as machine learning algorithms have been attempted to predict and optimize those complex procedures and to obtain feasible immobilization conditions (Shi et al. 2022). Moreover, some natural biomass and/or their derivatives such as biochar would be expected to be good and low-cost alternatives because of the renewable raw materials. The reported uses of these materials, i.e., natural biomass and bio-resourced derivates in facilitating pollutant removal are shown in Tables 4 and 5, respectively.

Henna plant (Lawsonia inermis) biomass (HPB) has been reported to be a natural source of lawsone (an effective RM), with lawsone comprising up to 1.8% of the plant’s dry weight (Almeida et al. 2012). Previous studies indicated that natural HPB, including stem, leaf, flower, and seed, could serve as both an electron donor and RM source to improve the bio-reduction of azo dyes and Cr(VI) in batch and continuous bioreactors, all showing positive effects of HPB on pollutants removal (Huang et al. 2014, 2015, 2016a 2021; Tang et al. 2016; Rau et al. 2002). Moreover, HPB could be co-fermented with waste sludge in municipal wastewater treatment plants and enhance the production of VFAs (Huang et al. 2016b). As a protein-rich biomass, waste sludge could be firstly hydrolyzed to amino acid, and then converted to carboxylic acid and NH4+-N by the Stickland redox reaction (Huang et al. 2019). The associated RM of lawsone in HPB could speed up the transfer of electron from one amino acid to another, resulting in higher VFAs production. Moreover, another waste biomass of Punica granatum peal was also approved as a natural RM to efficiently degrade orange G dye (Bibi et al. 2019). In contrast, some extractions from plant biomass or algae, such as chlorophyll, laccase, syringaldehyde, and acetosyringone, were also reported to serve as natural RMs for enhanced denitrification, decolorization, and electricity generation (Ma et al. 2017; Mani et al. 2018; Lu et al. 2020; Song et al. 2021). Furthermore, a novel biomass-derived lignin with electron transfer 3D networks was developed to act as cathode interlayer, achieving high-efficient performance of solar cells (Hu et al. 2020). However, raw plant biomass or its extractions without proper treatment can cause problems such as secondary pollutants (e.g., organic substances lost from biomass as well as wash-out of associated RMs). To address these problems in engineering applications, researchers have suggested that natural biomass could be prepared as biochar to retain or modify most of the active electron transferral groups.

Table 4. Natural Biomass Acting as Redox Mediator

Biomass-resourced derivatives including biochar and natural HS are listed in Table 5. Biochar is a porous bio-resourced carbon material that has rich functional groups and can be widely obtained from natural biomass, especially cellulosic, hemicellulosic, and lignin agricultural and forestry residues. The biochar is usually prepared via pyrolysis, forming graphitized structure and redox active sites (functional groups such as quinone groups) in its surface, and is functioned as extracellular electron mediation for pollutant removal, contamination transformation, and in-situ remediation (Chen et al. 2019; Zhang et al. 2020; Fang et al. 2020; Yuan et al. 2022). Pyrolytic temperature, heating rate, and holding time are the effective impactors for shaping surface properties of biochar. The electron transfer/exchange capacity (EEC) and hydrophilicity of biochars prepared under different conditions were significantly varied. Although higher pyrolytic temperature (700 to 800 °C) can improve the hydrophobicity, the density of electron transfer groups, such as quinone groups, on the biochar surface decreased, resulting in weakened EEC for redox process (Oliveira et al. 2017; Yuan et al. 2017). In contrast, lower pyrolytic temperature (300 to 600 °C) could enrich the functional mediating groups for the reduction of hydrophilic contaminants (Zhao et al. 2018; Zhang et al. 2022); however, the surface hydrophobicity decreased, which is unfavorable for the removal of hydrophobic contaminants (Xu et al. 2020). Therefore, optimizing the pyrolytic temperature is an effective method to regulate the surface hydrophilic/hydrophobicity and electron transfer capability of the designed biochar, targeting the fast and efficient transformation of specific contaminants.

Natural humic substances (HS), derived from dead biomass underground through thousands of years, is another bio-resourced RM usually containing redox functional groups (carboxylic or phenolic groups) that facilitate the electron transfer process (Guo et al. 2018). Natural HS is vastly present in the soil, sedimentation, and natural water environment (Yang et al. 2021), implying an inexpensive but promising method for in-situ remediation and elimination of biorefractory pollutants (Table 5).

Table 5. Bio-resourced Derivatives Acting as Redox Mediator

Moreover, HS has even been reported to accelerate the acidification step during the fermentation process and to decrease (or even halt) the activities of acetotrophic methanogens by inhibiting the key bio-reaction of acetyl-CoA → 5-methyl-THMPT (Liu et al. 2015), preventing the electron sink to methane (CH4). Thus, the presence of HS in anaerobic condition potentially mitigates the release of greenhouse gases, such as CH4, contributing to the carbon neutralization society (Zhang et al. 2020). However, as a typical NOM, HS is a precursor of carcinogenic disinfection byproducts, limiting its application in aquatic systems (Krasner et al. 2009). This provides motivation for the immobilization of them on insoluble carriers, which has been mentioned earlier (Table 2). Moreover, insoluble humic-mineral complexes have also been demonstrated to be effective solid-phase RMs that fulfill the electron-mediating function during the reductive transformation of poly-halogenated contaminants (Zhang et al. 2014a; Cruz-Zavala et al. 2016).

CONCLUSIONS AND FUTURE PERSPECTIVES

The immobilization of soluble redox mediators (RMs) on inert carriers is a promising environmental technology to promote biorefractory pollutant removal and bioelectricity generation by speeding up the electron transfer. The RMs have been found to enable the reuse of redox mediators in scaled-up engineering applications and in-situ remediation. This article reviewed the currently reported RM-immobilizing technologies such as entrapment, adsorption, and surface modification. Entrapment entails the physical fixation of soluble RMs into a polymeric matrix; adsorption concentrates soluble RMs into special solid materials; and surface modification involves chemical, thermal, and electro-chemical modification of specific materials to enhance functional groups that are responsible for electron shuttling. Furthermore, renewable natural and biomass-derived RMs, such as biochar, containing functional groups that transfer electrons, were also included in this review.

However, the immobilization procedure of RMs is still complex and expensive, requiring the process optimization and techno-economic analysis when using specific immobilizing methods and carriers. Although the use of natural RMs or some biomass derivatives, such as biochar, could reduce the operational costs, the densities of functional groups are at a low level. Moreover, properties of natural RM or biochar are varied from different areas, requiring extensive experiments to select suitable ones and their optimal preparation conditions for the effective transformation of specific contaminants. At present, data-driven approaches, such as machine learning algorithms, have been widely applied to assist research results towards engineering applications. In future work, to reduce the labor-consuming and repeated experiments, powerful machine learning algorithms are suggested to aid the design/selection of RM immobilization procedures, feasible carriers, and optimal operating parameters.

ACKNOWLEDGMENTS

The authors are grateful for the support of Postdoctoral Advance Programs of Zhejiang Province (ZJ2021048), and the project of PowerChina Huadong Engineering Corporation Limited (KY2021-GH-02-03).

 

REFERENCES CITED

Adachi, M., Shimomura, T., Komatsu, M., Yakuwa, H., and Miya, A. (2008). “A novel mediator-polymer-modified anode for microbial fuel cells,” Chem. Commun. 2008(17), 2055-2057. DOI: 10.1039/b717773a

Almeida, P. J., Borrego, L., Pulido-Melian, E., and Gonzalez-Diaz, O. (2012). “Quantification of p-phenylenediamine and 2-hydroxy-1,4-naphthoquinone in henna tattoos,” Contact Dermatitis 66(1), 33-37. DOI: 10.1111/j.1600-0536.2011.01992.x

Alvarez, L. H., Perez-Cruz, M. A., Rangel-Mendez, J. R., and Cervantes, F. J. (2010). “Immobilized redox mediator on metal-oxides nanoparticles and its catalytic effect in a reductive decolorization process,” J. Hazard. Mater. 184(1-3), 268-272. DOI: 10.1016/j.jhazmat.2010.08.032

Alvarez, L. H., Jimenez-Bermudez, L., Hernandez-Montoya, V., and Cervantes, F. J. (2012). “Enhanced dechlorination of carbon tetrachloride by immobilized fulvic acids on alumina particles,” Water Air Soil Pollut. 223(4), 1911-1920. DOI: 10.1007/s11270-011-0994-3

Amezquita-Garcia, H. J., Rangel-Mendez, J. R., Cervantes, F. J., and Razo-Flores, E. (2016). “Activated carbon fibers with redox-active functionalities improves the continuous anaerobic biotransformation of 4-nitrophenol,” Chem. Eng. J. 286(15), 208-215. DOI: 10.1016/j.cej.2015.10.085

Anson, C. W., and Stahl, S. S. (2020). “Mediated fuel cells: soluble redox mediators and their applications to electrochemical reduction of O2 and oxidation of H2, alcohols, biomass, and complex fuels,” Chem. Rev. 120(8), 3749-3786. DOI: 10.1021/acs.chemrev.9b00717

Anwer, A. H., Khan, N., Khan, M. D., Shakeel, S., and Khan, M. Z. (2021). “Redox mediators as cathode catalyst to boost the microbial electro-synthesis of biofuel product from carbon dioxide,” Fuel 302(15), article ID 121124. DOI: 10.1016/j.fuel.2021.121124

Bhushan, B., Halasz, A., and Hawari, J. (2006). “Effect of iron (III), humic acids and anthraquinone-2,6-disulfonate on biodegradation of cyclic nitramines by Clostridium sp. EDB2,” J. Appl. Microbiol. 100(3), 555-563. DOI: 10.1111/j.1365-2672.2005.02819.x

Bibi, I., Javed, S., Ata, S., Majid, F., Kamal, S., Sultan, M., Jilani, K., Umair, M., Khan, M. I., and Iqbal, M. (2019). “Biodegradation of synthetic orange G dye by Pleurotus sojar-caju with Punica granatum peal as natural mediator,” Biocatal. Agric. Biotechnol. 22, article ID 101420. DOI: 10.1016/j.bcab.2019.101420

Borch, T., Inskeep, W. P., Harwood, J. A., and Gerlach, R. (2005). “Impact of ferrihydrite and anthraquinone-2,6-disulfonate on the reductive transformation of 2,4,6-trinitrotoluene by a Gram-positive fermenting bacterium,” Environ. Sci. Technol. 39(18), 7126-7133. DOI: 10.1021/es0504441

Buckel, W., and Thauer, R. K. (2018). “Flavin-based electron bifurcation, a new mechanism of biological energy coupling,” Chem. Rev. 118(7), 3862-3886. DOI: 10.1021/acs.chemrev.7b00707

Cervantes, F. J., Garcia-Espinosa, A., Antonieta Moreno-Reynosa, M., and Rangel-Mendez, J. R. (2010). “Immobilized redox mediators on anion exchange resins and their role on the reductive decolorization of azo dyes,” Environ. Sci. Technol. 44(5), 1747-1753. DOI: 10.1021/es9027919

Cervantes, F. J., Gonzalez-Estrella, J., Marquez, A., Alvarez, L. H., and Arriaga, S. (2011). “Immobilized humic substances on an anion exchange resin and their role on the redox biotransformation of contaminants,” Bioresource Technol. 102(2), 2097-2100. DOI: 10.1016/j.biortech.2010.08.021

Cervantes, F. J., Martinez, C. M., Gonzalez-Estrella, J., Marquez, A., and Arriaga, S. (2013). “Kinetics during the redox biotransformation of pollutants mediated by immobilized and soluble humic acids,” Appl. Microbiol. Biotechnol. 97(6), 2671-2679. DOI: 10.1007/s00253-012-4081-5

Chen, W., Meng, J., Han, X., Lan, Y., and Zhang, W. (2019). “Past, present, and future of biochar,” Biochar 1(1), 75-87. DOI: 10.1007/s42773-019-00008-3

Chen, T., Xiao, W., Wang, Z., Xie, T., Yi, C., and Xu, Z. (2022). “Design and engineering of heterogeneous nitroxide-mediated catalytic systems for selective oxidation: Efficiency and sustainability,” Mater. Today Chem. 24, article 100872. DOI: 10.1016/j.mtchem.2022.100872

Chiorcea-Paquim, A. M., Pauliukaite, R., Brett, C. M. A., and Oliveira-Brett, A. M. (2008). “AFM nanometer surface morphological study of in situ electropolymerized neutral red redox mediator oxysilane sol-gel encapsulated glucose oxidase electrochemical biosensors,” Biosens. Bioelectron. 24(2), 297-305. DOI: 10.1016/j.bios.2008.04.001

Ciriminna, R., Blum, J., Avnir, D., and Pagliaro, M. (2000). “Sol-gel entrapped TEMPO for the selective oxidation of methyl α-D-glucopyranoside,” Chem. Commun. (15), 1441-1442. DOI: 10.1039/B003096L

Ciriminna, R., Pandarus, V., Béland, F., and Pagliaro, M. (2018). “Sol–gel entrapped nitroxyl radicals: Catalysts of broad scope,” Chem. Cat. Chem. 10(8), 1731-1738. DOI: 10.1002/cctc.201701910

Colunga, A., Rene Rangel-Mendez, J., Celis, L. B., and Cervantes, F. J. (2015). “Graphene oxide as electron shuttle for increased redox conversion of contaminants under methanogenic and sulfate-reducing conditions,” Bioresource Technol. 175, 309-314. DOI: 10.1016/j.biortech.2014.10.101

Costa, M. C., Mota, S., Nascimento, R. F., and Dos Santos, A. B. (2010). “Anthraquinone-2,6-disulfonate (AQDS) as a catalyst to enhance the reductive decolourisation of the azo dyes Reactive Red 2 and Congo Red under anaerobic conditions,” Bioresource Technol. 101(1), 105-110. DOI: 10.1016/j.biortech.2009.08.015

Cruz-Zavala, A. S., Pat-Espadas, A. M., Rene Rangel-Mendez, J., Chazaro-Ruiz, L. F., Ascacio-Valdes, J. A., Aguilar, C. N., and Cervantes, F. J. (2016). “Immobilization of metal-humic acid complexes in anaerobic granular sludge for their application as solid-phase redox mediators in the biotransformation of iopromide in UASB reactors,” Bioresource Technol. 207, 39-45. DOI: 10.1016/j.biortech.2016.01.125

Dai, R., Chen, X., Ma, C., Xiang, X., and Li, G. (2016). “Insoluble/immobilized redox mediators for catalyzing anaerobic bio-reduction of contaminants,” Rev. Environ. Sci. Bio/Technol. 15(3), 379-409. DOI: 10.1007/s11157-016-9404-z

Del Ángel, Y. A., García-Reyes, R. B., Celis, L. B., Serrano-Palacios, D., Gortáres, P., and Alvarez, L. H. (2021). “Quinone-reducing enrichment culture enhanced the direct and mediated biotransformation of azo dye with soluble and immobilized redox mediator,” J. Water Process Eng. 44, article 102424. DOI: 10.1016/j.jwpe.2021.102424

Di, L., and Hua, Z. (2011). “Porous silica beads supported TEMPO and adsorbed NOx (PSB‐TEMPO/NOx): An efficient heterogeneous catalytic system for the oxidation of alcohols under mild conditions,” Adv. Synth. Catal. 353(8), 1253-1259. DOI: 10.1002/adsc.201000876

dos Santos, A. B., Bisschops, I. A. E., Cervantes, F. J., and van Lier, J. B. (2005). “The transformation and toxicity of anthraquinone dyes during thermophilic (55 °C) and mesophilic (30 °C) anaerobic treatments,” J. Biotechnol. 115(4), 345-353. DOI: 10.1016/j.jbiotec.2004.09.007

Fang, Z., Gao, Y., Bolan, N., Shaheen, S. M., Xu, S., Wu, X., Xu, X., Hu, H., Lin, J., Zhang, F., et al. (2020). “Conversion of biological solid waste to graphene-containing biochar for water remediation: A critical review,” Chem. Eng. J. 390(15), article ID 124611. DOI: 10.1016/j.cej.2020.124611

Feng, C., Ma, L., Li, F., Mai, H., Lang, X., and Fan, S. (2010). “A polypyrrole/ anthraquinone-2,6-disulphonic disodium salt (PPy/AQDS)-modified anode to improve performance of microbial fuel cells,” Biosens. Bioelectron. 25(6), 1516-1520. DOI: 10.1016/j.bios.2009.10.009

Feng, E., Peng, H., Zhang, Z., Li, J., and Lei, Z. (2017). “Polyaniline-based carbon nanospheres and redox mediator doped robust gel films lead to high performance foldable solid-state supercapacitors,” New J. Chem. 41(17), 9024-9032. DOI: 10.1039/c7nj01478c

Fernandez-Fernandez, M., Angeles Sanroman, M., and Moldes, D. (2013). “Recent developments and applications of immobilized laccase,” Biotechnol. Adv. 31(8), 1808-1825. DOI: 10.1016/j.biotechadv.2012.02.013

Gao, Z., Yi, Y., Zhao, J., Xia, Y., Jiang, M., Cao, F., Zhou, H., Wei, P., Jia, H., and Yong, X. (2018). “Co-immobilization of laccase and TEMPO onto amino-functionalized magnetic Fe3O4 nanoparticles and its application in acid fuchsin decolorization,” Bioresour. Bioprocess. 5(1), 1-8. DOI: 10.1186/s40643-018-0215-7

Gharehkhani, S., Zhang, Y., and Fatehi, P. (2019). “Lignin-derived platform molecules through TEMPO catalytic oxidation strategies,” Prog. Energ. Combust. 72, 59-89. DOI: 10.1016/j.pecs.2019.01.002

Guo, J., Zhou, J., Wang, D., Tian, C., Wang, P., Uddin, M. S., and Yu, H. (2007). “Biocatalyst effects of immobilized anthraquinone on the anaerobic reduction of azo dyes by the salt-tolerant bacteria,” Water Res. 41(2), 426-432. DOI: 10.1016/j.watres.2006.10.022

Guo, J., Kang, L., Yang, J., Wang, X., Lian, J., Li, H., Guo, Y., and Wang, Y. (2010). “Study on a novel non-dissolved redox mediator catalyzing biological denitrification (RMBDN) technology,” Bioresource Technol. 101(11), 4238-4241. DOI: 10.1016/j.biortech.2010.01.029

Guo, P., Zhang, C., Wang, Y., Yu, X., Zhang, Z., and Zhang, D. (2018). “Effect of long-term fertilization on humic redox mediators in multiple microbial redox reactions,” Environ. Pollut. 234, 107-114. DOI: 10.1016/j.envpol.2017.10.106

Hu, H., Xu, H., Wu, J., Li, L., Yue, F., Huang, L., Chen, L., Zhang, X., and Ouyang, X. (2020). “Secondary bonds modifying conjugate-blocked linkages of biomass-derived lignin to form electron transfer 3D networks for efficiency exceeding 16% nonfullerene organic solar cells,” Adv. Funct. Mater. 30(23), article ID 2001494. DOI: 10.1002/adfm.202001494

Huang, J., Chu, S., Chen, J., Chen, Y., and Xie, Z. (2014). “Enhanced reduction of an azo dye using henna plant biomass as a solid-phase electron donor, carbon source, and redox mediator,” Bioresource Technol. 161, 465-468. DOI: 10.1016/j.biortech.2014.03.143

Huang, J., Wu, M., Chen, J., Liu, X., Chen, T., Wen, Y., Tang, J., and Xie, Z. (2015). “Enhanced azo dye removal in a continuously operated up-flow anaerobic filter packed with henna plant biomass,” J. Hazard. Mater. 299(15), 158-164. DOI: 10.1016/j.jhazmat.2015.05.044

Huang, J., Wu, M., Tang, J., Zhou, R., Chen, J., Han, W., and Xie, Z. (2016a). “Enhanced bio-reduction of hexavalent chromium by an anaerobic consortium using henna plant biomass as electron donor and redox mediator,” Desalination Water Treat. 57(32), 15125-15132. DOI: 10.1080/19443994.2015.1070292

Huang, J., Zhou, R., Chen, J., Han, W., Chen, Y., Wen, Y., and Tang, J. (2016b). “Volatile fatty acids produced by co-fermentation of waste activated sludge and henna plant biomass,” Bioresource Technol. 211, 80-86. DOI: 10.1016/j.biortech.2016.03.071

Huang, W., Chen, J., Hu, Y., Chen, J., Sun, J., and Zhang, L. (2017). “Enhanced simultaneous decolorization of azo dye and electricity generation in microbial fuel cell (MFC) with redox mediator modified anode,” Int. J. Hydrogen Energy 42(4), 2349-2359. DOI: 10.1016/j.ijhydene.2016.09.216

Huang, J., Chen, S., Wu, W., Chen, H., Guo, K., Tang, J., and Li, J. (2019). “Insights into redox mediator supplementation on enhanced volatile fatty acids production from waste activated sludge,” Environ. Sci. Pollut. Res. 26(26), 27052-27062. DOI: 10.1007/s11356-019-05927-z

Huang, J., Guo, K., Shi, B., Han, W., Hou, P., and Tang, J. (2020). “Achieving high volatile fatty acid production from raw henna (Lawsonia inermis) biomass at mild alkaline conditions,” BioResources 15(2), 3707-3716. DOI: 10.15376/biores.15.2.3707-3716

Huang, J., Shi, B., Han, W., Qiu, S., Li, H., Hou, P., Wu, W., and Tang, J. (2021). “Effect of pH on hexavalent chromium removal driven by henna (Lawsonia inermis) fermentation,” Biochem. Eng. J. 167, Article ID 107919. DOI: 10.1016/j.bej.2020.107919

Krasner, S. W., Westerhoff, P., Chen, B., Rittmann, B. E., and Amy, G. (2009). “Occurrence of disinfection byproducts in United States wastewater treatment plant effluents,” Environ. Sci. Technol. 43(21), 8320-8325. DOI: 10.1021/es901611m

Kwon, M. J., and Finneran, K. T. (2006). “Microbially mediated biodegradation of hexahydro-1,3,5-trinitro-1,3,5-triazine by extracellular electron shuttling compounds,” Appl. Environ. Microbiol. 72(9), 5933-5941. DOI: 10.1128/aem.00660-06

Kwon, M. J., and Finneran, K. T. (2008). “Biotransformation products and mineralization potential for hexahydro-1,3,5-trinitro-1,3,5-triazine (RDX) in abiotic versus biological degradation pathways with anthraquinone-2,6-disulfonate (AQDS) and Geobacter metallireducens,” Biodegradation 19(5), 705-715. DOI: 10.1007/s10532-008-9175-5

Lang, X., Wan, Q., Feng, C., Yue, X., Xu, W., Li, J., and Fan, S. (2010). “The role of anthraquinone sulfonate dopants in promoting performance of polypyrrole composites as pseudo-capacitive electrode materials,” Synth. Met. 160(15-16), 1800-1804. DOI: 10.1016/j.synthmet.2010.06.023

Li, L., Wang, J., Zhou, J., Yang, F., Jin, C., Qu, Y., Li, A., and Zhang, L. (2008). “Enhancement of nitroaromatic compounds anaerobic biotransformation using a novel immobilized redox mediator prepared by electropolymerization,” Bioresource Technol. 99(15), 6908-6916. DOI: 10.1016/j.biortech.2008.01.037

Li, L., Zhou, J., Wang, J., Yang, F., Jin, C., and Zhang, G. (2009). “Anaerobic biotransformation of azo dye using polypyrrole/anthraquinonedisulphonate modified active carbon felt as a novel immobilized redox mediator,” Sep. Purif. Technol. 66(2), 375-382. DOI: 10.1016/j.seppur.2008.12.019

Li, H., Guo, J., Lian, J., Xi, Z., Zhao, L., Liu, X., Zhang, C., and Yang, J. (2014). “Study the biocatalyzing effect and mechanism of cellulose acetate immobilized redox mediators technology (CE-RM) on nitrite denitrification,” Biodegradation 25(3), 395-404. DOI: 10.1007/s10532-013-9668-8

Li, E., Wang, N., He, H., and Chen, H. (2016a). “Improved thermoelectric performances of SrTiO3 ceramic doped with Nb by surface modification of nanosized titania,” Nanoscale Res. Lett. 11, article 188. DOI: 10.1186/s11671-016-1407-8

Li, L., Liu, Q., Wang, Y., Zhao, H., He, C., Yang, H., Gong, L., Mu, Y., and Yu, H. (2016b). “Facilitated biological reduction of nitroaromatic compounds by reduced graphene oxide and the role of its surface characteristics,” Sci. Rep. 6, article ID 30082. DOI: 10.1038/srep30082

Li, X., Guo, W., Liu, Z., Wang, R., and Liu, H. (2017). “Quinone-modified NH2-MIL-101(Fe) composite as a redox mediator for improved degradation of bisphenol A,” J. Hazard. Mater. 324, 665-672. DOI: 10.1016/j.jhazmat.2016.11.040

Li, J., Li, Y., Chen, P., Sathishkumar, K., Lu, Y., Naraginti, S., Wu, Y., and Wu, H. (2022a). “Biological mediated synthesis of reduced graphene oxide (rGO) as a potential electron shuttle for facilitated biological denitrification: Insight into the electron transfer process,” J. Environ. Chem. Eng. 10(5), article ID 108225. DOI: 10.1016/j.jece.2022.108225

Li, Q., Jia, Z., Fu, J., Yang, X., Shi, X., and Chen, R. (2022b). “Biochar enhances partial denitrification/anammox by sustaining high rates of nitrate to nitrite reduction,” Bioresource Technol. 349, article ID 126869. DOI: 10.1016/j.biortech.2022.126869

Lian, J., Hu, Z., Li, Z., Guo, J., Xu, Z., Guo, Y., Li, M., and Yang, J. (2016). “Effects of non-dissolved redox mediators on a hexavalent chromium bioreduction process,” Biotechnol. Biotechnol. Equip. 30(2), 292-298. DOI: 10.1080/13102818.2015.1134277

Liu, G., Zhou, J., Wang, J., Zhou, M., Lu, H., and Jin, R. (2009). “Acceleration of azo dye decolorization by using quinone reductase activity of azoreductase and quinone redox mediator,” Bioresource Technol. 100(11), 2791-2795. DOI: 10.1016/j.biortech.2008.12.040

Liu, H., Guo, J., Qu, J., Lian, J., Guo, Y., Jefferson, W., and Yang, J. (2012a). “Biological catalyzed denitrification by a functional electropolymerization biocarrier modified by redox mediator,” Bioresource Technol. 107, 144-150. DOI: 10.1016/j.biortech.2011.12.071

Liu, H., Guo, J., Qu, J., Lian, J., Jefferson, W., Yang, J., and Li, H. (2012b). “Catalyzing denitrification of Paracoccus versutus by immobilized 1,5-dichloroanthraquinone,” Biodegradation 23(3), 399-405. DOI: 10.1007/s10532-011-9518-5

Liu, K., Chen, Y., Xiao, N., Zheng, X., and Li, M. (2015). “Effect of humic acids with different characteristics on fermentative short-chain fatty acids production from waste activated sludge,” Environ. Sci. Technol. 49(8), 4929-4936. DOI: 10.1021/acs.est.5b00200

Liu, J., Huang, J., Shi, B., Guo K., Li, J., and Tang, J. (2021). “Riboflavin enhanced denitrification of artificial wastewater under low C/N condition in cold season,” BioResources 16(1), 1949-1957. DOI: 10.15376/biores.16.1.1949-1957

Lou, X., Zhi, F., Sun, X., Wang, F., Hou, X., Lv, C., and Hu, Q. (2023). “Construction of co-immobilized laccase and mediator based on MOFs membrane for enhancing organic pollutants removal,” Chem. Eng. J. 451, article ID 138080. DOI: 10.1016/j.cej.2022.138080

Lu, H., Zhou, J., Wang, J., Si, W., Teng, H., and Liu, G. (2010). “Enhanced biodecolorization of azo dyes by anthraquinone-2-sulfonate immobilized covalently in polyurethane foam,” Bioresource Technol. 101(18), 7185-7188. DOI: 10.1016/j.biortech.2010.04.007

Lu, C., Xie, Z., Guo, J., Song, Y., Xing, Y., Han, Y., Li, H., and Hou, Y. (2020). “Chlorophyll as natural redox mediators for the denitrification process,” Int. Biodeterior. Biodegrad. 148, article ID 104895. DOI: 10.1016/j.ibiod.2020.104895

Lukyanov, D. A., Kalnin, A. Y., Rubicheva, L. G., Potapenkov, V. V., Bakulina, O. Y., and Levin, O. V. (2022). “Application of a TEMPO-polypyrrole polymer for NOx-mediated oxygen electroreduction,” Catalysts 12(11), article 1466. DOI: 10.3390/catal12111466

Ma, J., Wang, Z., Zhang, J., Waite, T. D., and Wu, Z. (2017). “Cost-effective Chlorella biomass production from dilute wastewater using a novel photosynthetic microbial fuel cell (PMFC),” Water Res. 108, 356-364. DOI: 10.1016/j.watres.2016.11.016

Mani, P., Kumar, V. T. F., Keshavarz, T., Chandra, T. S., and Kyazze, G. (2018). “The role of natural laccase redox mediators in simultaneous dye decolorization and power production in microbial fuel cells,” Energies 11(12), article 3455. DOI: 10.3390/en11123455

Martinez, C. M., Celis, L. B., and Cervantes, F. J. (2013). “Immobilized humic substances as redox mediator for the simultaneous removal of phenol and Reactive Red 2 in a UASB reactor,” Appl. Microbiol. Biotechnol. 97(22), 9897-9905. DOI: 10.1007/s00253-013-5190-5

Martinez, C. M., Zhu, X., and Logan, B. E. (2017). “AQDS immobilized solid-phase redox mediators and their role during bioelectricity generation and RR2 decolorization in air-cathode single-chamber microbial fuel cells,” Bioelectrochemistry 118, 123-130. DOI: 10.1016/j.bioelechem.2017.07.007

Martins, L. R., Lobo Baeta, B. E., Alves Gurgel, L. V., de Aquino, S. F., and Gil, L. F. (2015). “Application of cellulose-immobilized riboflavin as a redox mediator for anaerobic degradation of a model azo dye Remazol Golden Yellow RNL,” Ind. Crop. Prod. 65, 454-462. DOI: 10.1016/j.indcrop.2014.10.059

Moyo, M., Okonkwo, J. O., and Agyei, N. M. (2012). “Recent advances in polymeric materials used as electron mediators and immobilizing matrices in developing enzyme electrodes,” Sensors 12(1), 923-953. DOI: 10.3390/s120100923

O’Loughlin, E. J. (2008). “Effects of electron transfer mediators on the bioreduction of lepidocrocite (γ-FeOOH) by Shewanella putrefaciens CN32,” Environ. Sci. Technol. 42(18), 6876-6882. DOI: 10.1021/es800686d

Olivo-Alanis, D., Bernardo Garcia-Reyes, R., Alvarez, L. H., and Garcia-Gonzalez, A. (2018). “Mechanism of anaerobic bio-reduction of azo dye assisted with lawsone-immobilized activated carbon,” J. Hazard. Mater. 347, 423-430. DOI: 10.1016/j.jhazmat.2018.01.019

Oliveira, F. R., Patel, A. K., Jaisi, D. P., Adhikari, S., Lu, H., and Khanal, S. K. (2017). “Environmental application of biochar: Current status and perspectives,” Bioresource Technol. 246, 110-122. DOI: 10.1016/j.biortech.2017.08.122

Palmisano, G., Ciriminna, R., and Pagliaro, M. (2006). “Waste-free electrochemical oxidation of alcohols in water,” Adv. Synth. Catal. 348(15), 2033-2037. DOI:10.1002/adsc.200606199

Pelton, R., Ren, P., Liu, J., and Mijolovic, D. (2011). “Polyvinylamine-graft-TEMPO adsorbs onto, oxidizes, and covalently bonds to wet cellulose,” Biomacromolecules 12(4), 942-948. DOI: 10.1021/bm200101b

Pereira, L., Pereira, R., Pereira, M. F. R., van der Zee, F. P., Cervantes, F. J., and Alves, M. M. (2010). “Thermal modification of activated carbon surface chemistry improves its capacity as redox mediator for azo dye reduction,” J. Hazard. Mater. 183(1-3), 931-939. DOI: 10.1016/j.jhazmat.2010.08.005

Pereira, R. A., Pereira, M. F. R., Alves, M. M., and Pereira, L. (2014). “Carbon based materials as novel redox mediators for dye wastewater biodegradation,” Appl. Catal. B 144, 713-720. DOI: 10.1016/j.apcatb.2013.07.009

Rau, J., Knackmuss, H. J., and Stolz, A. (2002). “Effects of different quinoid redox mediators on the anaerobic reduction of azo dyes by bacteria,” Environ. Sci. Technol. 36(7), 1497-1504. DOI: 10.1021/es010227

Ren, Z., Wang, Z., Lv, L., Ma, P., Zhang, G., Li, Y., Qin, Y., Wang, P., Liu, X., and Gao, W. (2022). “Fe-N complex biochar as a superior partner of sodium sulfide for methyl orange decolorization by combination of adsorption and reduction,” J. Environ. Manag. 316, article ID 115213. DOI: 10.1016/j.jenvman.2022.11521

Rocha‐Martín, J., Betancor, L., and López‐Gallego, F. (2021). “Immobilization techniques for the preparation of supported biocatalysts: Making better biocatalysts through protein immobilization,” Biocatalysis for Practitioners: Techniques, Reactions and Applications, 63-88. DOI: 10.1002/9783527824465.ch3

Romero-Soto, I. C., Martinez-Perez, R. B., Rodriguez, J. A., Camacho-Ruiz, R. M., Barbachano-Torres, A., Martin del Campo, M., Napoles-Armenta, J., Pliego-Sandoval, J. E., Concha-Guzman, M. O., and Angeles Camacho-Ruiz, M. (2021). “Galactomannans for entrapment of Gliomastix murorum laccase and their use in reactive Blue 2 decolorization,” Sustainability 13(16), article 9019. DOI: 10.3390/su13169019

Sakurada, Y., Takeda, K., Ohno, H., and Nakamura, N. (2017). “Immobilization of pyrroloquinoline quinone-dependent alcohol dehydrogenase with a polyion complex and redox polymer for a bioanode,” Catalysts 7(10), article 296. DOI: 10.3390/catal7100296

Saratale, R. G., Saratale, G. D., Chang, J. S., and Govindwar, S. P. (2011). “Bacterial decolorization and degradation of azo dyes: A review,” J. Taiwan Inst. Chem. Eng. 42(1), 138-157. DOI: 10.1016/j.jtice.2010.06.006

Sharma, S. C. D., Sun, Q., Li, J., Wang, Y., Suanon, F., Yang, J., and Yu, C. (2016). “Decolorization of azo dye methyl red by suspended and co-immobilized bacterial cells with mediators anthraquinone-2,6-disulfonate and Fe3O4 nanoparticles,” Int. Biodeterior. Biodegrad. 112, 88-97. DOI: 10.1016/j.ibiod.2016.04.035

Shi, Z., Ding, Y., Zhang, Q., and Sun, J. (2022). “Electrocatalyst modulation toward bidirectional sulfur redox in Li–S batteries: From strategic probing to mechanistic understanding,” Adv. Energy Mater. 12(9), article 2201056. DOI: 10.1002/aenm.202201056

Song, Y., Jiang, J., Qin, W., Li, J., Zhou, Y., and Gao, Y. (2021). “Enhanced transformation of organic pollutants by mild oxidants in the presence of synthetic or natural redox mediators: A review,” Water Res. 189, article ID 116667. DOI: 10.1016/j.watres.2020.116667

Somu, P., Narayanasamy, S., Gomez, L. A., Rajendran, S., Lee, Y. R., and Balakrishnan, D. (2022). “Immobilization of enzymes for bioremediation: A future remedial and mitigating strategy,” Environ. Res. 212, article 113411. DOI: 10.1016/j.envres.2022.113411

Su, Y., Zhang, Y., Wang, J., Zhou, J., Lu, X., and Lu, H. (2009). “Enhanced bio-decolorization of azo dyes by co-immobilized quinone-reducing consortium and anthraquinone,” Bioresource Technol. 100(12), 2982-2987. DOI: 10.1016/j.biortech.2009.01.029

Sun, W., Kong, J., and Deng, J. (1997). “Electrocatalytic reduction of hemoglobin at a chemically modified electrode containing riboflavin,” Electroanalysis 9(2), 115-119. DOI: 10.1002/elan.1140090205

Sun, Q., Yang, J., Fan, Y., Cai, K., Lu, Z., He, Z., Xu, Z., Lai, X., Zheng, Y., Liu, C., Wang, F., and Sun, Z. (2022). “The role of trace N-Oxyl compounds as redox mediator in enhancing antiviral ribavirin elimination in UV/chlorine process,” Appl. Catal. B: Environ. 317, article 121709. DOI: 10.1016/j.apcatb.2022.121709

Tang, J., Wu, M., Huang, J., Chen, J., Han, W., and Xie, Z. (2016). “Effects of different henna plant parts on enhanced removal of an azo dye Orange II: Biotic and abiotic contributions,” Environ. Prog. Sustainable Energy 35(2), 404-410. DOI: 10.1002/ep.12248

Uchimiya, M., and Stone, A. T. (2009). “Reversible redox chemistry of quinones: Impact on biogeochemical cycles,” Chemosphere 77(4), 451-458. DOI: 10.1016/j.chemosphere.2009.07.025

van der Zee, F. R., and Cervantes, F. J. (2009). “Impact and application of electron shuttles on the redox (bio)transformation of contaminants: A review,” Biotechnol. Adv. 27(3), 256-277. DOI: 10.1016/j.biotechadv.2009.01.004

van der Zee, F. P., Bouwman, R. H., Strik, D. P., Lettinga, G., and Field, J. A. (2001). “Application of redox mediators to accelerate the transformation of reactive azo dyes in anaerobic bioreactors,” Biotechnol. Bioeng. 75(6), 691-701. DOI: 10.1002/bit.10073

van der Zee, F. P., Bisschops, I. A. E., Lettinga, G., and Field, J. A. (2003). “Activated carbon as an electron acceptor and redox mediator during the anaerobic biotransformation of azo dyes,” Environ. Sci. Technol. 37(2), 402-408. DOI: 10.1021/es025885o

Viveiros, R., Maia, L. B., Corvo, M. C., Bonifácio, V. D., Heggie, W., and Casimiro, T. (2022). “Enzyme-inspired dry-powder polymeric catalyst for green and fast pharmaceutical manufacturing processes,” Catal. Commun. 172, article 106537. DOI: 10.1016/j.catcom.2022.106537

Wang, J., Li, L., Zhou, J., Lu, H., Lu, G., Jin, R., and Yang, F. (2009). “Enhanced biodecolorization of azo dyes by electropolymerization-immobilized redox mediator,” J. Hazard. Mater. 168(2-3), 1098-1104. DOI: 10.1016/j.jhazmat.2009.02.152

Wang, D. C., Li, Y. H., Li, D., Xia, Y. Z., and Zhang, J. P. (2010). “A review on adsorption refrigeration technology and adsorption deterioration in physical adsorption systems,” Renew. Sust. Energy Rev. 14(1), 344-353. DOI: 10.1016/j.catcom.2022.106537

Wang, K., Liu, Y., and Chen, S. (2011). “Improved microbial electrocatalysis with neutral red immobilized electrode,” J. Power Sources 196(1), 164-168. DOI: 10.1016/j.jpowsour.2010.06.056

Wang, J., Lu, H., Zhou, Y., Song, Y., Liu, G., and Feng, Y. (2013). “Enhanced biotransformation of nitrobenzene by the synergies of Shewanella species and mediator-functionalized polyurethane foam,” J. Hazard. Mater. 252-253, 227-232. DOI: 10.1016/j.jhazmat.2013.02.040

Wang, G., Yang, S., Ding, J., Chen, C., Zhong, L., Ding, L., Ma, M., Sun, G., Huang, Z., and Ren, N. Q. (2021). “Immobilized redox mediators on modified biochar and their role on azo dye biotransformation in anaerobic biological systems: Mechanisms, biodegradation pathway and theoretical calculation,” Chem. Eng. J. 423, article 130300. DOI: 10.1016/j.cej.2021.130300

Wang, W., Wang, T., Liu, Q., Wang, H., Xue, H., Zhang, Z., and Wang, Y. (2022). “Biochar-mediated DNRA pathway of anammox bacteria under varying COD/N ratios,” Water Res. 212, article ID 118100. DOI: 10.1016/j.watres.2022.118100

Wen, Y., Yuan, Z., Liu, X., Qu, J., Yang, S., Wang, A., Wang, C., Wei, B., Xu, J., and Ni, Y. (2019). “Preparation and characterization of lignin-containing cellulose nanofibril from poplar high-yield pulp via TEMPO-mediated oxidation and homogenization,” ACS Sustain. Chem. Eng. 7(6), 6131-6139. DOI: 10.1021/acssuschemeng.8b06355

Xu, H., Quan, X., Xiao, Z., and Chen, L. (2017). “Cathode modification with peptide nanotubes (PNTs) incorporating redox mediators for azo dyes decolorization enhancement in microbial fuel cells,” Int. J. Hydrogen Energy 42(12), 8207-8215. DOI: 10.1016/j.ijhydene.2017.01.025

Xu, Q., Guo, J., Niu, C., Lian, J., Hou, Z., Guo, Y., and Li, S. (2015). “The denitrification characteristics of novel functional biocarriers immobilised by non-dissolved redox mediators,” Biochem. Eng. J. 95, 98-103. DOI: 10.1016/j.bej.2014.12.004

Xu, Z., Xu, X., Zhang, Y., Yu, Y., and Cao, X. (2020). “Pyrolysis-temperature depended electron donating and mediating mechanisms of biochar for Cr(VI) reduction,” J. Hazard. Mater. 388, article ID 121794. DOI: 10.1016/j.jhazmat.2019.121794

Yang, F., Tang, C., and Antonietti, M. (2021). “Natural and artificial humic substances to manage minerals, ions, water, and soil microorganisms,” Chem. Soc. Rev. 50(10), 6221-6239. DOI: 10.1039/d0cs01363c

Yin, C., Aroua, M. K., and Daud, W. M. A. W. (2007). “Review of modifications of activated carbon for enhancing contaminant uptakes from aqueous solutions,” Sep. Purif. Technol. 52(3), 403-415. DOI: 10.1016/j.seppur.2006.06.009

Yin, X., Qiao, S., Yu, C., Tian, T., and Zhou, J. (2015). “Effects of reduced graphene oxide on the activities of anammox biomass and key enzymes,” Chem. Eng. J. 276, 106-112. DOI: 10.1016/j.cej.2015.04.073

Yuan, S., Lu, H., Wang, J., Zhou, J. T., Wang, Y., and Liu, G. (2012). “Enhanced bio-decolorization of azo dyes by quinone-functionalized ceramsites under saline conditions,” Process Biochem. 47(2), 312-318. DOI: 10.1016/j.procbio.2011.11.015

Yuan, Y., Bolan, N., Prevoteau, A., Vithanage, M., Biswas, J. K., Ok, Y. S., and Wang, H. (2017). “Applications of biochar in redox-mediated reactions,” Bioresource Technol. 246, 271-281. DOI: 10.1016/j.biortech.2017.06.15

Yuan, J., Wen, Y., Dionysiou, D. D., Sharma, V. K., and Ma, X. (2022). “Biochar as a novel carbon-negative electron source and mediator: Electron exchange capacity (EEC) and environmentally persistent free radicals (EPFRs): A review,” Chem. Eng. J. 429, article ID 132313. DOI: 10.1016/j.cej.2021.132313

Zhai, S., Cheng, H., Wang, Q., Zhao, Y., Wang, A., and Ji, M. (2022). “Reinforcement of denitrification in a biofilm electrode reactor with immobilized polypyrrole/ anthraquinone-2, 6-disulfonate composite cathode,” J. Environ. Manage. 315, article 115203. DOI: 10.1016/j.jenvman.2022.115203

Zhan, J. L., Wu, M. W., Wei, D., Wei, B. Y., Jiang, Y., Yu, W., and Han, B. (2019). “4-HO-TEMPO-catalyzed redox annulation of cyclopropanols with oxime acetates toward pyridine derivatives,” ACS Catal. 9(5), 4179-4188. DOI: 10.1021/acscatal.9b00832

Zhang, C., Zhang, D., Li, Z., Akatsuka, T., Yang, S., Suzuki, D., and Katayama, A. (2014a). “Insoluble Fe-humic acid complex as a solid-phase electron mediator for microbial reductive dechlorination,” Environ. Sci. Technol. 48(11), 6318-6325. DOI: 10.1021/es501056n

Zhang, G., Yang, F., Gao, M., Fang, X., and Liu, L. (2008). “Electro-Fenton degradation of azo dye using polypyrrole/anthraquinonedisulphonate composite film modified graphite cathode in acidic aqueous solutions,” Electrochim. Acta 53(16), 5155-5161. DOI: 10.1016/j.electacta.2008.01.008

Zhang, H., and Hu, X. (2017). “Catalytic reduction of NACs by nano Fe3O4/quinone composites in the presence of a novel marine exoelectrogenic bacterium under hypersaline conditions,” RSC Advances 7(20), 11852-11861. DOI: 10.1039/c7ra00365j

Zhang, H., Lu, H., Zhang, S., Liu, G., Li, G., Zhou, J., and Wang, J. (2014b). “A novel modification of poly (ethylene terephthalate) fiber using anthraquinone-2-sulfonate for accelerating azo dyes and nitroaromatics removal,” Sep. Purif. Technol. 132, 323-329. DOI: 10.1016/j.seppur.2014.05.042

Zhang, H., Shi, Z., Bai, R., Wang, D., Cui, F., Zhang, J., and Strathmann, T. J. (2021). “Role of TEMPO in enhancing permanganate oxidation toward organic contaminants,” Environ. Sci.Technol. 55(11), 7681-7689. DOI: 10.1021/acs.est.1c01824

Zhang, J., Yang, X., Shi, J., Zhao, M., Yin, W., Wang, X., Wang, S., and Zhang, C. (2022). “Carbon matrix of biochar from biomass modeling components facilitates electron transfer from zero-valent iron to Cr(VI),” Environ. Sci. Pollut. Res. 29(16), 24309-24321. DOI: 10.1007/s11356-021-17713-x

Zhang, Y., Zhang, Z., Liu, W., and Chen, Y. (2020). “New applications of quinone redox mediators: Modifying nature-derived materials for anaerobic biotransformation process,” Sci. Total Environ. 744, article ID 140652. DOI: 10.1016/j.scitotenv.2020.140652

Zhao, N., Yin, Z., Liu, F., Zhang, M., Lv, Y., Hao, Z., Pan, G., and Zhang, J. (2018). “Environmentally persistent free radicals mediated removal of Cr (VI) from highly saline water by corn straw biochars,” Bioresource Technol. 260, 294-301. DOI: 10.1016/j.biortech.2018.03.116

Zhou, W., Chen, X., Ismail, M., Wei, L., and Hu, B. (2021). “Simulating the synergy of electron donors and different redox mediators on the anaerobic decolorization of azo dyes: Can AQDS-chitosan globules replace the traditional redox mediators?” Chemosphere, 275, article 130025. DOI: 10.1016/j.chemosphere.2021.130025

Zor, E., Oztekin, Y., Ramanaviciene, A., Anusevicius, Z., Voronovic, J., Bingol, H., Barauskas-Memenas, D., Labanauskas, L., and Ramanavicius, A. (2016). “Evaluation of 1,10-phenanthroline-5,6-dione as redox mediator for glucose oxidase,” J. Anal. Chem. 71(1), 77-81. DOI: 10.1134/s1061934816010044

Zou, Y., Xiang, C., Yang, L., Sun, L. X., Xu, F., and Cao, Z. (2008). “A mediatorless microbial fuel cell using polypyrrole coated carbon nanotubes composite as anode material,” Int. J. Hydrogen Energy 33(18), 4856-4862. DOI: 10.1016/j.ijhydene.2008.06.061

Article submitted: September 25, 2022; Peer review completed: November 18, 2022; Revised version received and accepted: December 10, 2022; Published: December 14, 2022.

DOI: 10.15376/biores.18.1.Guo