Next Article in Journal
Social and Economic Aspects of Water Use in Specialty Crop Production in the USA: A Review
Next Article in Special Issue
Reductive/Oxidative Sequential Bioelectrochemical Process for Perchloroethylene Removal
Previous Article in Journal
An Improved Simplified Urban Storm Inundation Model Based on Urban Terrain and Catchment Modification
Previous Article in Special Issue
Bioelectrochemical Systems for Groundwater Remediation: The Development Trend and Research Front Revealed by Bibliometric Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Progress Towards Bioelectrochemical Remediation of Hexavalent Chromium

1
Department of Civil and Environmental Engineering, Politecnico di Milano, Piazza Leonardo da Vinci 32, 20133 Milano, Italy
2
Department of Earth and Environmental Sciences, University of Milano-Bicocca, Piazza della Scienza 1, 20126 Milano, Italy
3
Department of Agriculture, Food, Environment and Forestry, University of Florence, Piazzale delle Cascine 18, 50144 Firenze, Italy
*
Authors to whom correspondence should be addressed.
Water 2019, 11(11), 2336; https://doi.org/10.3390/w11112336
Submission received: 1 October 2019 / Revised: 2 November 2019 / Accepted: 3 November 2019 / Published: 7 November 2019

Abstract

:
Chromium is one of the most frequently used metal contaminants. Its hexavalent form Cr(VI), which is exploited in many industrial activities, is highly toxic, is water-soluble in the full pH range, and is a major threat to groundwater resources. Alongside traditional approaches to Cr(VI) treatment based on physical-chemical methods, technologies exploiting the ability of several microorganisms to reduce toxic and mobile Cr(VI) to the less toxic and stable Cr(III) form have been developed to improve the cost-effectiveness and sustainability of remediating hexavalent chromium-contaminated groundwater. Bioelectrochemical systems (BESs), principally investigated for wastewater treatment, may represent an innovative option for groundwater remediation. By using electrodes as virtually inexhaustible electron donors and acceptors to promote microbial oxidation-reduction reactions, in in situ remediation, BESs may offer the advantage of limited energy and chemicals requirements in comparison to other bioremediation technologies, which rely on external supplies of limiting inorganic nutrients and electron acceptors or donors to ensure proper conditions for microbial activity. Electron transfer is continuously promoted/controlled in terms of current or voltage application between the electrodes, close to which electrochemically active microorganisms are located. Therefore, this enhances the options of process real-time monitoring and control, which are often limited in in situ treatment schemes. This paper reviews research with BESs for treating chromium-contaminated wastewater, by focusing on the perspectives for Cr(VI) bioelectrochemical remediation and open research issues.

1. Introduction

Hexavalent Cr(VI) and trivalent Cr(III) chromium are the most common forms of this element in the environment. Cr(VI) is water-soluble in the full pH range and extremely toxic to human health and all living organisms because of its mutagenic and carcinogenic properties [1]. The U.S. EPA has classified Cr(VI) as one of the 17 most dangerous elements for human health [2,3]. In aqueous systems, Cr(VI) can be present in different species: primarily as chromic acid [H2CrO4] and its salts, the hydrogen chromate ion [(HCrO4)], and the chromate ion [(CrO4)2−]. The chemical equilibrium of the different chromium species depends on Cr(VI) concentration, oxidation-reduction potential (ORP), and pH of the system [3]. Cr(III) is less toxic and more stable than Cr(VI) in aquatic environments, since under natural conditions, at moderately acidic or alkaline pH, it tends to precipitate as chromium hydroxide or oxide (Cr(OH)3 or Cr2O3) [4,5].
Hexavalent chromium has been, and it is still used, in many industrial activities such as chromite ore processing, electroplating, in the production of dyes and pigments, pharmaceuticals, in leather tanning, in wood preservation/processing, and in the metallurgical industry for alloy preparation [6,7,8]. Improper management of Cr(VI)-containing effluents or wastes led to widespread chromium environmental contamination around the world. As a consequence of its subsequent mobility in water, groundwater resources are especially vulnerable to Cr(VI) contamination, with levels often shown to exceed the internationally acceptable exposure limit in water of 0.05 mg Cr(VI)/L [9,10].
Hexavalent chromium in industrial wastewater is typically treated by reduction, and subsequent precipitation, to the non-toxic Cr(III) by means of reductants (for instance, FeSO4, Na2S2O5, and SO2) [11] or electrochemical processes (including electrocoagulation, electro-reduction, electrodialysis, and electro-deionization) [12]. Generally, these processes are easy to implement and efficient at high or moderate Cr(VI) concentrations. However, the use of reductants produces a large amount of metallic sludge, and both the above-mentioned approaches are either ineffective or not cost-effective when applied for trace Cr(VI) treatment. Adsorption on adsorbents, such as activated carbons or zeolites, is also adopted, since this process exhibits several advantages of simple operation, low cost, and high efficiency [13,14]. Nevertheless, this strategy presents some disadvantages such as needing regeneration of adsorption media and no degradation/detoxification of Cr(VI) to Cr(III) is achieved [15].
Conventional methods for Cr(VI) contaminated site remediation, soil excavation, and groundwater pump and treat (P&T), despite still being widely applied, are energy and chemical-intensive and entail high costs [16]. Furthermore, the P&T method, which involves the extraction of contaminated water from the aquifer and above-ground treatment, tends to be poorly efficient when dealing with contaminated plumes at relatively low Cr(VI) concentrations [17].
In situ chemical strategies are applied using Fe(0) permeable reactive barriers or by injecting nano-Fe(0), Na2S2O4 or hydrogen sulphide to develop reactive zones in the aquifer [7,18,19]. The applicability and effectiveness of these approaches are limited by high costs, poor chemical distributions, and undesired side reactions in the subsurface [20].
Recent studies have shown that microbial activity can indirectly promote Cr(VI) reduction and immobilization in contaminated groundwater and/or soil. Bio-reduction relies on injections of biodegradable organic substrates (e.g., molasses), whose rapid microbial degradation by indigenous heterotrophic microorganisms prompts anaerobic conditions in the subsurface and produces reductants, such as S2−, Fe(II), and fermentation metabolites, which are able to mediate Cr(VI) chemical reduction and precipitation [21,22,23,24].
Selected microorganisms can directly bio-reduce [25,26] and bio-sorb [27] Cr(VI) as well. Nevertheless, the presence of suitable microorganisms in conjunction with favorable environmental conditions is essential. The performance of in situ Cr(VI) bio-remediation, by relying on either induced—or direct—bioreduction/biosorption mechanisms, is, therefore, influenced by the abundance of suitable electron acceptors and donors, carbon sources, and nutrients for the microorganisms, which often require external supplements for a balance [20]. The control of supply rates can be a crucial step to avoid unwanted side reactions and the accumulation of undesired substances (e.g., excess substrates or nutrients, fermentation products) that contribute to the deterioration of soil and water quality.
Cr(VI), like any metal, is not actually removed by in situ remediation, but only changes its form and valence. This requires careful evaluation of the long-term stability of reduced products, post-treatment, and of possible re-oxidation mechanisms [28,29]. Cr(III) in the precipitates formed during remediation, whether exposed to environmentally common Mn oxides, such as the birnessite, have the potential of rapid re-oxidation to Cr(VI). Even highly oxidized Mn(III, IV) oxides are likely unstable under reducing conditions, in case these are caused by the addition of electron donors as part of the Cr(VI) remediation strategy. Cr could be re-mobilized when the donor addition stops and the site returns to natural oxidizing conditions [28].
A promising strategy to be explored for in situ Cr(VI) bioremediation is the application of bioelectrochemical systems (BESs), which is an emerging platform technology combining microbial processes with electrochemical systems. In BESs, the ability of electrochemically active microorganisms (EAM) to use electrodes as inexhaustible electron acceptors/donors are exploited, via a process typically referred to as extracellular electron transfer [30,31,32]. BES reactors essentially consist of electrodes, an anode and a cathode, immersed in an electrolytic medium/solution and an optional ion-exchange membrane to separate the compartments. At the anode, oxidation of reduced species generates a flow of electrons to the cathode, where reduction reactions take place. At least one or both reactions are microbially mediated. Membranes provide a separation structure to isolate different bulk liquids in the anode and cathode chambers, to optimize the operating condition without affecting the microbial community, to prevent undesired substrate transport, and to facilitate transfer of ionic species from one chamber to another for charge balance, increasing, however, the internal resistance of the system [33]. In case of thermodynamically favorable redox reactions, BESs can result in direct electricity production (microbial fuel cells, MFCs) [30,34,35] or, by external energy supply, in enhancement of thermodynamically unfavorable processes [36], with production of less toxic or value-added chemicals (such as hydrogen, H2O2, methane, or even organic molecules) (in microbial electrolysis cells, MECs) [37].
BESs have been extensively studied and intensively developed, especially during the last 10 years, for wastewater treatment, valorization, and reuse [38,39]. In environmental remediation, BESs, through biologically-mediated oxidation (at the anode) and reduction (at the cathode), potentially provide a flexible platform for treating many pollutants frequently found at contaminated sites, in co-contamination cases [40,41]. Solid electrodes can serve, in fact, either as an electron sink, for the oxidation of petroleum hydrocarbons [42,43,44] or As(III) [45], or as electron donor, for reduction of chlorinated hydrocarbons [46,47], nitrate [48], or oxidized metals, including Cr(VI) [49].
In in situ treatments, it would be possible to directly introduce the electrodes in the aquifer and stimulate biologic activity with no external chemicals or a minimal external chemical supply [50], which creates an advantage in terms of cost-effectiveness and sustainability in comparison to current approaches [51]. The electrical signal generated in BESs provides opportunities for real-time monitoring of Cr(VI) concentration [52,53,54] and in situ microbial activity [55]. Cr(III) deposition next to the electrode theoretically offers the chance of recovering the metal itself through the electrode [39].
Several experimental works investigated Cr(VI) reduction in MFCs with bioanodes and either abiotic cathodes, relying on Cr(VI) electrochemical reduction, or biocathodes, while taking advantage of biological activity. However, no research, to our knowledge, has yet specifically addressed Cr(VI) contaminated groundwater remediation. This review provides a comprehensive analysis of the current knowledge and experiences in bio-electrochemical treatment of Cr(VI) contaminated water streams, in order to explore BESs opportunities for in situ groundwater bioremediation.

2. Principles of Cr(VI) Reduction in Bio-Electrochemical Systems

Bio-electrochemical Cr(VI) reduction essentially relies on cathodic reduction (Figure 1), with most research focused on wastewater treatment coupled with energy recovery in MFCs, with biotic anodes and either abiotic or biotic cathodes [56]. Only a single study evaluated Cr(VI) detoxification at the bioanode, via bacterial protection mechanisms [57].

2.1. Electrochemical Reduction of Cr(VI)

Due to its high standard reduction potential, which is comparable, or, in certain conditions, even higher than those of other commonly used electron acceptors in BESs, Cr(VI) has been initially investigated as a theoretically favorable electron acceptor, to get reduced at abiotic cathodes in a typical MFC configuration for power production. This concept has been demonstrated for the first time in a dual-chamber MFC (2CMFC) by Wang et al. [59] who, using acetate as electron donor and Cr(VI) solution at a pH of 2 as acceptor, observed higher power densities than for O2 and hexacyanoferrate.
The half-cell Cr(VI) reduction potential and the stoichiometry of the reaction are, however, strongly dependent on chromium species, concentration, and pH conditions (Table 1).
In water solutions, the dichromate Cr2O72− form prevails for total chromium concentrations above approximately 1 g/L [3,17]. At lower concentrations, which typically occurs in groundwater plumes or natural surface water, the dominant species is HCrO4 at a pH between 1 and about 6 to 6.5, and CrO42− at neutral or alkaline conditions [60]. High positive standard reduction potentials (Eh0 vs. SHE) for both Cr2O72− and HCrO4 indicate a thermodynamically favorable reaction, conducive to high power density generation in BESs, only in acidic environments [39,59,61]. On the contrary, CrO42− lower potential limits chromium electrochemical reduction in the neutral pH range, which often makes the external energy supply necessary. Furthermore, at pH < 4, the predominant form of chromium reduction is dissolved Cr3+, whereas in the 5–8 pH range, soluble Cr(OH)2+ and Cr(OH)2+, coexist with Cr(OH)3 or Cr2O3 precipitates, which are responsible for the progressive deterioration of the reduction rates as Cr(III) deposits onto the cathode surface, especially at a pH above 6.5 [62,63].
Effective abiotic electrochemical reduction requires strongly acidic conditions (optimally at a pH of 2) that greatly limit its applicability to environmental remediation.
Biologically mediated Cr(VI) reduction, relying on several microbiological mechanisms, may overcome the current issues of abiotic electrochemical reduction, by offering opportunities to hexavalent chromium treatment in the environmentally compatible neutral pH range. Moreover, the biofilm on the cathode may somewhat protect and improve the long-term efficiency of the electrode by preventing or delaying Cr(III) deposition [64,65].

2.2. Microbiological Mechanisms of Cr(VI) Reduction

Several mechanisms of bacterial Cr(VI) reduction have been described both under aerobic and anaerobic conditions [66,67].
In the presence of oxygen, the reduction of Cr(VI) is commonly associated with soluble chromate reductases and requires reduced nicotinamide-adenine dinucleotide phosphate (NAD(P)H) as electron donor [66,68]. The mechanisms associated with Cr(VI) reduction can involve a direct, one-step or two-step electron transfer. Escherichia coli YieF Cr(VI) reductase transfers three electrons to Cr(VI) in one step to produce Cr(III), and one to molecular oxygen generating reactive oxygen species (ROS) [69]. The Cr(VI) reductase ChrR from Pseudomonas putida involves a one/two steps mechanism in which one/two electrons are donated from NAD(P)H to generate the intermediate Cr(V)/Cr(IV) that is further reduced to Cr(III) by one/two additional electrons [66,70]. Under aerobic conditions, most Cr(VI)-resistant microorganisms tolerate up to 1500 mg Cr(VI)/L [71]. However, the rate of chromium reduction is directly related to the concentration of the contaminant and physical parameters, such as pH and temperature [71,72]. One of the first studies with Cr(VI)-reducing bacteria, achieved almost 100% of chromate reduction in 2.0 mg/L Cr(VI) solution within 90 h by P. putida PRS2000 and P. fluorescens LB303 [73]. Similar results were obtained by the soil-isolated strains Bacillus sp. E29 and Arthrobacter crystallopoietes strain ES32 that achieved reductions of 82% and 90% of Cr(VI) in less than 6 h and 12 h, respectively [71]. Much higher Cr(VI) concentrations were removed by Serratia proteamaculans. Within 48 h, 100 mg Cr(VI)/L were reduced (corresponding to 100% of dichromate added) under aerobic conditions [72]. In the same study, the authors demonstrated that S. proteamaculans was also able to reduce chromate anaerobically, but the process was more efficient in the presence of oxygen.
Chromium-resistant microorganisms have also been found in marine environments. A Cr(VI)-resistant bacterium isolated from seawater and identified as Exiguobacterium indicum achieved nearly 92%, 50%, and 46% reduction for 100, 500, and 1000 mg Cr(VI)/L, respectively, after 192 h of incubation [74].
Under anaerobic conditions, Cr(VI) can serve as the final electron acceptor in a process that usually involves membrane-bound reductases [66], but also soluble enzymes (e.g., soluble cytochrome c3 from Desulfovibrio vulgaris) were observed to reduce Cr(VI) [75]. The overall reaction is provided in Equation (1), with glucose as electron donor [76].
C6H12O6 + 8CrO42−(aq) + 14H2O → 8Cr(OH)3(s) + 10OH(aq) + 6HCO(aq)
Cr(VI) reduction in anaerobic conditions was reported in several microorganisms. Both P. dechromaticans and in Enterobacter cloacae are capable to use Cr(VI) as terminal electron acceptor [66,77]. Gene expression of Shewanella oneidensis MR-1 during Cr(VI) reduction was studied [78]. Under Cr(VI) reducing conditions, 83 genes were upregulated. Among the others, genes involved in the reduction of Fe(III) and Mn(IV) were also upregulated. Further studies with mutant strains confirmed the involvement of mtrA, mtrB, mtrC, and omcA in the reduction of Cr(VI) [78,79].
In anaerobic environments, iron(II) and sulphide can also play a role in Cr(VI) reduction [66,80]. Iron-reducing bacteria (IRB) reduce Fe(III) to Fe(II), and biologically produced Fe(II) can be re-oxidized by reducing Cr(VI) to Cr(III) [76]. Sulphate is an electron acceptor widely used by several bacterial groups for the degradation of the organic matter in anaerobic environments [81]. In sulphate-rich environments, Cr(VI) can react with sulphide produced by sulphate-reducing bacteria (SRB) to produce Cr(III) that precipitates [76].
In addition to laboratory studies focused on the elucidation of possible mechanisms used by bacteria for Cr(VI) reduction, the ability to reduce Cr(VI) in soil-aquifer systems has been reported but needs to be further investigated. Clostridium chromiireducens sp. a Cr(VI)-resistant, Gram-positive, spore-forming, obligate anaerobe, was identified for its ability to reduce Cr(VI) at a contaminated site [82]. Many previous reports also confirmed autotrophic reduction of chromium, by mostly using hydrogen as electron donor [83,84]. A Gram-negative bacterium, capable of reducing hexavalent chromium, was also isolated from a contaminated site. 16S rRNA analysis revealed that it belonged to the Pseudomonas genus, with high similarity to P. synxantha [85,86]. Marsh and McInerney [2] demonstrated reduction of Cr(VI) with hydrogen and carbon dioxide/NaHCO3 as an electron donor and carbon source, respectively, carried out by an anaerobic mixed culture developed from aquifer sediment.
As previously reported, the reduction of the Cr(VI) can occur through the action of soluble cytochromes [66], membrane reductase mtrCAB, and omcA [78,79]. Kracke and colleagues [87] report that the mtrCAB terminal reductase complex and the omcA cytochrome are able to interact with the electrode directly or via mediators (flavins). This scientific evidence combined with the well documented ability of IRB (i.e., Shewanella oneidensis MR-1) [78] and SRB (i.e., Desulfovibrio desulfuricans 27774) [88] to reduce Cr(VI) and to exchange electrons with solid materials are good reasons to consider the feasibility of a bio-electrochemical system for reducing Cr(VI).
All the previously described biological mechanisms can take place in BESs. The inoculated cathode may act as an electron donor for electrochemical or bio-electrochemical Cr(VI) reduction. Electroactive Cr-reducing microorganisms in the cathodic biofilm or the production of hydrogen at the cathode of a BES could also favor the autotrophic reduction of Cr(VI) by hydrogenotrophic bacteria [2,83,84], or IRB or SRB involved in bio-electrochemical processes may facilitate indirect reduction of Cr(VI).

3. Cr(VI) Biocathodic Reduction

Table 2 summarizes the available research experiences of Cr(VI) reduction with biocathodes, by focusing on the electrode materials and inoculum, cathode potential, pH, Cr(VI) concentrations, the observed Cr(VI) removal rates, and efficiencies.
These experiments evaluated Cr(VI) reduction under anaerobic conditions. So far, no literature report a reduction of Cr(VI) in open-air biocathodes. Oxygen is unquestionably the preferred final electron acceptor for microorganisms. However, possible advantages to Cr(VI)-reducing bacteria with an aerobic cathode may exist. Cr(VI)-reducing bacteria may be favored over other species in an environment with specific toxicity, even though tolerance to high Cr(VI) concentrations, up to 1 g Cr(VI)/L or even above, have often been documented [103].
The first test with Cr(VI) reducing biocathode was performed by Tandukar et al. [89], who inoculated the cathodic compartment of a 2CMFC (PEM membrane) with a mix of a denitrifying and methanogenic mixed culture, dosing bicarbonate as sole carbon source. Anaerobic mixed culture fed with acetate served as anode inoculum. With graphite plate electrodes and an external 1000 Ω resistor, the authors reported power densities of 7.0 mW/m2 and 55.5 mW/m2 depending on initial Cr(VI) concentration (22 and 63 mg/L, respectively). The maximum specific Cr(VI) reduction rate, about 0.46 mgCrVI/gVSS/h, was registered at a 63 mgCr(VI)/L initial concentration. Analysis of the Cr(VI) reduction community by 16S rRNA gene sequences showed a predominance of phylotypes related to Trichococcus pasteurii and P. aeruginosa. Even considering the small amount of substrate that can leak from the anode, even when an ion-exchange membrane is used [104], and organic carbon released in cell lysis, most Cr(VI) reduction was obtained with autotrophic conditions.
In a batch-fed 2CMFC, the cathode is inoculated with a mixed microbial consortium from a Cr(VI) contaminated site and 39.2 mg Cr(VI)/L. Huang et al. [90] observed a specific reduction rate of about 2.4 mgCrVI/gVSS/h, and 3.9 W/m2 maximum power production at a current density of 11.1 mA/m2.
Anaerobic pure cultures were also tested [91,92,102]. Hsu et al. [91] compared Cr(VI) reduction by six Shewanella strains at the cathode of MFCs in repeated cycles, observing initially the use of the electrode as the sole electron source in all tested strains. The variability in Cr(VI) reduction was associated with different mechanisms of chromium reduction, not identified, for each Shewanella strain evaluated, and other factors such as biofilm attachment to the electrode. Repeated Cr(VI) injections resulted in a general decrease in the MFCs performances and high residual Cr(VI) concentrations, which were explained with microorganisms‘ finite tolerance limit to Cr(VI) exposure and gradual fouling of the system by biological or reduced chromium species, which limit the active surface area of the cathode. Xafenias et al. [92] inoculated the cathode of an MFC and a MEC with S. oneidensis MR-1 fed with lactate. The combined use of the electrode and lactate as electron donors allowed bio-electrochemical and non-bio-electrochemical Cr(VI) reduction at the same time, even the contribution of the two different mechanisms to the overall process was not recognized. In Wu et al. [102], Bacillus sp. showed efficient Cr(VI)-reducing ability in both heterotrophic and autotrophic environments. The Cr(VI) removal rate reached 2.56 mg/L/h, which was 1.75 times higher than that of the MFC with the sterile control cathode.

3.1. Effects of pH and Cr(VI) Concentration

Extreme pH values (indicatively pH <5 or >8) and/or high chromium concentrations, typically 10–100 mg/L, can inhibit microbial activity. Tandukar et al. [89] reported that initial Cr(VI) concentrations above 80 mg/L inhibited the reduction rates in a denitrifying community. Li et al. [93] observed 10 mg Cr(VI)/L to irreversibly inhibit microbial activity in a single chamber MFC inoculated with municipal wastewater. Below toxic levels, increased initial Cr(VI) concentration, and following thermodynamics, was associated with an improved specific chromium reduction rate and MFC’s power production [90,93].
The pH, with its effects on the surface properties of the cells, including cell surface hydrophobicity, net surface electrostatic charge, and biofilm structure, may also heavily affect complex biological and electrochemical reactions at the biocathode. Variation in pH may also affect enzymatic activity, and produced Cr(III) precipitation or bio-adsorption [105]. In Huang et al. [64,65], 50 mg/L initial Cr(VI) concentration inhibited the catalytic activity of electrochemical bacteria in the biocathode, whereas, at a 20 mg/L Cr(VI) concentration, chromium reduction efficiencies increased (+27.3%) and decreased (−21%) in acidic (pH = 5) and alkaline catholyte (pH = 8), with respect to neutral pH. A 0.22 cell net potential increase, from 0.54 V at a pH of 8.0 to 0.76 V at a pH of 5.0, beyond the theoretical value of 0.177 V derived by Nernst’s law, was associated with a pH decrease in the cathodic compartment, which actually indicates a positive response of microorganisms’ activity associated with a pH decline [64]. Similar effects have been reported for denitrifying biocathodes [106]. Clearly, pH also affected the Cr(III) precipitation, with 9.3 mg/L dissolved Cr(III) at the end of the test at pH 5.0, in comparison to 0.3 mg/L at a pH of 8.0.

3.2. Effects of Cathode Potential

Tests with potentiostatically controlled cathodes pointed out an optimal potential range that typically exists for enhancing Cr(VI) reduction performances in biocathodes [95,107]. Theoretically, from Nernst’s law, in MFC with a chromium-reducing cathode and acetate-oxidizing bioanode, the open-circuit voltage at pH 7.0 and 25 °C is about 0.68 V, which results in about 0.4 V theoretical cathode potential [90].
Lower set cathode potentials would promote the Cr(VI) reduction process. Huang et al. [64] compared the behavior of a potentiostatically controlled BES (with cathode operated at 200, −150, −300, and −450 mV vs. SHE) to an MFC operating with 200 Ω external load. Cathode at −150/−300 mV set potential promoted fast start-up time (19 days compared to 26 days in the uncontrolled MFC or 28 days in +200 mV set cathode system) and Cr(VI) reduction, with almost complete removal of 20 mg/L in 24 h, with respect to 43–70% with the other systems. Furthermore, +200 mV and −450 mV poised cathode limited bacterial growth, whereas −150 and −300 mV had beneficial effects. In all the tests, the reduction of Cr(VI) was attributed to microorganisms by directly accepting electrons from the electrode surface and transferring them to Cr(VI), as, even in the test, at the most negative potential, no production of hydrogen gas was observed. Optimal set potential can provide an appropriate selective pressure for adaptation of the microbial community in the system, which leads to enhancements of microbial electrochemical interaction with the cathode. The difference between Cr(VI) reduction potential and the cathode set potential represents the maximum energy to be gained by the cathodic microorganisms. Thus, the lower the set cathode potential is, the more energy microorganisms will potentially obtain. However, in case the cathode potential is set too low and goes beyond the self-regulation capability of microbial consortia, the energy gain by the cathodic microorganisms gets lost. Likely −150 and −300 mV set potentials allowed the biomass to gain more energy than when 200 mV set the potential. Although, theoretically most favorable, −450 mV may have exceeded the self-regulation capability of the microbial consortia, with no positive effect on power generation and Cr(VI) reduction [64].
Xafenias et al. [92] demonstrated the positive impact on Cr(VI) reduction of riboflavin, which is a naturally produced mediator, in potentiostatically-controlled Shewanella oneidensis MR-1 biocathodes. Different configurations, with lactate supplied as electron donor in inoculated and abiotic systems, with or without riboflavin addiction, were tested. At 20 mg/L initial Cr(VI) concentration, in a −300 mV poised biocathode fed with lactate (30 mM, or equivalently 2700 mg/L), up to 45% Cr(VI) reduction was observed in 4 h in comparison with 5% Cr(VI) reduction in a biotic system with no lactate and 15% reduction in abiotic systems with lactate. In 2CMFC with S. oneidensis MR-1 fed with lactate in both anodic and cathodic compartments, Cr(VI) reduction at the cathode (10 mgCr(VI)/L initial concentration) was coupled with 32.5 mA/m2 maximum current density production [92].

3.3. Effects of Materials, Reactor Design, and Other Operational Parameters

Huang et al. [90] identified, together with Cr(VI) concentration, high conductivity of the electrolyte, (i.e., improved ion transport between the biofilm and bulk phase), as a key factor for efficient Cr(VI) reduction and power production. Increased conductivity of the solution, from 1.5 mS/cm to 10.6 mS/cm, increased the specific Cr(VI) reduction rate by about 25%, from 2.4 mg/(L gVSS h) to about 3.0 mg/(L gVSS h).
As to electrode materials, most experiences tested graphite or carbon-based electrodes. Huang et al. [90] tested graphite electrodes’ specific surface, by covering the cathode with graphite granules, to promote bacterial attachment and electrical connection between bacteria and the electrode surface. In a tubular 2CMFC, with a cathode to anode surface ratio (C/A) of 3, at a pH of 7, and 22 °C, the graphite fiber biocathode showed a higher specific Cr(VI) reduction rate and power generation than either graphite felt and granular graphite ones [65]. Specific Cr(VI) reduction rates on the graphite fiber cathodes, 12.4–20.6 mg/ gVSS /h, were about 10–100 folds higher than the values reported for biocatalyzed carbon plate or graphite granule cathode in H-type MFCs with about the same Cr(VI) concentrations [89,90].
These results underline the coordinated role of the cathode surface area and reactor architecture on the biocathode performance. In Wu et al. [108], NaX zeolite-modified graphite felts were used as electrodes (anode and cathode) in 2CMFCs. NaX zeolite proved to enhance the hydrophilicity of the graphite felt by facilitating bacterial adhesion and electrochemical reaction, and by decreasing mass transport resistances. Two different fabrication methods for the NaX zeolite-modified graphite felts were testedin which the first one was tested without any pre-treatment of the felt and the second one was tested with HNO3 pre-treatment. Both methods, especially the latter, resulted in excellent performance, with significant improvement in both electricity generation and Cr(VI) reduction rates, in comparison with graphite felts MFC. The HNO3 pre-process remarkably enhanced NaX loading mass on the graphite felt, by decreasing the organic residues on the graphite surface. NaX zeolite-modified graphite felts MFC at an initial Cr(VI) concentration of 20 mg/L resulted in more than 410 mV maximum voltage, 29 mW/m2 power density, and complete removal of Cr(VI) in 3 h, with an 8.2 times faster rate than simple graphite felts MFC. Nanostructured graphene also reduced Cr(VI) [99]. The maximum power density in an MFC with graphene biocathode was 5.7 times higher than the one produced with graphite felt biocathode. Electricity production, in fact, increased from 28.6 to 164 mW/m2. Furthermore, improved efficiency in Cr(VI) reduction was obtained, with 100% reduction in a 40 mgCr(VI)/L solution within 48 h, in comparison to only 58% reduction with graphite felt.
The most widely tested configuration with biocathodes is the 2CMFC. Even a study with a single-chamber reactor exists [93]. Organic substrate removal at the anode and cathodic chromium reduction were reflected in the open circuit potential of the system and Cr(III) deposition on the cathode, as revealed by scanning electron microscopy and energy-dispersive X-ray spectroscopy [93]. Cr(VI) conversion efficiencies ranged from 89% to 99% depending on initial Cr(VI) concentrations (89% at 1.1 mg/L, 95% at 3 mg/L, and 99% at 10 mg/L). In the open-circuit control, Cr(VI) conversion efficiency was lower and decreased with Cr(VI) concentrations (88% at 1.1mg/L, 63% at 3 mg/L, and 28% at 10 mg/L). This means the SCMFC took advantage of electroactive bacteria using Cr(VI) as electron acceptor, as the other Cr(VI) reduction mechanisms, including bio-adsorption or bio-reduction by inelectroactive bacteria, as the open-circuit control pointed out, were inhibited at high Cr(VI) concentrations [93].
To increase microbial concentration and prevent premature cathode passivation due to Cr(III) precipitates during the system set-up, Wu et al. [96] proposed an ex situ acclimatization method for Cr(VI)-reducing biocathodes. The electrode was initially enriched with exoelectrogenic biofilm as an MFC anode, and the system was subsequently established using the anode as biocathode. This method allowed for the development of a mature biofilm in a shorter period of acclimatization (<19 days in the authors’ experience) compared to traditional in situ methods, with Cr(VI) removal reaching 79% in 24 h, which is about four times higher than the one observed in the MFC with an in situ acclimated cathode. The improved performance was attributed not only to avoidance of premature formation of Cr(III) precipitates on the electrode, during biofilm acclimatization, but also to the enhanced bacterial growth rates in the heterotrophic anodic environment, which leads to high microbial density and bacterial coverage of the electrode. This may limit the effects of Cr(VI) toxicity on the microorganisms, at the anode/cathode inversion.

3.4. Simultaneous Reduction of Cr(VI) and Other Metals

As abiotic cathodes, biocathodes are also being tested for the simultaneous reduction of multiple metals, usually present in a variety of metal-processing wastewaters. Huang et al. [94] demonstrated that bacterial communities in biocathodes could adaptively evolve to utilize solutions containing mixtures of metals. Reduction rates of Cr(VI), Cu(II), and Cd(III) in BESs with biocathodes individually acclimated to the three different metals or acclimated to increased concentrations of a mixture of metalswere compared. In a Cr(VI) acclimated biocathode, the Cr(VI) reduction rate decreased from about 1.21 mg/L/h to 0.49 mg/L/h in the presence of 5 mg/L Cu(II) and 5 mg/L Cd(II). Acclimatization, by gradually increasing concentrations of mixed metals, allowed complete removal of Cr(VI) at a rate similar to that originally obtained with the reactor acclimated to Cr(VI) only. Analysis of bacterial communities showed different communities on the biocathodes of the reactors acclimated to the mixed metal solutions, compared to reactors acclimated only to a single metal. The decrease in diversity of the microbial communities was likely due to the greater toxicity of the mixed metals compared to only single metals. At the phylum level, compared to Cr(VI) acclimated biofilm, the relative abundance of Proteobacteria, Actinobacteria, Firmicutes, and Tenericutes increased in biofilms acclimated to mixed metals. Synergistetes, to the contrary, were absent following acclimation to the mixed metals.
Huang et al. [100] examined Cr(VI) and Cd(II) reductions on biocathode in pure culture MFC experiments with known electrochemically active bacteria, Stenotrophomonas sp. YS1, Stenotrophomonas maltophilia YS2, Serratia marcescens YS3, and Achromobacter xylosoxidans YS8. Cr(VI) reduction in the MFCs decreased in the presence of Cd(II) for all the pure cultures, with removals in a 5-h period ranging from 63% to 71%, depending on the species, compared to a range of 73–82% when Cd(II) was absent. Cr(VI) removal in biocathodes was higher than in the abiotic cathode, limited to about 39%. Cd(II) removal, on the contrary, was not impacted by the presence of Cr(VI).

4. Cr(VI) Reduction at Bioanode

Yeon et al. [57] enriched electroactive Cr(VI)-reducing bacteria in the anode compartment of MFC with the air cathode, using Cr(VI)-containing sludge from a leather tanning wastewater treatment plant fed with synthetic wastewater. At the end of the enrichment procedure, Cr(VI) removal capability of such biofilm was observed with a 93% reduction of 5 mg/L Cr(VI) and 61% of 25 mg/L Cr(VI). MFC-mediated Cr(VI) removal was attributed to either physical adsorption on the carbon felt anode (about 20%) and biological reactions including biosorption or reduction to Cr(III).
The bacterial community analysis by polymerase chain reaction—denaturing gradient gel electrophoresis (PCR-DGGE) of 16S rDNA, after enrichment, pointed out the microbial consortium is composed of both Cr(VI) reducers with either electrochemical activity (such as Clostridium sp.) or not (like Acinetobacter sp.), and non-Cr(VI) reducers with/without electrochemical activity (as Actinobacteria sp). Electroactive bacteria were responsible for electricity production in the MFC. However, reducers with electrochemical activity used Cr(VI) as an electron acceptor instead of interacting with the electrode, which causes a decrease in the current. Cr(VI) reduction was likely performed also by Cr(VI) reducers without electrochemical activity, with the required protons supplied by the metabolism of fermentative bacteria.

5. Cr(VI) Bio-Electrochemical Remediation

This review of published research, targeting treatment of Cr(VI) contaminated wastewater/industrial effluents, offers the first proof of concept for the chance of bio-electrochemical Cr(VI) remediation. Experiences with conventional bioreduction processes, under both aerobic and anaerobic conditions, by either pure cultures or mixed consortia, refer of Cr(VI) bioreduction rates in a 0.1–13.5 mg/L/h range [22,62,109,110,111], which are fully comparable to the values, 0.1–6.6 mg/L/h, observed in Cr(VI) reducing biocathodes.
Therefore, with comparable results, bioelectrochemical Cr(VI) remediation over the conventional bioremediation presents the following advantages: (i) the chance of excluding any external chemical supply to support the microorganisms’ actions, (ii) ease of monitoring and control of the process, and (iii) the potential recovery of reduced chromium deposited on the electrode.
It should be noted that all the reviewed studies were batch laboratory tests, under conditions quite different from natural Cr(VI) contaminated water/groundwater. It would, therefore, be useful to perform evaluations under dynamic water flow conditions as in contaminated aquifers, with real groundwater, as reported by Gregory and Lovley [49] for uranium-contaminated aquifers.
For groundwater remediation, it is important to take into consideration specific properties that potentially affect BES operation [20]. Although Cr(VI) contamination likely increases the specific conductivity, typical low specific conductivity value of groundwater (well below 2 mS/cm) can be negatively impacted on BESs by implying higher ohmic and transport losses [30]. Moreover, pH shifts due to electrochemical Cr(VI) reduction in low buffering capacity systems may directly harm the electroactive bacteria and their removal performance [59]. Another challenge for bio-electrochemical treatment of contaminated groundwater is the presence, in addition to the contaminants, of a mixture of various naturally occurring inorganic (calcium, magnesium, carbonate, nitrates and sulphates, metals) and organic chemicals (e.g., humic acids) [112]. Magnesium and calcium can produce precipitates that could passivate the cathode with the consequent reduction of surface exchange active area [113]. Bio-electrochemical reduction has been recently reported for nitrate [97,101,114] and sulphate [115]. So far, the study of co-contaminants with BESs is limited, but theoretically, since reduction potentials of nitrate and sulphate are similar to the reduction potential of several pollutants, they can be electron competitors in the remediation process and affect the microbial community at the biocathode [116]. Wang et al. [101] evaluated the simultaneous autotrophic denitrification and the reduction of Cr(VI) under different pH conditions (6, 7, and 8). The highest removal efficiencies for nitrates (97%) and Cr(VI) (73%) were obtained at a pH of 7. The stable combined reduction was mainly ascribed to Pseudomonas, Halomonas, and Thauera species.
Chen and colleagues [97] used a 3 L cylindrical single-chamber reactor with a graphite felt cathode and a central carbon rod anode. The reactor was filled with sulphur granules and inoculated with anaerobic sludge. The reactor was continuously fed with 100 mg/L Cr(VI) synthetic wastewater with no organic C source (16 h of a hydraulic retention time) and run in the galvanostatic mode (current 10–60 mA). Cr(VI) reduction in the effluent ranged between 43% and 97%, which is proportional to the externally supplied current. This observation, together with SO42− in the effluent, highlighted both sulphur and hydrogen autotrophic bacteria were responsible for Cr(VI) reduction by using the S granules in the reactor and H2 produced by the cathode as electron donors. A similar system, which is a single chamber cylindrical reactor operated in galvanostatic (200 mA) continuous flow mode (20 h HRT), was adopted by Wang et al. [101] for removing Cr(VI) and nitrates from synthetic wastewater.
A continuous-flow BES was proposed for the simultaneous removal of p-fluoronitrobenzene (p-FNB), nitrates, and hexavalent chromium from synthetic wastewater as well [97]. In this co-contaminated system, the competition for electrons, for the carbon source and metabolism of microorganisms negatively influenced the degradation rates in comparison with the single pollutant control tests. The biodegradation of p-FNB in the co-contaminated system produced an additional organic carbon source to the microorganisms that promoted Cr(VI) and nitrates removal (nitrate and Cr(VI) removal through degradation of p-FNB). Instead, the p-FNB removal rate was controlled by electron availability (p-FNB degradation increased at currents above 40 mA) [97].
It is also interesting to consider the full cycle of sulphur in BESs. The role of the sulphur cycle during the electro-bioremediation of oil spills has been recently reviewed [43]. Sulphide produced by sulphate reducers can be oxidized to elemental sulphur on the anode surface [117]. Elemental sulphur can be back oxidized to sulphate [118] or can be reduced again to sulphide [119]. The sulphur cycle in BESs can, thus, be effective in enhancing current production (i.e., via sulphide recycling) or in supplying electron acceptors for biodegradation of reduced pollutants, such as hydrocarbons (i.e., via back oxidation of sulphur to sulphate). In this context, it is crucial to understand the possible role of Cr(VI) in environments in which co-contamination occurs. Reduction of Cr(VI) to Cr(III) can represent a sink for sulphide oxidation, which affects the performance of the process. Whether the effect is positive or negative on the anodic oxidation is still an open question. During oxidation of BTEX (benzene, toluene, ethylbenzene, xylenes) mixtures in BESs where sulphate was present in the medium (250 mg/L), the bacterial communities enriched on the anodes were dominated by microorganisms linked to the sulphur cycle. However, bacteria able to oxidize hydrocarbons and to perform direct electron transfer to the electrode (i.e., Geobacter spp.) were also detected [120]. In similar conditions, the competition between the anode and chromium for the scavenging of sulphide could facilitate the enrichment on the electrode of microorganisms not directly linked to the sulphur cycle. No study of Cr(VI) bioelectrochemical remediation in the presence of nitrate and sulphate is available. However, interferences with chromium reduction are likely to occur. Therefore, further understanding of chemical species that coexist with the target pollutant in groundwater is required.
In view of in situ applications, the effects of soil particles on pollutant partitioning and bioavailability, as well as system conductance need to be assessed. Soil type and external resistance significantly affected the current and Cr(VI) removal efficiency in soil MFCs tests operated at external resistances of 100 and 1000 Ω for 16 days [121]. The current production and Cr(VI) reduction in red soil and fluvo-aquic soil MFCs were compared. Red soil MFC performed better in the current production, but showed a lower Cr(VI) removal than fluvo-aquic soil MFC, which implies red soil may contain more electron acceptors that competed with the Cr(VI) reduction reaction [121]. About 60% to 90% of Cr(VI) was removed in 16 days of operation of a soil MFCs, while only 32–46% was removed in the open circuit control. Experiences integrating plants, microbes, and electrochemistry revealed promising applications of BESs to shallow contaminations [122,123], since plants can rely on atmospheric CO2 for photosynthesis and secrete root exudates that can serve as carbon sources and electron donors for microbes in the rhizosphere to promote biodegradation/biostabilization.

6. Conclusions

Ever since the discovery of microbial remediation methods for Cr(VI), many technological approaches have been developed, and some are already used in full-scale treatments. As compared to other energy-intensive technologies, bioremediation is considered a promising cost-efficient and sustainable option. Microbial electrochemical systems have been recently proposed as an alternative platform for bioremediation of Cr(VI) and other toxic chemicals. BESs, in comparison with other bioremediation techniques, are particularly appealing for in situ applications, since they do not require relevant chemical addition in the subsurface and may entail a low energy supply. Reducing energy costs and chemical soil amendments implies lower operating costs for the BESs, which is particularly valuable considering the typical timeframe required for groundwater remediation. Moreover, in the Cr(VI) contamination treatment, the deposition of Cr(III) onto the electrode may enable extraction and recovery of Cr(III).
In lab-scale studies, BESs were competent in reducing at the cathode Cr(VI) in contaminated water streams with initial concentrations as low as 1 mg/L up to about 1 g/L. Even though positive results have been reported with abiotic cathode systems, biocathodes offer several advantages in Cr(VI) reduction from the perspective of groundwater remediation, such as the effectiveness in the natural waters’ pH range and the exploitation of microbial catalysis, which limits cathode passivation due to Cr(III) precipitation.
Nonetheless, much work is still needed to improve the Cr(VI) reduction rate at the biocathode to maximize the advantages of biocathode BESs over conventional biological processes. A lack in the long-term pilot and scale-up research suggests that more focus should be given to key factors that need to be considered for fully-practiced feasibility studies and full-scale applications. For example, the stability of processes and equipment and the area of influence that each BES module can effectively cover have to be addressed. In addition, flexible configurations to adapt to different site-specific characteristics (for example, water table depth, soil types, etc.) are required for in situ remediation. No preliminary models have been developed to start addressing this issue. Further research is needed to understand the mechanisms driving BES function and, thus, to fully exploit BES potential in real contaminated soil and groundwater. Many challenges, including the selection of microorganisms, cheap materials, and an optimal reactor configuration, need to be addressed during the technology scale-up.

Author Contributions

Conceptualization, G.B., A.F.M. and E.S.; literature review of microbial mechanisms of Cr(VI) reduction G.B., A.E.T. and M.D.; literature review of electrochemical and bio-cathodic Cr(VI) reduction, G.B., A.F.M. and E.S.; organization of the information, synthesis and critical analysis, G.B., A.E.T., A.F.M. and E.S.; writing—original draft preparation, A.F.M. and E.S.; writing—review and editing, G.B., A.E.T., M.D. and E.S.; project supervision, S.S. and A.F.

Funding

The Fondazione Cariplo in the framework of the project BEvERAGE—BioElEctrochemical RemediAtion of Groundwater plumes (2015-0195) supported this work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Humphries, A.C.; Nott, K.P.; Hall, L.D.; Macaskie, L.E. Reduction of Cr(VI) by immobilized cells of desulfovibrio vulgaris NCIMB 8303 and Microbacterium sp. NCIMB 13776. Biotechnol. Bioeng. 2005, 90, 589–596. [Google Scholar] [CrossRef] [PubMed]
  2. Marsh, T.L.; McInerney, M.J. Relationship of Hydrogen Bioavailability to Chromate Reduction in Aquifer Sediments. Appl. Environ. Microbiol. 2001, 67, 1517–1521. [Google Scholar] [CrossRef] [PubMed]
  3. Palmer, C.D.; Puls, R.W. Natural Attenuation of Hexavalent Chromium in Ground Water and Soils; EPA/540/S-94/505; U.S. Environmental Protection Agency Ground Water Issue; Superfund Technology Support Center for Ground Water—Robert S. Kerr Environmental Research Laboratory: Ada, OK, USA, 1994; pp. 1–13. [Google Scholar]
  4. Rai, D.; Eary, L.E.; Zachara, J.M. Environmental chemistry of chromium. Sci. Total Environ. 1989, 86, 15–23. [Google Scholar] [CrossRef]
  5. Beverskog, B.; Puigdomenech, I. Revised Pourbaix diagrams for Chromium at 25–300 °C. Corros. Sci. 1997, 39, 43–57. [Google Scholar] [CrossRef]
  6. Zayed, A.M.; Terry, N. Chromium in the environment: Factors affecting biological remediation. Plant Soil 2003, 249, 139–156. [Google Scholar] [CrossRef]
  7. Chirwa, E.M.; Molokwane, P.P.P. Biological Cr(VI) reduction: Microbial diversity, kinetics and biotechnological solutions to pollution. In Biodiversity; Sofo, A., Ed.; InTech: Rijeka, Croatia, 2011; pp. 75–100. [Google Scholar]
  8. Jadhav, U.U.; Hocheng, H. A review of recovery of metals from industrial waste. J. Achiev. Mater. Manuf. Eng. 2012, 54, 159–167. [Google Scholar]
  9. Independent Environmental Technical Evaluation Group. Chromium (VI) Handbook; Guertin, J., Jacobs, J.A., Avakian, C.P., Eds.; CRC Press: New York, NY, USA, 2005. [Google Scholar]
  10. European Food Safety Authority. Scientific Opinion on the risks to public health related to the presence of chromium in food and drinking water. EFSA J. 2014, 12, 3595. [Google Scholar]
  11. Malaviya, P.; Singh, A. Physicochemical Technologies for Remediation of Chromium-Containing Waters and Wastewaters. Crit. Rev. Environ. Sci. Technol. 2011, 41, 1111–1172. [Google Scholar] [CrossRef]
  12. Jin, W.; Du, H.; Zheng, S.; Zhang, Y. Electrochemical processes for the environmental remediation of toxic Cr(VI): A review. Electrochim. Acta 2015, 191, 1044–1055. [Google Scholar] [CrossRef]
  13. Brungesh, K.; Nagabhushana, B.; Raveendra, R.; Krishna, K.; Prashantha, P.; Nagabhushana, H. Adsorption of Cr(VI) from Aqueous Solution onto a Mesoporous Carbonaceous Material Prepared from Naturally Occurring Pongamia pinnata Seeds. Environ. Anal. Toxicol. 2015, 5, 1. [Google Scholar]
  14. Caputo, D.; De Gennaro, B.; Pansini, M.; Colella, C. Chromium removal from water by ion exchange using zeolites and solidification of the resulting sludge in a cement matrix. In Studies in Surface Science and Catalysis; Kiricsi, I., Nagy, J.B., Karge, H.G., Palyi, G., Eds.; Elsevier Science Publisher B.V.: Amsterdam, the Netherlands, 1999; Volume 125, pp. 723–730. [Google Scholar]
  15. Rajeswari, V.; Janaki, V.; Shanthi, K.; Kamala-Kannan, S. Adsorption and subsequent detoxification of hexavalent chromium in aqueous solution using polypyrrole-bacterial extracellular polysaccharide nanocomposite. Environ. Prog. Sustain. Energy 2016, 35, 1293–1297. [Google Scholar] [CrossRef]
  16. Jeyasingh, J.; Philip, L. Bioremediation of chromium contaminated soil: Optimization of operating parameters under laboratory conditions. J. Hazard. Mater. 2005, 118, 113–120. [Google Scholar] [CrossRef] [PubMed]
  17. Palmer, C.D.; Wittbrodt, P.R. Processes affecting the remediation of chromium-contaminated sites. Environ. Health Perspect. 1991, 92, 25–40. [Google Scholar] [CrossRef] [PubMed]
  18. Tseng, J.K.; Bielefeldt, A.R. Low-Temperature Chromium(VI) Biotransformation in Soil with Varying Electron Acceptors. J. Environ. Qual. 2002, 31, 1831–1841. [Google Scholar] [CrossRef] [PubMed]
  19. Turick, C.E.; Graves, C.; Apel, W.A. Bioremediation potential of Cr(VI)-contaminated soil using indigenous microorganisms. Bioremediat. J. 1998, 2, 1–6. [Google Scholar] [CrossRef]
  20. Pous, N.; Balaguer, M.D.; Colprim, J.; Puig, S. Opportunities for groundwater microbial electro-remediation. Microb. Biotechnol. 2018, 11, 119–135. [Google Scholar] [CrossRef] [PubMed]
  21. Brodie, E.L.; Joyner, D.C.; Faybishenko, B.; Conrad, M.E.; Rios-Velazquez, C.; Malave, J.; Martinez, R.; Mork, B.; Willett, A.; Koenigsberg, S.; et al. Microbial community response to addition of polylactate compounds to stimulate hexavalent chromium reduction in groundwater. Chemosphere 2011, 85, 660–665. [Google Scholar] [CrossRef]
  22. Somenahally, A.C.; Mosher, J.J.; Yuan, T.; Podar, M.; Phelps, T.J.; Brown, S.D.; Yang, Z.K.; Hazen, T.C.; Arkin, A.P.; Palumbo, A.V.; et al. Hexavalent chromium reduction under fermentative conditions with lactate stimulated native microbial communities. PLoS ONE 2013, 8, e83909. [Google Scholar] [CrossRef]
  23. Deng, B.; Stone, A.T. Surface-Catalyzed Chromium(VI) Reduction: The TiO2−CrVI−Mandelic Acid System. Environ. Sci. Technol. 1996, 30, 463–472. [Google Scholar] [CrossRef]
  24. Beretta, G.; Mastorgio, A.F.; Pedrali, L.; Saponaro, S.; Sezenna, E. Support tool for identifying in situ remediation technology for sites contaminated by hexavalent chromium. Water 2018, 10, 1344. [Google Scholar] [CrossRef]
  25. Elangovan, R.; Philip, L.; Chandraraj, K. Hexavalent chromium reduction by free and immobilized cell-free extract of arthrobacter rhombi-RE. Appl. Biochem. Biotechnol. 2010, 160, 81–97. [Google Scholar] [CrossRef] [PubMed]
  26. Yang, J.; He, M.; Wang, G. Removal of toxic chromate using free and immobilized Cr(VI)-reducing bacterial cells of intrasporangium sp. Q5-1. World J. Microbiol. Biotechnol. 2009, 25, 1579–1587. [Google Scholar] [CrossRef]
  27. Liu, H.; Guo, L.; Liao, S.; Wang, G. Reutilization of immobilized fungus Rhizopus sp. LG04 to reduce toxic chromate. J. Appl. Microbiol. 2012, 112, 651–659. [Google Scholar] [CrossRef] [PubMed]
  28. Butler, E.C.; Chen, L.; Hansel, C.M.; Krumholz, L.R.; Elwood Madden, A.S.; Lan, Y.; Madden, A.S.E.; Lan, Y. Biological versus mineralogical chromium reduction: Potential for reoxidation by manganese oxide. Environ. Sci. Process. Impacts 2015, 17, 1930–1940. [Google Scholar] [CrossRef] [PubMed]
  29. Varadharajan, C.; Beller, H.R.; Bill, M.; Brodie, E.L.; Conrad, M.E.; Han, R.; Irwin, C.; Larsen, J.T.; Lim, H.C.; Molins, S.; et al. Reoxidation of Chromium(III) Products Formed under Different Biogeochemical Regimes. Environ. Sci. Technol. 2017, 51, 4918–4927. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Logan, B.E.; Hamelers, B.; Rozendal, R.; Schröder, U.; Keller, J.; Freguia, S.; Aelterman, P.; Verstraete, W.; Rabaey, K. Microbial fuel cells: Methodology and technology. Environ. Sci. Technol. 2006, 40, 5181–5192. [Google Scholar] [CrossRef] [PubMed]
  31. Shi, L.; Dong, H.; Reguera, G.; Beyenal, H.; Lu, A.; Liu, J.; Yu, H.-Q.Q.; Fredrickson, J.K. Extracellular electron transfer mechanisms between microorganisms and minerals. Nat. Rev. Microb. 2016, 14, 651–662. [Google Scholar] [CrossRef]
  32. Rabaey, K.; Angenent, L.T.T.; Schröder, U.; Keller, J. Bioelectrochemical Systems: From Extracellular Electron Transfer to Biotechnological Application; IWA Publishing: London, UK, 2009. [Google Scholar]
  33. Kokabian, B.; Gude, V.G. Role of membranes in bioelectrochemical systems. Membr. Water Treat. 2015, 6, 53–75. [Google Scholar] [CrossRef]
  34. Gude, V.G.; Kokabian, B.; Gadhamshetty, V.G. Beneficial Bioelectrochemical Systems for Energy, Water, and Biomass Production. J. Microb. Biochem. Technol. 2013, S6:005, 1–14. [Google Scholar]
  35. Santoro, C.; Arbizzani, C.; Erable, B.; Ieropoulos, I. Microbial fuel cells: From fundamentals to applications. A review. J. Power Sources 2017, 356, 225–244. [Google Scholar] [CrossRef]
  36. Rabaey, K.; Rozendal, R.A. Microbial electrosynthesis—Revisiting the electrical route for microbial production. Nat. Rev. Microbiol. 2010, 8, 706–716. [Google Scholar] [CrossRef] [PubMed]
  37. Pant, D.; Singh, A.; Van Bogaert, G.; Irving Olsen, S.; Singh Nigam, P.; Diels, L.; Vanbroekhoven, K. Bioelectrochemical systems (BES) for sustainable energy production and product recovery from organic wastes and industrial wastewaters. RSC Adv. 2012, 2, 1248–1263. [Google Scholar] [CrossRef]
  38. Li, X.; Abu-Reesh, I.; He, Z. Development of Bioelectrochemical Systems to Promote Sustainable Agriculture. Agriculture 2015, 5, 367–388. [Google Scholar] [CrossRef] [Green Version]
  39. Nancharaiah, Y.V.; Venkata Mohan, S.; Lens, P.N.L. Metals removal and recovery in bioelectrochemical systems: A review. Bioresour. Technol. 2015, 195, 102–114. [Google Scholar] [CrossRef] [PubMed]
  40. Lovley, D.R. Powering microbes with electricity: Direct electron transfer from electrodes to microbes. Environ. Microbiol. Rep. 2011, 3, 27–35. [Google Scholar] [CrossRef] [PubMed]
  41. Wang, H.; Luo, H.; Fallgren, P.H.; Jin, S.; Ren, Z.J. Bioelectrochemical system platform for sustainable environmental remediation and energy generation. Biotechnol. Adv. 2015, 33, 317–334. [Google Scholar] [CrossRef]
  42. Palma, E.; Daghio, M.; Franzetti, A.; Petrangeli Papini, M.; Aulenta, F. The bioelectric well: A novel approach for in situ treatment of hydrocarbon-contaminated groundwater. Microb. Biotechnol. 2018, 11, 112–118. [Google Scholar] [CrossRef]
  43. Daghio, M.; Aulenta, F.; Vaiopoulou, E.; Franzetti, A.; Arends, J.B.A.; Sherry, A.; Suárez-Suárez, A.; Head, I.M.; Bestetti, G.; Rabaey, K. Electrobioremediation of oil spills. Water Res. 2017, 114, 351–370. [Google Scholar] [CrossRef] [Green Version]
  44. Zhang, T.; Gannon, S.M.; Nevin, K.P.; Franks, A.E.; Lovley, D.R. Stimulating the anaerobic degradation of aromatic hydrocarbons in contaminated sediments by providing an electrode as the electron acceptor. Environ. Microbiol. 2010, 12, 1011–1020. [Google Scholar] [CrossRef]
  45. Nguyen, V.K.; Park, Y.; Yu, J.; Lee, T. Simultaneous arsenite oxidation and nitrate reduction at the electrodes of bioelectrochemical systems. Environ. Sci. Pollut. Res. 2016, 23, 19978–19988. [Google Scholar] [CrossRef]
  46. Lai, A.; Aulenta, F.; Mingazzini, M.; Palumbo, M.T.; Papini, M.P.; Verdini, R.; Majone, M. Bioelectrochemical approach for reductive and oxidative dechlorination of chlorinated aliphatic hydrocarbons (CAHs). Chemosphere 2017, 169, 351–360. [Google Scholar] [CrossRef] [PubMed]
  47. Leitão, P.; Nouws, H.; Danko, A.S.; Aulenta, F. Bioelectrochemical Dechlorination of 1,2-DCA with an AQDS-Functionalized Cathode Serving as Electron Donor. Fuel Cells 2017, 17, 612–617. [Google Scholar] [CrossRef]
  48. Cecconet, D.; Devecseri, M.; Callegari, A.; Capodaglio, A.G. Effects of process operating conditions on the autotrophic denitrification of nitrate-contaminated groundwater using bioelectrochemical systems. Sci. Total Environ. 2018, 613, 663–671. [Google Scholar] [CrossRef] [PubMed]
  49. Gregory, K.B.; Lovley, D.R. Remediation and Recovery of Uranium from Contaminated Subsurface Environments with Electrodes. Environ. Sci. Technol. 2005, 39, 8943–8947. [Google Scholar] [CrossRef]
  50. Harnisch, F.; Aulenta, F.; Schröder, U. Microbial Fuel Cells and Bioelectrochemical Systems: Industrial and Environmental Biotechnologies Based on Extracellular Electron Transfer. In Comprehensive Biotechnology, 2nd ed.; Moo-Young, M., Ed.; Academic Press: Burlington, NJ, USA, 2011; Volume 6, pp. 643–659. [Google Scholar]
  51. Wang, H.; Ren, J.; Ren, Z.J.; Ren, J.; Ren, Z.J. Bioelectrochemical metal recovery from wastewater: A review. Water Res. 2014, 66, 219–232. [Google Scholar] [CrossRef]
  52. Wu, Y.; Qi, H.; Gao, Y.; Wang, L. Heavy Metal Sensor Research Based on Microbial Fuel Cell. Int. J. Environ. Monit. Anal. 2018, 6, 53–64. [Google Scholar] [CrossRef]
  53. Zhao, S.; Liu, P.; Niu, Y.; Chen, Z.; Khan, A.; Zhang, P.; Li, X. A Novel Early Warning System Based on a Sediment Microbial Fuel Cell for In Situ and Real Time Hexavalent Chromium Detection in Industrial Wastewater. Sensors 2018, 18, 642. [Google Scholar] [CrossRef]
  54. Zhou, T.; Han, H.; Liu, P.; Xiong, J.; Tian, F.; Li, X. Microbial fuels cell-based biosensor for toxicity detection: A review. Sensors 2017, 17, 2230. [Google Scholar] [CrossRef]
  55. Williams, K.H.; Nevin, K.P.; Franks, A.; Englert, A.; Long, P.E.; Lovley, D.R. Electrode-based approach for monitoring in situ microbial activity during subsurface bioremediation. Environ. Sci. Technol. 2010, 44, 47–54. [Google Scholar] [CrossRef]
  56. Sophia, A.C.; Saikant, S. Reduction of chromium(VI) with energy recovery using microbial fuel cell technology. J. Water Process Eng. 2016, 11, 39–45. [Google Scholar] [CrossRef]
  57. Yeon, R.E.; Kim, M.; Lee, S.J. Characterization of microbial fuel cells enriched using Cr(VI)-containing sludge. J. Microbiol. Biotechnol. 2011, 21, 187–191. [Google Scholar]
  58. Clauwaert, P.; Aelterman, P.; Pham, T.H.; De Schamphelaire, L.; Carballa, M.; Rabaey, K.; Verstraete, W. Minimizing losses in bio-electrochemical systems: The road to applications. Appl. Microbiol. Biotechnol. 2008, 79, 901–913. [Google Scholar] [CrossRef] [PubMed]
  59. Wang, G.; Huang, L.; Zhang, Y. Cathodic reduction of hexavalent chromium [Cr(VI)] coupled with electricity generation in microbial fuel cells. Biotechnol. Lett. 2008, 30, 1959–1966. [Google Scholar] [CrossRef] [PubMed]
  60. Richard, F.C.; Bourg, A.C.M. Aqueous geochemistry of chromium: A review. Water Res. 1991, 25, 807–816. [Google Scholar] [CrossRef]
  61. Sophia, A.C.; Sai, S. Modified microbial fuel cell for Cr(VI) reduction and simultaneous bio-electricity production. J. Environ. Chem. Eng. 2016, 4, 2402–2409. [Google Scholar] [CrossRef]
  62. Molokwane, P.E.; Meli, K.C.; Nkhalambayausi-Chirwa, E.M. Chromium (VI) reduction in activated sludge bacteria exposed to high chromium loading: Brits culture (South Africa). Water Res. 2008, 42, 4538–4548. [Google Scholar] [CrossRef] [Green Version]
  63. Li, M.; Zhou, S.; Xu, Y.; Liu, Z.; Ma, F.; Zhi, L.; Zhou, X. Simultaneous Cr(VI) reduction and bioelectricity generation in a dual chamber microbial fuel cell. Chem. Eng. J. 2018, 334, 1621–1629. [Google Scholar] [CrossRef]
  64. Huang, L.; Chai, X.; Chen, G.; Logan, B.E. Effect of set potential on hexavalent chromium reduction and electricity generation from biocathode microbial fuel cells. Environ. Sci. Technol. 2011, 45, 5025–5031. [Google Scholar] [CrossRef]
  65. Huang, L.; Chai, X.; Cheng, S.; Chen, G. Evaluation of carbon-based materials in tubular biocathode microbial fuel cells in terms of hexavalent chromium reduction and electricity generation. Chem. Eng. J. 2011, 166, 652–661. [Google Scholar] [CrossRef]
  66. Thatoi, H.; Das, S.; Mishra, J.; Rath, B.P.; Das, N. Bacterial chromate reductase, a potential enzyme for bioremediation of hexavalent chromium: A review. J. Environ. Manag. 2014, 146, 383–399. [Google Scholar] [CrossRef]
  67. Viti, C.; Marchi, E.; Decorosi, F.; Giovannetti, L. Molecular mechanisms of Cr(VI) resistance in bacteria and fungi. FEMS Microbiol. Rev. 2014, 38, 633–659. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Horitsu, H.; Futo, S.; Miyazawa, Y.; Ogai, S.; Kawai, K. Enzymatic reduction of hexavalent chromium by hexavalent chromium tolerant Pseudomonas ambigua G-1. Agric. Biol. Chem. 1987, 51, 2417–2420. [Google Scholar] [CrossRef]
  69. Ackerley, D.F.; Gonzalez, C.F.; Park, C.H.; Ii, R.B.; Matin, A.; Keyhan, M. Chromate-Reducing Properties of Soluble Flavoproteins from Pseudomonas putida and Escherichia coli. Am. Soc. Microbiol. 2004, 70, 873–882. [Google Scholar] [CrossRef] [PubMed]
  70. Cheung, K.H.; Gu, J.D. Mechanism of hexavalent chromium detoxification by microorganisms and bioremediation application potential: A review. Int. Biodeterior. Biodegrad. 2007, 59, 8–15. [Google Scholar] [CrossRef]
  71. Camargo, F.A.O.; Bento, F.M.; Okeke, B.C.; Frankenberger, W.T. Hexavalent chromium reduction by an actinomycete, Arthrobacter crystallopoietes ES 32. Biol. Trace Elem. Res. 2004, 97, 183–194. [Google Scholar] [CrossRef]
  72. Tahri Joutey, N.; Bahafid, W.; Sayel, H.; Ananou, S.; El Ghachtouli, N. Hexavalent chromium removal by a novel Serratia proteamaculans isolated from the bank of Sebou River (Morocco). Environ. Sci. Pollut. Res. 2014, 21, 3060–3072. [Google Scholar] [CrossRef]
  73. Ishibashi, Y.; Cervantes, C.; Silver, S. Chromium reduction in Pseudomonas putida. Appl. Environ. Microbiol. 1990, 56, 2268–2270. [Google Scholar]
  74. Mohapatra, R.K.; Parhi, P.K.; Thatoi, H.; Panda, C.R. Bioreduction of hexavalent chromium by Exiguobacterium indicum strain MW1 isolated from marine water of Paradip Port, Odisha, India. Chem. Ecol. 2017, 33, 114–130. [Google Scholar] [CrossRef]
  75. Lovley, D.R.; Phillips, E.J.P. Reduction of Chromate by Desulfovibrio-Vulgaris and Its c3 Cytochrome. Appl. Environ. Microbiol. 1994, 60, 726–728. [Google Scholar]
  76. Barrera-Díaz, C.E.; Lugo-Lugo, V.; Bilyeu, B. A review of chemical, electrochemical and biological methods for aqueous Cr(VI) reduction. J. Hazard. Mater. 2012, 223, 1–12. [Google Scholar] [CrossRef]
  77. Wang, P.C.; Mori, T.; Komori, K.; Sasatsu, M.; Toda, K.; Ohtake, H. Isolation and characterization of an Enterobacter cloacae strain that reduces hexavalent chromium under anaerobic conditions. Appl. Environ. Microbiol. 1989, 55, 1665–1669. [Google Scholar]
  78. Bencheikh-Latmani, R. Global transcriptional profiling of Shewanella oneidensis MR-1 during Cr (VI) and U (VI) reduction. Appl. Environ. Microbiol. 2005, 71, 7453–7460. [Google Scholar] [CrossRef]
  79. Belchik, S.M.; Kennedy, D.W.; Dohnalkova, A.C.; Wang, Y.; Sevinc, P.C.; Wu, H.; Lin, Y.; Lu, H.P.; Fredrickson, J.K.; Shi, L. Extracellular reduction of hexavalent chromium by cytochromes MtrC and OmcA of Shewanella oneidensis MR-1. Appl. Environ. Microbiol. 2011, 77, 4035–4041. [Google Scholar] [CrossRef]
  80. Němeček, J.; Pokorný, P.; Lhotský, O.; Knytl, V.; Najmanová, P.; Steinová, J.; Černík, M.; Filipová, A.; Filip, J.; Cajthaml, T. Combined nano-biotechnology for in-situ remediation of mixed contamination of groundwater by hexavalent chromium and chlorinated solvents. Sci. Total Environ. 2016, 563, 822–834. [Google Scholar] [CrossRef]
  81. Jørgensen, B.B. Mineralization of organic matter in the sea bed—The role of sulphate reduction. Nature 1982, 296, 643–645. [Google Scholar] [CrossRef]
  82. Inglett, K.S.; Bae, H.S.; Aldrich, H.C.; Hatfield, K.; Ogram, A.V. Clostridium chromiireducens sp. nov., isolated from Cr(VI)-contaminated soil. Int. J. Syst. Evol. Microbiol. 2011, 61, 2626–2631. [Google Scholar] [CrossRef]
  83. Battaglia-Brunet, F.; Foucher, S.; Denamur, A.; Ignatiadis, I.; Michel, C.; Morin, D. Reduction of chromate by fixed films of sulfate-reducing bacteria using hydrogen as an electron source. J. Ind. Microbiol. Biotechnol. 2002, 28, 154–159. [Google Scholar] [CrossRef]
  84. Chung, J.; Nerenberg, R.; Rittmann, B.E. Bio-reduction of soluble chromate using a hydrogen-based membrane biofilm reactor. Water Res. 2006, 40, 1634–1642. [Google Scholar] [CrossRef]
  85. McLean, J.S.; Beveridge, T.J.; Phipps, D. Isolation and characterization of a chromium-reducing bacterium from a chromated copper arsenate-contaminated site. Environ. Microbiol. 2000, 2, 611–619. [Google Scholar] [CrossRef]
  86. Mclean, J.; Beveridge, T.J. Chromate Reduction by a Pseudomonad Isolated from a Site Contaminated with Chromated Copper Arsenate. Appl. Environ. Microbiol. 2001, 67, 1076–1084. [Google Scholar] [CrossRef] [Green Version]
  87. Kracke, F.; Vassilev, I.; Krömer, J.O. Microbial Electron Transport and Energy Conservation—The Foundation for Optimizing Bioelectrochemical Systems. Front. Microbiol. 2015, 6, 575. [Google Scholar] [CrossRef]
  88. Cordas, C.M.; Guerra, L.T.; Xavier, C.; Moura, J.J.G. Electroactive biofilms of sulphate reducing bacteria. Electrochim. Acta 2008, 54, 29–34. [Google Scholar] [CrossRef]
  89. Tandukar, M.; Huber, S.J.; Onodera, T.; Pavlostathis, S.G. Biological chromium(VI) reduction in the cathode of a microbial fuel cell. Environ. Sci. Technol. 2009, 43, 8159–8165. [Google Scholar] [CrossRef]
  90. Huang, L.; Chen, J.; Quan, X.; Yang, F. Enhancement of hexavalent chromium reduction and electricity production from a biocathode microbial fuel cell. Bioprocess Biosyst. Eng. 2010, 33, 937–945. [Google Scholar] [CrossRef]
  91. Hsu, L.; Masuda, S.A.; Nealson, K.H.; Pirbazari, M. Evaluation of microbial fuel cell Shewanella biocathodes for treatment of chromate contamination. RSC Adv. 2012, 2, 5844–5855. [Google Scholar] [CrossRef]
  92. Xafenias, N.; Zhang, Y.; Banks, C.J. Enhanced performance of hexavalent chromium reducing cathodes in the presence of Shewanella oneidensis MR-1 and lactate. Environ. Sci. Technol. 2013, 47, 4512–4520. [Google Scholar] [CrossRef]
  93. Li, Y.; Wu, Y.; Puranik, S.; Lei, Y.; Vadas, T.; Li, B. Metals as electron acceptors in single-chamber microbial fuel cells. J. Power Sources 2014, 269, 430–439. [Google Scholar] [CrossRef]
  94. Huang, L.; Wang, Q.; Jiang, L.; Zhou, P.; Quan, X.; Logan, B.E. Adaptively Evolving Bacterial Communities for Complete and Selective Reduction of Cr(VI), Cu(II), and Cd(II) in Biocathode Bioelectrochemical Systems. Environ. Sci. Technol. 2015, 49, 9914–9924. [Google Scholar] [CrossRef]
  95. Xafenias, N.; Zhang, Y.; Banks, C.J. Evaluating hexavalent chromium reduction and electricity production in microbial fuel cells with alkaline cathodes. Int. J. Environ. Sci. Technol. 2015, 12, 2435–2446. [Google Scholar] [CrossRef]
  96. Wu, X.; Zhu, X.; Song, T.; Zhang, L.; Jia, H.; Wei, P. Effect of acclimatization on hexavalent chromium reduction in a biocathode microbial fuel cell. Bioresour. Technol. 2015, 180, 185–191. [Google Scholar] [CrossRef]
  97. Chen, D.; Wang, H.; Yang, K. Effective biodegradation of nitrate, Cr(VI) and p-fluoronitrobenzene by a novel three dimensional bioelectrochemical system. Bioresour. Technol. 2016, 203, 370–373. [Google Scholar] [CrossRef]
  98. Wu, X.; Tong, F.; Yong, X.; Zhou, J.; Zhang, L.; Jia, H.; Wei, P. Effect of NaX zeolite-modified graphite felts on hexavalent chromium removal in biocathode microbial fuel cells. J. Hazard. Mater. 2016, 308, 303–311. [Google Scholar] [CrossRef]
  99. Song, T.; Jin, Y.; Bao, J.; Kang, D.; Xie, J. Graphene/biofilm composites for enhancement of hexavalent chromium reduction and electricity production in a biocathode microbial fuel cell. J. Hazard. Mater. 2016, 317, 73–80. [Google Scholar] [CrossRef]
  100. Huang, L.; Zhou, P.; Quan, X.; Logan, B.E. Removal of binary Cr(VI) and Cd(II) from the catholyte of MFCs and determining their fate in EAB using fluorescence probes. Bioelectrochemistry 2018, 122, 61–68. [Google Scholar] [CrossRef]
  101. Wang, H.; Zhang, S.; Wang, J.; Song, Q.; Zhang, W.; He, Q.; Song, J.; Ma, F. Comparison of performance and microbial communities in a bioelectrochemical system for simultaneous denitrification and chromium removal: Effects of pH. Process Biochem. 2018, 73, 154–161. [Google Scholar] [CrossRef]
  102. Wu, X.; Ren, X.; Owens, G.; Brunetti, G.; Zhou, J.; Yong, X.; Wei, P.; Jia, H. A Facultative Electroactive Chromium(VI)-Reducing Bacterium Aerobically Isolated From a Biocathode Microbial Fuel Cell. Front. Microbiol. 2018, 9, 2883. [Google Scholar] [CrossRef] [Green Version]
  103. Sagar, S.; Dwivedi, A.; Yadav, S.; Tripathi, M.; Kaistha, S.D. Hexavalent chromium reduction and plant growth promotion by Staphylococcus arlettae Strain Cr11. Chemosphere 2012, 86, 847–852. [Google Scholar] [CrossRef]
  104. Jung, R.K.; Cheng, S.; Oh, S.E.; Logan, B.E. Power generation using different cation, anion, and ultrafiltration membranes in microbial fuel cells. Environ. Sci. Technol. 2007, 41, 1004–1009. [Google Scholar]
  105. Malaviya, P.; Singh, A. Bioremediation of chromium solutions and chromium containing wastewaters. Crit. Rev. Microbiol. 2016, 42, 607–633. [Google Scholar] [CrossRef]
  106. Clauwaert, P.; Desloover, J.; Shea, C.; Nerenberg, R.; Boon, N.; Verstraete, W. Enhanced nitrogen removal in bio-electrochemical systems by pH control. Biotechnol. Lett. 2009, 31, 1537–1543. [Google Scholar] [CrossRef]
  107. Liang, P.; Fan, M.; Cao, X.; Huang, X. Evaluation of applied cathode potential to enhance biocathode in microbial fuel cells. J. Chem. Technol. Biotechnol. 2009, 84, 794–799. [Google Scholar] [CrossRef]
  108. Wu, J.; Zhang, J.; Xiao, C. Focus on factors affecting pH, flow of Cr and transformation between Cr(VI) and Cr(III) in the soil with different electrolytes. Electrochim. Acta 2016, 211, 652–662. [Google Scholar] [CrossRef]
  109. Chai, L.; Huang, S.; Yang, Z.; Peng, B.; Huang, Y.; Chen, Y. Cr(VI) remediation by indigenous bacteria in soils contaminated by chromium-containing slag. J. Hazard. Mater. 2009, 167, 516–522. [Google Scholar] [CrossRef]
  110. Kavita, B.; Keharia, H. Reduction of hexavalent chromium by Ochrobactrum intermedium BCR400 isolated from a chromium-contaminated soil. 3 Biotech 2012, 2, 79–87. [Google Scholar] [CrossRef]
  111. Sharma, S.; Adholeya, A. Hexavalent chromium reduction in tannery effluent by bacterial species isolated from tannery effluent contaminated soil. J. Environ. Sci. Technol. 2012, 5, 142–154. [Google Scholar] [CrossRef]
  112. Squillace, P.J.; Scott, J.C.; Moran, M.J.; Nolan, B.T.; Kolpin, D.W. VOCs, pesticides, nitrate, and their mixtures in groundwater used for drinking water in the United States. Environ. Sci. Technol. 2002, 36, 1923–1930. [Google Scholar] [CrossRef]
  113. Santini, M.; Marzorati, S.; Fest-Santini, S.; Trasatti, S.; Cristiani, P. Carbonate scale deactivating the biocathode in a microbial fuel cell. J. Power Sources 2017, 356, 400–407. [Google Scholar] [CrossRef]
  114. Duca, M.; Koper, M.T.M. Powering denitrification: The perspectives of electrocatalytic nitrate reduction. Energy Environ. Sci. 2012, 5, 9726. [Google Scholar] [CrossRef]
  115. Coma, M.; Puig, S.; Pous, N.; Balaguer, M.D.; Colprim, J. Biocatalysed sulphate removal in a BES cathode. Bioresour. Technol. 2013, 130, 218–223. [Google Scholar] [CrossRef]
  116. Bardiya, N.; Bae, J.H. Dissimilatory perchlorate reduction: A review. Microbiol. Res. 2011, 166, 237–254. [Google Scholar] [CrossRef]
  117. Daghio, M.; Vaiopoulou, E.; Patil, S.A.; Suárez-Suárez, A.; Head, I.M.; Franzetti, A.; Rabaey, K. Anodes stimulate anaerobic toluene degradation via sulfur cycling in marine sediments. Appl. Environ. Microbiol. 2016, 82, 297–307. [Google Scholar] [CrossRef]
  118. Zhang, T.; Bain, T.S.; Barlett, M.A.; Dar, S.A.; Snoeyenbos-West, O.L.; Nevin, K.P.; Lovley, D.R. Sulfur oxidation to sulfate coupled with electron transfer to electrodes by Desulfuromonas strain TZ1. Microbiology 2014, 160, 123–129. [Google Scholar] [CrossRef] [Green Version]
  119. Dutta, P.K.; Keller, J.; Yuan, Z.; Rozendal, R.A.; Rabaey, K. Role of sulfur during acetate oxidation in biological anodes. Environ. Sci. Technol. 2009, 43, 3839–3845. [Google Scholar] [CrossRef]
  120. Daghio, M.; Espinoza Tofalos, A.; Leoni, B.; Cristiani, P.; Papacchini, M.; Jalilnejad, E.; Bestetti, G.; Franzetti, A. Bioelectrochemical BTEX removal at different voltages: Assessment of the degradation and characterization of the microbial communities. J. Hazard. Mater. 2018, 341, 120–127. [Google Scholar] [CrossRef]
  121. Wang, C.; Deng, H.; Zhao, F. The Remediation of Chromium (VI)-Contaminated Soils Using Microbial Fuel Cells. Soil Sediment Contam. 2016, 25, 1–12. [Google Scholar] [CrossRef]
  122. Habibul, N.; Hu, Y.; Wang, Y.K.; Chen, W.; Yu, H.Q.; Sheng, G.P. Bioelectrochemical Chromium(VI) Removal in Plant-Microbial Fuel Cells. Environ. Sci. Technol. 2016, 50, 3882–3889. [Google Scholar] [CrossRef]
  123. Guan, C.Y.; Hu, A.; Yu, C.P. Stratified chemical and microbial characteristics between anode and cathode after long-term operation of plant microbial fuel cells for remediation of metal contaminated soils. Sci. Total Environ. 2019, 670, 585–594. [Google Scholar] [CrossRef]
Figure 1. Schematic overview of a BES for Cr(VI) reduction as MFC with energy harvesting or MEC with external supply ([58], modified). In MFCs, oxidation of the electron donor at the anode is coupled with a reduction of species with comparable or higher redox potential at the cathode. The net potential of the MFC, as the sum of anodic and cathodic potentials, is positive. Therefore, spontaneous electron flow from the anode to the cathode occurs. Conversely, in MEC, thanks to external power input to force electron flow, the oxidation of an electron donor at the anode can be coupled with the reduction of lower redox potential species at the cathode. CEM/PEM: Cation/Proton Exchange Membrane. EAB: Electrochemically Active Bacteria
Figure 1. Schematic overview of a BES for Cr(VI) reduction as MFC with energy harvesting or MEC with external supply ([58], modified). In MFCs, oxidation of the electron donor at the anode is coupled with a reduction of species with comparable or higher redox potential at the cathode. The net potential of the MFC, as the sum of anodic and cathodic potentials, is positive. Therefore, spontaneous electron flow from the anode to the cathode occurs. Conversely, in MEC, thanks to external power input to force electron flow, the oxidation of an electron donor at the anode can be coupled with the reduction of lower redox potential species at the cathode. CEM/PEM: Cation/Proton Exchange Membrane. EAB: Electrochemically Active Bacteria
Water 11 02336 g001
Table 1. Standard potentials (Eh0) at 25 °C (V vs. Standard Hydrogen Electrode, SHE) and theoretical potential (Eh) at pH = 7 and 25 °C (V vs. SHE) of Cr(VI) and other selected species of interest for BESs application in Cr(VI) groundwater remediation, sourced from [32,39,59].
Table 1. Standard potentials (Eh0) at 25 °C (V vs. Standard Hydrogen Electrode, SHE) and theoretical potential (Eh) at pH = 7 and 25 °C (V vs. SHE) of Cr(VI) and other selected species of interest for BESs application in Cr(VI) groundwater remediation, sourced from [32,39,59].
ReactionEh0 (V vs. SHE)Eh’ at pH 7 (V vs. SHE)
Cr2O72− + 14H+ + 6e → 2Cr3+ +7H2O1.330.33
HCrO4− + 7H+ + 3e → Cr3+ + 4H2O1.350.35
HCrO4− + 4H+ + 3e → Cr(OH)2+ + 3H2O1.310.76
CrO42− + 4H2O + 3e → Cr(OH)3(s, hydrated) + 5OH−0.130.21
O2 + 4H+ + 4e → H2O1.230.805 (pO2 = 0.2 bar)
NO3 + 6H+ + 5e → 1/2 N2 + 3H2O1.240.71 (pN2 = 0.8 bar)
Fe3+ + e → Fe2+0.770.77
SO42− + 10H+ + 8e → H2S + 4H2O0.30−0.21
HCO3 + 9H+ + 8e → CH4(g) + 3H2O0.20−0.26
2HCO3 + 9H+ + 8e → CH3COO + 4H2O0.19−0.29
2H+ + 2e → H2(g)0−0.41 (pH2 = 1 bar)
MnO2(s) + 4H+ + 2e → Mn2+ + 2H2O1.230.402
Table 2. Summary of key studies regarding Cr(VI) reduction in BESs.
Table 2. Summary of key studies regarding Cr(VI) reduction in BESs.
BES Type 1Anode MaterialAnodic Inoculum/Mediator 2Cathode MaterialCathodic Inoculum/MediatorCathode Potential (V vs. SHE)Initial Cr(VI) (mg/L)Initial pHTest Period
(h) 3
CrVI Removal (%)Rate (mgCrVI/L/h) (Specific Rate (mgCrVI/ gVSS /h)) 4References
2CMFC-PEM, H(b)graphite plateanaerobic sludge. ED: acetategraphite plateCr(VI) enriched denitrifying and anaerobic mixed cultureNA227.2–7.65521000.14 (0.18)[89]
31(0.22)
40(0.36)
630.45 (0.46)
2CMFC, PEM, H(b)graphite plateMFC effluent. ED: acetategraphite plate and granulesAnaerobic mixed culture from contaminated soilNA13≈771003.8 (2)[90]
NA395.3 (2.4)
2CMFC, CEM, TR(b)graphite fibereffluent and acclimated anode from MFCgraphite fiber (C/A = 3)MFC anaerobic effluent0.013205398.56.57[65]
0.038 614.07
NA4073.5758.57
−0.05503NA(5.2)
NA2051004.08 (12.4)
graphite fiber (C/A = 10)4 5.12 (15.5)
graphite fiber (C/A = 20)3 6.8 (20.6)
graphite fiber (C/A = 3)590.253.61 ± 0.1
(11.3 ± 2.2)
graphite felt
(C/A = 3)
76.753.07 ± 0.12
(9.5 ± 1)
graphite granules (C/A = 3)86.53.46 ± 0.09
(5.6 ± 0.6)
2CMFC, CEM, TR(b)graphite brushanaerobic wastewater ED: acetategraphite granulesWWTP primary clarifier effluentNA20724700.6[64]
2CBES (poised cathode), CEM, TR(b)graphite brushgraphite granules−0.45550.43
−0.31000.83
−0.151000.83
0.2430.36
2CMFC, CEM, H(sb)reticulated vitreous carbonShewanella oneidensis MR-1. ED: lactatereticulated vitreous carbonShewanella oneidensis MR1NA2.5 (3 cycles)772 (each cycle)60–1000.02–0.04[91]
Shewanella putrefaciens92–1000.03–0.04
Shewanella amazonensis72–1000.025–0.04
Shewanella sp ANA332–1000.01–0.04
Shewanella loihica44–1000.015–0.04
Shewanella sp MR-420–1000.007–0.04
2CMFC, PEM, H(b)graphite feltS.oneidensis MR-1. ED: lactategraphite feltShewanella oneidensis MR-1NA10 (6 cycles)7300900.1–1.1[92]
2CBES (poised cathode), PEM, H(b)graphite feltShewanella oneidensis MR-1. ED: lactategraphite feltShewanella oneidensis MR1100.5
Shewanella oneidensis MR1; MED: riboflavin130.65
Shewanella oneidensis MR1; ED: lactate−0.32074452.25
SCMFC (b)carbon brushmunicipal wastewater. ED: acetatecarbon cloth with Ptmunicipal wastewater; ED: acetate2.9316.5120890.01[93]
3.03395.70.02
3.131098.80.08
2CMFC, CEM, TR(b)graphite brushMFC anodic effluent. ED: acetategraphite feltWWTP primary clarifier effluent acclimated to Cr(VI)−0.07455.841001.21[94]
NA5 (with Cu(II), Cd(II))390.49
WWTP primary clarifier effluent, acclimated to Cr(VI), Cu(II) and Cd(II)1001.24
2CMFC, PEM, H(b)graphite feltShewanella oneidensisgraphite feltShewanella oneidensisNA10 (8 cycles)884090–1000.9–1.95[95]
2CMFC, PEM, C(b)graphite feltanaerobic sludge. ED: glucosegraphite feltmixed culture from MFC anode (ex situ)NA2072479.30.66[96]
anaerobic digester sludge enriched in presence of Cr(VI) (in situ)20.20.17
SCMEC, Cyl(c)carbon rodanaerobic sludgegraphite feltanaerobic sludgeNA100 16 HRT43.12–96.68 2.69–6.04[97]
100 (200 NO3, 100 p-FNB)41.382.59
100 (200 NO3, 150 p-FNB)49.143.07
100 (200 NO3, 200 p-FNB)55.213.45
100 (200 NO3, 300 p-FNB)58.933.68
2CMFC-PEM, C(b)graphite feltanaerobic sludge. ED: glucosegraphite feltacclimated MFC anode NA207528.31.13 ± 0.01[98]
graphite felt /NaXNA692.76 ± 0.09
graphite felt /NaX-HNO3NA310010.39 ± 0.28
2CMFC-PEM, H(b)graphite feltsewage sludge. ED: glucosegraphite feltacclimated MFCNA4074858.30.49[99]
graphene-modified felt1000.83
2CMFC, CEM, Cyl(b)graphite feltMFC anodic effluent. ED: acetategraphite feltStenotrophomonas sp., S. maltophilia, Serratia marcescens, Achromobacter xylosoxidans−0.04205.8574–832.96–3.32[100]
−0.0520 (20 mg/L Cd(II))63–712.52–2.84
SCMEC, Cyl(c)graphite rodactivated sludgecarbon feltactivated sludgeNA30 (20 mg/L NO3)620 HRT58.960.88[101]
772.651.09
865.080.98
2CMFC-CEM, H(b)graphite feltanaerobic sludgegraphite feltBacillus cereusNA277251002.56[102]
Notes: 1 2CMFC: Double chamber Microbial Fuel Cell. 2CBES-poised cathode: potentiostatically controlled bio-electrochemical systems. SCMFC: Single-Chamber Microbial Fuel Cell. SCMEC: Single-Chamber Microbial Electrolysis Cell. 3CMDC: three-chamber Microbial Desalinization Cell. 3CMFC: three-chamber Microbial Fuel Cell (anode and double cathode). PEM: Proton Exchange Membrane. CEM: Cation Exchange Membrane. BPM: Bipolar membrane. H = H-type reactor. Cyl = Cylindrical reactor. C = cubic chambers reactor. TR = tubular reactor. (b) = batch mode operation. (sb) = semi-batch mode operation, i.e., continuous flow anode chamber and batch cathodic chamber. (c) continuous flow operation 2 ED: Electron Donor. MED: mediator for chemical or biological activity. WWTP (Wastewater Treatment Plant). PPy (polypyrrole). AQS (9,10-anthraquinone-2-sulfonic acid sodium salt). AQDS (anthraquinone-2-sulfonate) 3 HRT: Hydraulic Retention Time. 4 Specific rate normalized by the mass of volatile suspended solids (VSS) in the cathode compartment. NA: no available information.

Share and Cite

MDPI and ACS Style

Beretta, G.; Daghio, M.; Espinoza Tofalos, A.; Franzetti, A.; Mastorgio, A.F.; Saponaro, S.; Sezenna, E. Progress Towards Bioelectrochemical Remediation of Hexavalent Chromium. Water 2019, 11, 2336. https://doi.org/10.3390/w11112336

AMA Style

Beretta G, Daghio M, Espinoza Tofalos A, Franzetti A, Mastorgio AF, Saponaro S, Sezenna E. Progress Towards Bioelectrochemical Remediation of Hexavalent Chromium. Water. 2019; 11(11):2336. https://doi.org/10.3390/w11112336

Chicago/Turabian Style

Beretta, Gabriele, Matteo Daghio, Anna Espinoza Tofalos, Andrea Franzetti, Andrea Filippo Mastorgio, Sabrina Saponaro, and Elena Sezenna. 2019. "Progress Towards Bioelectrochemical Remediation of Hexavalent Chromium" Water 11, no. 11: 2336. https://doi.org/10.3390/w11112336

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