Next Article in Journal
Study on the Preparation and Properties of Talcum-Fly Ash Based Ceramic Membrane Supports
Next Article in Special Issue
Immobilisation and Release of Radical Scavengers on Nanoclays for Chemical Reinforcement of Proton Exchange Membranes
Previous Article in Journal
Reverse Electrodialysis: Co- and Counterflow Optimization of Multistage Configurations for Maximum Energy Efficiency
Previous Article in Special Issue
The Effects of Temperature and Humidity on the Microstructure of Sulfonated Syndiotactic–polystyrene Ionic Membranes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Insights into Advancements and Electrons Transfer Mechanisms of Electrogens in Benthic Microbial Fuel Cells

by
Mohammad Faisal Umar
1,
Syed Zaghum Abbas
2,*,
Mohamad Nasir Mohamad Ibrahim
3,
Norli Ismail
1 and
Mohd Rafatullah
1,*
1
Division of Environmental Technology, School of Industrial Technology, Universiti Sains Malaysia, Penang 11800, Malaysia
2
Biofuels Institute, School of Environment, Jiangsu University, Zhenjiang 212013, China
3
School of Chemical Sciences, Universiti Sains Malaysia, Penang 11800, Malaysia
*
Authors to whom correspondence should be addressed.
Membranes 2020, 10(9), 205; https://doi.org/10.3390/membranes10090205
Submission received: 7 August 2020 / Revised: 19 August 2020 / Accepted: 19 August 2020 / Published: 28 August 2020
(This article belongs to the Special Issue Ionic Conductive Membranes for Fuel Cells)

Abstract

:
Benthic microbial fuel cells (BMFCs) are a kind of microbial fuel cell (MFC), distinguished by the absence of a membrane. BMFCs are an ecofriendly technology with a prominent role in renewable energy harvesting and the bioremediation of organic pollutants through electrogens. Electrogens act as catalysts to increase the rate of reaction in the anodic chamber, acting in electrons transfer to the cathode. This electron transfer towards the anode can either be direct or indirect using exoelectrogens by oxidizing organic matter. The performance of a BMFC also varies with the types of substrates used, which may be sugar molasses, sucrose, rice paddy, etc. This review presents insights into the use of BMFCs for the bioremediation of pollutants and for renewable energy production via different electron pathways.

Graphical Abstract

1. Introduction

Different environmental pollutants, such as organic- and inorganic-based contaminants, remain a severe challenge to the sustainability of water resources [1,2]. This poses a serious threat to living organisms, including human beings and marine organisms [3]. Due to the depletion of natural water resources, there is an imbalance in the natural ecosystem, but simultaneously the commutability of renewable pure water resources has been enhanced. There is a plethora of potential sources of pollution in water bodies (e.g., oceans, lakes, rivers and reservoirs) stemming from human activity, and notably the chemical and oil filtration industries. The chemical substances emitted from these industries contain very harmful and potentially carcinogenic inorganic and organic pollutants [4]. These pollutants have a severe impact on living organisms and pose a serious threat to the environment.
Several techniques exist for the treatment of wastewater prior to irrigation, such as lagoon ponds, constructed wetlands, conventional wastewater treatment plants, membrane bioreactors and membrane filtration. Although these techniques have been shown to be effective, disadvantages remain, i.e., they require a large area for operation, along with high economic stability [5]. Recently, a novel approach was introduced for the treatment of wastewater: the microbial fuel cell. Microbial fuel cells (MFCs) are devices which utilize microbial activity to produce electricity from chemical energy stored in an organic substrate. Thus, MFCs are a promising technique for wastewater bioremediation and for generating electricity in an economical way.
Organic pollutant compounds are oxidized by microorganisms and the transfer of electrons to the anode of the MFC via exoelectrogens [6,7]. A new type of MFC, the benthic microbial fuel cell (BMFC), was designed to generate electricity from organic matter present in wastewater. As a result, like with MFCs, chemical energy is converted into electrical energy with exoelectrogens working as a catalyst, i.e., electrons (e) and protons (H+) are released. In this way, a potential difference exists between the anode and cathode. Here, we present information regarding recent developments using exoelectrogens on the anode by direct and indirect processes.

2. Benthic Microbial Fuel Cell (BMFC)

There is a need for sustainable and clean energy sources to meet growing energy demands. In 2014, the global percentage of electricity generated via the consumption of fossil fuels was 66%; however, only 11% of this was utilized together with renewable energy [8,9]. Organic substrates are used as bio sediments, and they protect the microbial ecosystem in various regions and provide a suitable environment for the bioremediation of accumulated pollutants via the electron donor–acceptor mechanism [10]. Currently, physiochemical processes, such as dredging, ozonation and electrochemical degradation, are used for the bioremediation of pollutants. These techniques are effective but require a lot of energy and are costly, limiting their application. Usually, the accumulation of reductive substances and the lack of electron acceptors are the main limitations for the remediation of sediment under anaerobic conditions.
In recent years, microbial fuel cells (MFC) have been considered as an alternative, cheap approach to the bioremediation of toxic organic pollutants via power generation. Recently, BMFCs have attracted the attention of many researchers due to their nonaggressive and easily controllable nature. BMFCs consist of an anode, which is embedded in organic matter, and a cathode, which is placed in the overlying water. The air diffuser provides a constant supply of oxygen which plays a vital role in the transfer of electrons and protons from the anode to cathode via an external circuit, where electrons react with oxygen and produce water [11,12].
Reimers et al. [13] were the first to employ BMFCs; their approach included a platinum mesh for the anode and carbon fiber for the cathode. A unique feature of the BMFC is its membrane-less assembly; this is possible thanks to the boundary organic substrate used as a substrate, which itself acts as a pseudo membrane. Nowadays, many researchers are working on improving ecofriendly systems, including BMFCs [14]. The prototype of a double chamber BMFC is shown in Figure 1.
An air cathode in the overlaying water connected with a benthic-integrating anode is the most common BMFC model. In a saline environment, conductivity is normally high, so the overpotential limits the BMFC performance; this is not the case in freshwater [15]. Under the latter scenario, the efficiency of the anode decreases because of anodic contamination, i.e., the accumulation of waste substrate in the anodic region. BMFCs are usually restricted in terms of the proximity of the electrode by the naturally forming spatial separation of oxic and anoxic zones [16]. The tubular air cathode designs along with the cathodic fabric assembly structure suggest that only low-cost fabric would separate the electrodes. In this configuration, the cathode catalytic layer was exposed to air and would allow a hydrogen oxidation reaction [17]. However, as this setup requires long tubes for air exposure, the BMFC’s setup cannot operate in deep-water environment. If the BMFC can adapt the cathode carbon cloth, then embedded cathode in the organic substrate can also be used optionally [2]. In the simple design of the BMFC, though, electrodes can be constructed from both graphite felt or carbon cloth.

3. Degradation of Organic Matter by BMFC

Like bio-electrochemical systems, BMFCs too have been shown to boost the organic compounds biodegradation, i.e., total petroleum hydrocarbons, total organic carbon, ignition loss and polycyclic aromatic hydrocarbons present in the wastewater, as shown in Figure 2. BMFC takes some time for the formation of a biofilm on the anode, which is the main requirement for the removal of the organic contents [18,19]. The anodic biofilm consists of two types of bacteria, the fermentative bacteria and the exoelectrogens. Fermentative bacteria are primarily involved in the complex organic matter hydrolysis and transform the products of hydrolysis into ethanol, H2, volatile organic acids and CO2 by acid-forming fermentation [20]. Ethanol, into which lactic acid can easily be converted, is volatile and readily escapes, allowing the reaction to proceed easily. CO2 is the other product, but is weakly acidic and even more volatile than ethanol. H2 is a substrate for methanogens and sulfate reducers, which keep the concentration of hydrogen low and favor the production of such an energy-rich compound, but hydrogen gas at a fairly high concentration can nevertheless be formed.
The metabolites of fermentative bacteria used by electrogenic bacteria as substrates, which produce electrons, CO2 and protons by oxidation, are shown in Equation (1). The protons are shifted to the overlying cathodic water and transfer few electrons towards the anode, which can be seen in Equation (2). These electrons are passed to the cathode through an external circuit and a redox reaction occurs that generates protons and dissolves oxygen, as mentioned in Equation (3) [21]. The existence of these electrodes has established a new microbial mechanism for metabolism, and to some degree it alters anodic microbial communities too. Recently, it has been reported that BMFCs alone cannot efficiently remove the organic pollutants. Wu et al. [18] reported that zero-valent iron (ZVI) has a high reducing ability (E0 = −0.44 V) and could react with the oxidizing contaminants. The hydroxyl radical formed through this method is a very durable oxidative degradation of bio-refractory organics, which allows for the common use of ZVI technology in the treatment of dyes, complexing agents, chlorinated organic compounds and preservatives. ZVI can also alter the metabolic pathways and redox capacity, regulate acidification and promote extracellular electron transfer. Estevezcanales et al. [22] cultivated Geobacter sulfurreducense with an iron-free substratum and found an abruptly reduced cytochrome c, which showed a limited capacity of outer membrane electrons transport. However, using ZVI alone, the desired effect cannot be guaranteed, particularly the final removal of certain refractory contaminants. The combined use of ZVI and BMFC technologies offers an enhanced substitute approach for eliminating organic contaminants.
Anode: (oxidation)
a ( OP ) +   bH 2 O cCO 2 +   ne + dH +
Cathode: (reduction)
eO 2 +   dH + + ne bH 2 O
Overall reaction: (redox reaction)
a ( OP ) +   eO 2 cCO 2 +   bH 2 O
a = number of organic pollutants (OP) molecules, b = number of water molecules, c = number of carbon dioxide molecules, d = number of protons, e = number of oxygen molecules and n = number of electrons.
The removal of organic contaminants from BMFC is the foremost priority for organic contents remediation. Many hydrocarbons, such as those consisting of nitro and chlorine aromatic compounds, can be employed as substrates in BMFC. For bioremediation, these compounds need bioreduction [23]. The amalgamation of bioremediation and the electrochemical system forms a synergistic connection among electrodes and bacteria and enables the bioreduction of perchloroethane and polycyclic aromatic hydrocarbons. The in-situ generation of oxygen and hydrogen can be employed for intermediates reduction. The energy efficiency and removal of these organic compounds can be upgraded by direct electron transfer to electrodes from exoelectrogens or the inclusion of dechlorinating species [24]. There is a proportional relationship among power production and the degradation of these organic compounds. This closed-circuit BMFC creates the optimum environment for the degradation of organic compounds. This system could have a negative impact on BMFC microbes if not used properly. During the remediation of the organic compound in BMFC, some common issue are encountered, such as cathodic pH becoming alkaline and anodic pH becoming acidic via water electrolysis [25]. Unequal nutrients distribution in the chamber, like nitrate and phosphate, accumulating in the cathode chamber and ammonium accumulating in the anode chamber are other issues encountered during the remediation. These issues not only effect the performance of BMFC but also the biological clogging. These issue can be resolved by reversal of electrodes polarity and with proper water circulation. The degradation of organic compounds is also influenced by the competitive reactions with nitrate and sulphate [26].

4. Electron Transfer Mechanism by Electrogens

The electrons transmission mechanism is essential in order to acquire a flawless knowledge for the application of BMFC at a large scale. In the anodic chamber of BMFC, organic substrates are reduced by microbes and transfer electrons to anodes, from where the electrons move to the cathode through external circuit to generate electricity [27]. Earlier, the microbes were exploited in the anodic chamber, but recently microbes are also exploited as biocathodes in the cathodic region to assist electrons transmission to the terminal electron acceptor (TEA) [28,29]. The power density, current density and coulombic efficiency can be measured by electron transfer rate. If the electrons transfer rate is higher than the electrons passing through the external circuit, more coulombic efficiency, power density and current density will be measured, leading to higher voltage production. The harvested bioenergy produced by the electron transfer towards electrodes from the respiration chain of electrogens is known as a new BMFC technology [30]. There are two means of electrons transfer in BMFC occupied by microbes: (i) direct electron transfer (direct contact between the microbes and the electrode surface) and (ii) indirect electron transfer (through the so-called electron mediators), as shown in Figure 3.
Recently, the application of electro-autotrophs in the Bioelectrochemical Systems (BES) has attracted the attention of researchers. The exoelectrogens use the electrodes or extracellular insoluble mineral as terminal electron acceptor (TEA), while electro-autotrophs accept the electrons from electrodes or solid compounds for CO2 reduction and produce multi-carbon compounds [31]. Gregory et al. first studied the electro-autotrophy in the Geobacter, which is a model exoelectrogen [32]. Most exoelectrogens are iron-oxidizing bacteria, which led to the hypothesis that dissimilatory iron-reducing bacteria can only accept the electrons from a cathode. Indeed, Mariprofundus ferrooxydans PV-1, Acidithiobacillus ferrooxidans and Rhodopseudomonas palustris have been selected as electro-autotrophs [33,34,35]. Furthermore, Methanobacterium archaeon strain IM1 and chemolithoautotrophic archea Methanococcus maripaludis were purified for electromethanogenesis with an electron donor (metallic iron) [36]. Many acetogenic bacteria like Sporumosa acidovorans, Sporomusa silvacetica, Sporomusa sphaeroides, Sporumosa malonica, Moorella thermoacetica, Sporomusa ovate, Clostridium aceticum and Clostridium ljungdahlii can also accept electrons from the cathode and reduce CO2 to organic acids [37]. Some sulphate-reducing autotrophs are believed to accept the electrons from cathode and generate hydrogen (H2) by reducing sulphate [38]. The cathodic electron consumption by bacteria causes anaerobic microbial-induced corrosion (MIC). The electro-autotrophs generate the corrosive hydrogen sulphide that results in chemically induced iron corrosion. The electro-autotrophs also stimulate the induced electrochemical corrosion by using cathodic hydrogen, which is generated by iron–water contact [39]. The benthic microbial fuel cells (BMFCs) were constructed for anaerobic exoelectrogenic enrichment, which separates the electrotrophic bacteria by opposing the anode to bio-cathode [40]. Recently, an MFC was developed initially with heterotrophic conditions that later alter with autotrophic conditions. After five batches of cultivation, the nonelectrochemical bacteria is dispersed into the liquid medium and only electro-autotrophs bacteria (Geobacter) were abundant in the MFC [41]. This electro-autotrophic process promotes the growth of exoelectrogens on the electrodes and reduces the number of nonelectrobiochemical bacteria, which finally increases the MFC’s efficiency. The electro-autotrophic enrichment of the bio-cathode offers a simplified approach to purify the bio-chemical from various inoculum sources. Initially, bacteria are grown heterotrophically on fructose, glycerol and glucose, followed by acclimation to the medium, and CO2 was provided as the sole electron acceptor [42]. The conventional cathode causes corrosion, denaturation and toxicity of material, but the bio-cathode is very cost-effective. The microbes must be chosen based on their capability to shift from heterotrophic to autotrophic metabolism. This pathway may help us to understand the metabolic pathways of different electron donors or acceptor microbes that have formed on bio-cathodes [43]. For the production of valuable organic and fuel commodities, pure culture was used because the diversified electro-autotrophs uptake the electrons from the negatively poised cathode for CO2 reduction with heavier coulombic efficiencies. The mixed cultures primarily generate the complex products and acetates, which maintains the microbial metabolism. The surfeit of products was generated by employing a viable BES system with pure culture of Clostridium ljungdahlii. Overall, though, very little research has been focused on the electro-autotrophs, particularly the electrons transfer pathways from cathode to bacteria and their applications.

4.1. Direct Electron Transfer

Electrons should interact between the outer membrane of the microbes and the electrode. The biofilm or electrically conductive nanowires (pili and flagella) were found over the surface of the anode formed by electrogens [44]. The transmission of electrons takes place by direct interaction without any external mediator through an external membrane’s cytochromes, nanowires and electron transport proteins in exchange with the microbial membranes. The external membrane’s cytochromes are bonded with nanowires and allow electrogens to use an electrode as an electron acceptor. Furthermore, the direct electron transfer mechanism fully depends on the electron transport proteins, and they play a crucial role in electron transfer from cytoplasm to mitochondrial membrane. The drawback of this mechanism is the very poor electron transfer rate, because the active sites of electron transmission are deeply embedded within the proteins [45]. Recently, many electrochemical bacteria like Shewanella and Geobacter nanowires have been folded for better electrons transmission [46,47]. For effective and fast electron transfer (coulombic efficiency), the nanowires form an electroactive layer instead of a normal single layer. Geobacter species are diverse in their current production ability; Geobacter hydrogenophilus and Geobacter metallireducens produced higher current densities (0.2 mAcm−2) than Geobacter bremensis, Geobacter chapellei, Geobacter humireducens, Geobacter uraniireducen and Geobacter bemidjiensis, which produced much lower current densities (0.05 mAcm−2) [48]. Some electrogesns reported direct electron transfer to electrodes, such as Geobacter sulfurreducens [49], Rhodopseudomonas palustris [50], Anaeromyxobacter dehalogenansc [51], Geobacter lovleyi [52], Pseudomonas aeruginosa [53], Thermincola potens [54], Shewanella oneidensis [55], Geothrix fermentans [56], Thermincola carboxydophila [57], Shewanella putrefaciens [58], and Escherichia coli [59].
Much less is known about direct electron transfer pathways in the electro-autotrophic bacteria. From the experiments, it is confirmed that the Fe species uptake the electrons secreted by the cathodic biofilm. It is also ventured that c-type cytochromes, which are crucial constituents of Fe extracellular electron uptake, also play a vital part in the electron transmission from cathode to electro-autotrophs [60]. In the light of this hypothesis, the metaproteomics and metagenomics of the diversified microbial community inhibit the self-regenerating biocathode’s effect whereby CO2 is reduced via c-type cytochromes directly acquiring electrons from the Chromatiaceae family and other proteins related with Fe(II) oxidation [61]. The Fe(0)-corroding sulphate reducing microbes (SRM) could also uptake the electrons [62], and this discovery paved the way for scientists to use these microbes in biocathodic BES employments. So, this negative metabolic character can be turned into a sustainable positive biotechnological solution. Up to now, though, only some pure SRM cultures are used as electro-autotrophs. The cathodic biofilm of Desulfovibrio desulfuricans ATCC 27774 exhibited electro-autotrophic characteristics at an employed cathodic potential (Ecath) of −0.169 V vs. SHE. After 20 days, lactate was supplemented as the carbon source, not CO2, and a stable negative current was measured [63]. Consequently, other species (Desulfovibrio caledoniensis and Desulfovibrio paquesii) of the genus Desulfovibrio were used for H2 and cathodic current generation, employing lactate or bicarbonate as the carbon source and Ecath that enabled abiotic H2 evolution [64].
The pure cultures of Desulfovibrio piger and Desulfosporosinus orientis displayed the electro-autotrophic properties at Ecath = −0.31 V vs. SHE, which has a higher positive potential than the neutral redox potential of H2 evolution (E0′H+/H2 = −0.41 V vs. SHE) and gaseous CO2 supplemented as an inorganic source [65]. Desulfovibrio piger (SRM Deltaproteobacterium) is a H2-oxidizing, Gram-negative, nonspore-forming electro-autotroph. It could oxidize organic matter, like lactate, pyruvate ethanol and, partly, acetate. Before this, its autotrophic metabolism effect on CO2, was not reported for other Desulfovibrio species. Desulfosporosinus orientis (SRM class Clostridia) is an acetogenic, capable of executing anaerobic sulfate respiration, and is a spore-forming electro-autotroph. The broad range of energy sources, such as pyruvate, ethanol, formate, methanol, H2, Fatty acids, lactate CO and CO2, can be used by D. orientis [66]. It can use various TEAs, such as sulphite, sulphate, sulphur dioxide and thiosulfate [67].
In BES, for the first time Desulfopila corrodens strain IS4 was identified as an Fe(0)-corroding SRM [38]. By using an electron donor (metallic iron), this Deltaproteobacterium (Gram-negative) was quarantined from marine sediment. This strain performs very fast hydrogen generation and sulphate reduction by consuming iron as an energy source as compared to orthodox hydrogen-foraging Desulfovibrio species. In BES, by using CO2 as the growth substrate at Ecath = −0.4 V vs. SHE, direct electron uptake was accomplished [38]. Currently, Desulfobacterium autotrophicum HRM2 (sulphate reducing bacteria) is being reported as an electro-autotroph at Ecath = −0.5 V vs. SHE. This Deltaproteobacterium, secluded from marine mud, is a fully SRM oxidizer having both directional pathways (Wood-Ljungdahl) and relating to the c-Cyt rich group [68]. D. autotrophicum HRM2 showed a high coulombic efficiency (83 ± 6%) and a capacity for acetate bio-electro synthesis [69].

4.2. Indirect Electron Transfer

Indirect electron transfer does not require direct physical interaction between the microbes and electrons acceptors. The small molecules and soluble mediator are involved in the inducement of this electron’s transfer mechanism. In this mechanism, the electrons mediator enters into the microbes, where the electrons are extracted by a metabolic reaction of electrogens, and finally these electrons are transferred to an anode [70]. Initially, at the first BMFC operative phase, the presence of electron mediators was considered as important. The electron mediators auxiliary in the BMFC anodic chamber are produced by electrogens. Several types of species had been investigated, as the synthesis of self-mediators known as endo-electrogens mediators, such as phenazine and pyocyanin, could be secreted by Shewanella and Pseudomonas species [71]. The potential differences between several electron mediators and redox proteins were reported in many studies, which significantly affects the electron transfer efficiency of different species [72]. However, the tendency of electrons transfer is affected by different chemical compounds known as exoelectricigens mediators, such as anthracenedione, thionine, neutral red, humic acid, riboflavin and methylene blue [73,74,75]; both exo-electrogens and endo-electrogens are shown in Table 1. These electrogens are exploited to transfer the electrons from inside of the cell towards the electrode, and different microbes have a different capability to transfer electrons from cell to electrode.
Very little is known about the electron uptake by acetogens from the cathode. Currently, by using a genetic system, it is being confirmed that Clostridium ljungdahlii (Gram-positive) exhibits protons pumps that cause proton motive force, which is necessary for its growth with CO2 as a carbon source [111]. This gives clues about the energy conservation mechanism in the electro-autotrophic acetogens. In Clostridium ljungdahlii, the electron uptake mechanism is differently predicted, because it cannot synthesize quinones or c-type cytochromes [112]. By using genetic toolbox, the properties and electron uptake pathways of Clostridium ljungdahlii could be clearer, and also give information about the electron uptake pathways of many Gram-positive bacteria. The genomic sequence of acetogenic Sporomusa ovata (Gram-negative) is available now. Genes coding for type IV pili and c-type cytochromes are present in the genomic sequence, which are the two main parts of the extracellular electron transfer mechanism [113]. The c-type cytochromes are a precarious factor for the extracellular electron transfer mechanism in both electrotrophs and electrogenic types. In Geobacter spp., pili type IV are long strings that exhibit the metal-like conduction of long-range electron transfer. The gene coding for Ubiquinone also present in the genome of Sporomusa ovate is also crucial for the electron transfer pathway [114]. Sporomusa ovate has many extracellular electron transfer components, which proves that the electron uptake mechanisms of Sporomusa ovata are similar to those of other electrotrophic and electrigenic bacteria. Sporomusa sphaeroides-related acetogens showed direct electron transfer mechanisms. This showed that Gram-negative acetogens could use this strategy of electron transfer in different environments [115].

5. Performance of BMFC Affected by Organic Substrate

In BMFC, the chemical reaction is replaced by a microbial reaction where the organic substrates are utilized as fuel for feeding the microbes and generating renewable energy. All these microorganisms that grow are nourished by varieties of substrates, which include simple carbohydrates or polysaccharides, amino acids, organic acids, cellulose and lignocellulose [95]. Marine sediments and aqueous ones were also employed in BMFC as a substrate [116]. The substrate not only facilitates the microbes in producing the biofilm on the surface of the anode, but is also designed to increase the performance of the BMFC by producing higher coulombic efficiency and power density [117]. Moreover, the diverse substrate processes fully depend on the biodegradability factor. The power density of BMFC is directly proportional to the quantity of organic contents in the organic substrate and the biodegradation by electrogens of the microorganism [118]. The mechanism of organic substrate degradation through electrogens using BMFC is shown in Figure 4.
Hassan et al. [119] studied the different organic substrates (glucose, fructose and sucrose) used in BMFC. Wang et.al. [120] developed a BMFC to generate a power density of about 12.7 mW/m2 using an electron-mediating agent at pH 4, with the help of acidophilic bacterium, Acidiphilium cryptum, utilizing glucose as the organic substrate. The pure bacterial strain Brevibacillus borstelensis STRI1 produced a power density of about 188.5 mW/m2 by using sugarcane molasses as the organic substrate [119]. The rice straw was also used as an organic substrate to generate a power density of about 293.33 ± 7.89 mW/m2 [121]. The existing literature reveals the different kinds of organic waste being used as organic substrates, with their corresponding capacities for power density generation by electrogens, as shown in Table 2.
In the BMFC, various kinds of substrates could be employed; these substrates can be starch, petroleum-based compounds, cysteine, glucose, dairy-based, acetate, molasses, glutamic acid, food-based wastewater, river water and vegetable-based. The substrate selection is based on their biodegradability behaviors. The power production by BMFC depends upon the degradation rate by the bacteria and the quantity of organic contents in the substrates [147]. In BMFCs, there is a continuous generation of power which is impeded by access to nutrients in the anodic media. The nutrients in BMFCs are regularly supplied with fresh matter from the decay of microbes and animals, giving the BMFC an indefinite life span in theory [148]. In the BMFC, one biodegradable fuel was also the bio-battery, but with this the power generation ultimately drops with time. Some substrates only support a single form of organic material. Different types of chitin were also used in BMFC anode as substrates. Chitin 80 and chitin 20 produced optimum power of about 84 ± 10 and 76 ± 25 mW/m2, respectively. The internal resistances of chitin 80 and chitin 20 were 650 ± 130 and 1300 ± 440, respectively. The electricity production could be enhanced by using substrates of precise size, and slowly degradable substrates. The substrates of precise size enhance the degradation surface area, and the slowly degradable substrates enhance the power production duration [146].

6. Conclusions

BMFC is a novel bio-technique that may be a potential solution to the two main problems, namely pollutants bioremediation and sustainable energy production. These BMFCs will open new possibilities for sustainable, cost-effective and controllable ways to generate power and bioremediate toxic pollutants. For power generation, there are two main routes of electron transfer: direct electron (physical contact between electrogens and anode) and indirect electron (conductive pili and flagella) transfer from the electrogens towards the anode of BMFC. The performance of BMFC depends on the use of different organic matters as the substrate. The novel BMFC technology will be encouraging for in situ pollutants bioremediation. The challenges of BMFCs will be addressed jointly by the efforts of scientists from many fields, such as environmental sciences, biotechnology, electrochemistry, electrical engineering, biology and material sciences.

Author Contributions

Conceptualization, M.F.U. and M.R.; writing—original draft preparation, M.F.U. and S.Z.A.; writing—review and editing, M.N.M.I., N.I. and M.R.; supervision, M.N.M.I., N.I., and M.R.; funding acquisition, M.R. All authors have read and agreed to the published version of the manuscript.

Funding

The authors are grateful to the Universiti Sains Malaysia for supporting this work through RUI grant (1001/PTEKIND/8011044).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Grey, D.; Garrick, D.; Blackmore, D.; Kelman, J.; Muller, M.; Sadoff, C. Water security in one blue planet: Twenty-first century policy challenges for science. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2013, 371, 20120406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Li, W.-W.; Yu, H.-Q. Stimulating sediment bioremediation with benthic microbial fuel cells. Biotechnol. Adv. 2015, 33, 1–12. [Google Scholar] [CrossRef] [PubMed]
  3. Khan, M.E.; Khan, M.M.; Min, B.-K.; Cho, M.H. Microbial fuel cell assisted band gap narrowed TiO2 for visible light-induced photocatalytic activities and power generation. Sci. Rep. 2018, 8, 1–12. [Google Scholar] [CrossRef] [Green Version]
  4. Khan, M.E.; Han, T.H.; Khan, M.M.; Karim, M.R.; Cho, M.H. Environmentally sustainable fabrication of Ag@g–C3N4 nanostructures and their multifunctional efficacy as antibacterial agents and photocatalysts. ACS Appl. Nano. Mater. 2018, 1, 2912–2922. [Google Scholar] [CrossRef]
  5. Khan, M.E.; Khan, M.M.; Cho, M.H. Environmentally sustainable biogenic fabrication of AuNP decorated-graphitic gC3N4 nanostructures towards improved photoelectrochemical performances. RSC Adv. 2018, 8, 13898–13909. [Google Scholar] [CrossRef] [Green Version]
  6. Yaqoob, A.A.; Khatoon, A.; Mohd Setapar, S.H.; Umar, K.; Parveen, T.; Mohamad Ibrahim, M.N.; Ahmad, A.; Rafatullah, M. Outlook on the role of microbial fuel cells in remediation of environmental pollutants with electricity generation. Catalysts 2020, 10, 819. [Google Scholar] [CrossRef]
  7. Yaqoob, A.A.; Mohamad Ibrahim, M.N.; Rafatullah, M.; Chua, Y.S.; Ahmad, A.; Umar, K. Recent advances in anodes for microbial fuel cells: An overview. Materials 2020, 13, 2078. [Google Scholar] [CrossRef]
  8. Liew, K.B.; Daud, W.R.W.; Ghasemi, M.; Leong, J.X.; Lim, S.S.; Ismail, M. Non-Pt catalyst as oxygen reduction reaction in microbial fuel cells: A review. Int. J. Hydrogen Energy 2014, 39, 4870–4883. [Google Scholar] [CrossRef]
  9. Abbas, S.Z.; Rafatullah, M.; Ismail, N.; Nastro, R.A. Enhanced bioremediation of toxic metals and harvesting electricity through sediment microbial fuel cell. Int. J. Energy Res. 2017, 41, 2345–2355. [Google Scholar] [CrossRef]
  10. Martins, G.; Peixoto, L.; Brito, A.G.; Nogueira, R. Phosphorus–iron interaction in sediments: Can an electrode minimize phosphorus release from sediments? Rev. Environ. Sci. Biotechnol. 2014, 13, 265–275. [Google Scholar] [CrossRef] [Green Version]
  11. Li, H.; Tian, Y.; Qu, Y.; Qiu, Y.; Liu, J.; Feng, Y. A pilot-scale benthic microbial electrochemical system (BMES) for enhanced organic removal in sediment restoration. Sci. Rep. 2017, 7, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Abbas, S.Z.; Rafatullah, M.; Ismail, N.; Syakir, M.I. The behaviour of membrane less sediment microbial fuel cell in the terms of bioremediation and power generation. Malays. J. Microbiol. 2018, 14, 108–112. [Google Scholar]
  13. Reimers, C.E.; Girguis, P.; Stecher, H.A.; Tender, L.M.; Ryckelynck, N.; Whaling, P. Microbial fuel cell energy from an ocean cold seep. Geobiology 2006, 4, 123–136. [Google Scholar] [CrossRef]
  14. Nielsen, M.E.; Reimers, C.E.; White, H.K.; Sharma, S.; Girguis, P.R. Sustainable energy from deep ocean cold seeps. Energy Environ. Sci. 2008, 1, 584–593. [Google Scholar] [CrossRef]
  15. Hong, S.W.; Chang, I.S.; Choi, Y.S.; Kim, B.H.; Chung, T.H. Responses from freshwater sediment during electricity generation using microbial fuel cells. Bioprocess Biosyst. Eng. 2009, 32, 389–395. [Google Scholar] [CrossRef]
  16. Rezaei, F.; Richard, T.L.; Brennan, R.A.; Logan, B.E. Substrate-enhanced microbial fuel cells for improved remote power generation from sediment-based systems. Environ. Sci. Technol. 2007, 41, 4053–4058. [Google Scholar] [CrossRef]
  17. Yuan, Y.; Zhou, S.; Zhuang, L. A new approach to in situ sediment remediation based on air-cathode microbial fuel cells. J. Soils. Sediments 2010, 10, 1427–1433. [Google Scholar] [CrossRef]
  18. Wu, M.; Xu, X.; Lu, K.; Li, X. Effects of the presence of nanoscale zero-valent iron on the degradation of polychlorinated biphenyls and total organic carbon by sediment microbial fuel cell. Sci. Total. Environ. 2019, 656, 39–44. [Google Scholar] [CrossRef]
  19. Abbas, S.Z.; Rafatullah, M.; Ismail, N.; Syakir, M.I. A review on sediment microbial fuel cells as a new source of sustainable energy and heavy metal remediation: Mechanisms and future prospective. Int. J. Energy Res. 2017, 41, 1242–1264. [Google Scholar] [CrossRef]
  20. Xu, X.; Zhao, Q.; Wu, M.; Ding, J.; Zhang, W. Biodegradation of organic matter and anodic microbial communities analysis in sediment microbial fuel cells with/without Fe (III) oxide addition. Bioresour. Technol. 2017, 225, 402–408. [Google Scholar] [CrossRef]
  21. Khan, M.E.; Khan, M.M.; Cho, M.H. Recent progress of metal–graphene nanostructures in photocatalysis. Nanoscale 2018, 10, 9427–9440. [Google Scholar] [CrossRef] [PubMed]
  22. Estevez-Canales, M.; Kuzume, A.; Borjas, Z.; Füeg, M.; Lovley, D.; Wandlowski, T.; Esteve-Núñez, A. A severe reduction in the cytochrome C content of Geobacter sulfurreducens eliminates its capacity for extracellular electron transfer. Environ. Microbiol. Rep. 2015, 7, 219–226. [Google Scholar] [CrossRef] [Green Version]
  23. Khan, M.E.; Khan, M.M.; Cho, M.H. Green synthesis, photocatalytic and photoelectrochemical performance of an Au–Graphene nanocomposite. RSC Adv. 2015, 5, 26897–26904. [Google Scholar] [CrossRef]
  24. Khan, M.E.; Khan, M.M.; Cho, M.H. Biogenic synthesis of a Ag–graphene nanocomposite with efficient photocatalytic degradation, electrical conductivity and photoelectrochemical performance. New J. Chem. 2015, 39, 8121–8129. [Google Scholar] [CrossRef]
  25. Song, X.; Wang, W.; Cao, X.; Wang, Y.; Zou, L.; Ge, X.; Zhao, Y.; Si, Z.; Wang, Y. Chlorella vulgaris on the cathode promoted the performance of sediment microbial fuel cells for electrogenesis and pollutant removal. Sci. Total. Environ. 2020, 728, 138011. [Google Scholar] [CrossRef] [PubMed]
  26. Li, W.; Quan, X.; Chen, L.; Zheng, Y. Application of slow-release carbon sources embedded in polymer for stable and extended power generation in microbial fuel cells. Chemosphere 2020, 244, 125515. [Google Scholar] [CrossRef]
  27. Song, H.-L.; Zhu, Y.; Li, J. Electron transfer mechanisms, characteristics and applications of biological cathode microbial fuel cells—A mini review. Arab. J. Chem. 2019, 12, 2236–2243. [Google Scholar] [CrossRef] [Green Version]
  28. Yang, Z.; Yang, A. Modelling the impact of operating mode and electron transfer mechanism in microbial fuel cells with two-species anodic biofilm. Biochem. Eng. J. 2020, 107560. [Google Scholar] [CrossRef]
  29. Abbas, S.Z.; Rafatullah, M.; Ismail, N.; Shakoori, F.R. Electrochemistry and microbiology of microbial fuel cells treating marine sediments polluted with heavy metals. RSC Adv. 2018, 8, 18800–18813. [Google Scholar] [CrossRef] [Green Version]
  30. Jothinathan, D.; Mylsamy, P.; Bruno, L.B. Electricigens: Role and prominence in microbial fuel cell performance. In Microbial Fuel Cell Technology for Bioelectricity; Springer: Berlin/Heidelberg, Germany, 2018; pp. 169–185. [Google Scholar]
  31. Tremblay, P.-L.; Angenent, L.T.; Zhang, T. Extracellular electron uptake: Among autotrophs and mediated by surfaces. Trends Biotechnol. 2017, 35, 360–371. [Google Scholar] [CrossRef]
  32. Lee, C.S.; Robinson, J.; Chong, M.F. A review on application of flocculants in wastewater treatment. Process Saf. Environ. Prot. 2014, 92, 489–508. [Google Scholar] [CrossRef]
  33. Carbajosa, S.; Malki, M.; Caillard, R.; Lopez, M.F.; Palomares, F.J.; Martín-Gago, J.A.; Rodríguez, N.; Amils, R.; Fernández, V.M.; De Lacey, A.L. Electrochemical growth of Acidithiobacillus ferrooxidans on a graphite electrode for obtaining a biocathode for direct electrocatalytic reduction of oxygen. Biosens. Bioelectron. 2010, 26, 877–880. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Summers, Z.M.; Gralnick, J.A.; Bond, D.R. Cultivation of an obligate Fe (II)-oxidizing lithoautotrophic bacterium using electrodes. mBio 2013, 4, e00420-12. [Google Scholar] [CrossRef] [Green Version]
  35. Bose, A.; Gardel, E.J.; Vidoudez, C.; Parra, E.; Girguis, P.R. Electron uptake by iron-oxidizing phototrophic bacteria. Nat. Commun. 2014, 5, 1–7. [Google Scholar] [CrossRef] [PubMed]
  36. Beese-Vasbender, P.F.; Grote, J.-P.; Garrelfs, J.; Stratmann, M.; Mayrhofer, K.J. Selective microbial electrosynthesis of methane by a pure culture of a marine lithoautotrophic archaeon. Bioelectrochemistry 2015, 102, 50–55. [Google Scholar] [CrossRef]
  37. Aryal, N.; Tremblay, P.-L.; Lizak, D.M.; Zhang, T. Performance of different Sporomusa species for the microbial electrosynthesis of acetate from carbon dioxide. Bioresour. Technol. 2017, 233, 184–190. [Google Scholar] [CrossRef]
  38. Beese-Vasbender, P.F.; Nayak, S.; Erbe, A.; Stratmann, M.; Mayrhofer, K.J. Electrochemical characterization of direct electron uptake in electrical microbially influenced corrosion of iron by the lithoautotrophic SRB Desulfopila corrodens strain IS4. Electrochim. Acta 2015, 167, 321–329. [Google Scholar] [CrossRef]
  39. Li, Y.; Xu, D.; Chen, C.; Li, X.; Jia, R.; Zhang, D.; Sand, W.; Wang, F.; Gu, T. Anaerobic microbiologically influenced corrosion mechanisms interpreted using bioenergetics and bioelectrochemistry: A review. J. Mater. Sci. Technol. 2018, 34, 1713–1718. [Google Scholar] [CrossRef]
  40. Pisciotta, J.M.; Zaybak, Z.; Call, D.F.; Nam, J.-Y.; Logan, B.E. Enrichment of microbial electrolysis cell biocathodes from sediment microbial fuel cell bioanodes. Appl. Environ. Microbiol. 2012, 78, 5212–5219. [Google Scholar] [CrossRef] [Green Version]
  41. Chen, K.-T.; Bai, M.-D.; Wu, S.-I.; Chen, C.-C.; Lu, W.-J.; Wan, H.-P.; Huang, C. Electro-autotrophs induced the growth of exoelectrogens on the anode in a microbial fuel cell. Biochem. Eng. J. 2019, 141, 29–34. [Google Scholar] [CrossRef]
  42. Peng, T.; Feng, C.; Hu, W.; Chen, N.; He, Q.; Dong, S.; Xu, Y.; Gao, Y.; Li, M. Treatment of nitrate-contaminated groundwater by heterotrophic denitrification coupled with electro-autotrophic denitrifying packed bed reactor. Biochem. Eng. J. 2018, 134, 12–21. [Google Scholar] [CrossRef]
  43. Jabeen, G.; Farooq, R. Bio-electrochemical synthesis of commodity chemicals by autotrophic acetogens utilizing CO2 for environmental remediation. J. Biosci. 2016, 41, 367–380. [Google Scholar] [CrossRef] [PubMed]
  44. Reguera, G.; McCarthy, K.D.; Mehta, T.; Nicoll, J.S.; Tuominen, M.T.; Lovley, D.R. Extracellular electron transfer via microbial nanowires. Nature 2005, 435, 1098–1101. [Google Scholar] [CrossRef]
  45. Zhao, F.; Slade, R.C.T.; Varcoe, J.R. Techniques for the study and development of microbial fuel cells: An electrochemical perspective. Chem. Soc. Rev. 2009, 38, 1926–1939. [Google Scholar] [CrossRef] [Green Version]
  46. Malvankar, N.S.; Lovley, D.R. Microbial nanowires for bioenergy applications. Curr. Opin. Biotechnol. 2014, 27, 88–95. [Google Scholar] [CrossRef] [PubMed]
  47. Abbas, S.Z.; Rafatullah, M.; Khan, M.A.; Siddiqui, M.R. Bioremediation and Electricity Generation by Using Open and Closed Sediment Microbial Fuel Cells. Front. Microbiol. 2019, 9, 3348. [Google Scholar] [CrossRef] [Green Version]
  48. Rotaru, A.-E.; Woodard, T.L.; Nevin, K.P.; Lovley, D.R. Link between capacity for current production and syntrophic growth in Geobacter species. Front. Microbiol. 2015, 6, 744. [Google Scholar] [CrossRef] [Green Version]
  49. Xing, D.; Zuo, Y.; Cheng, S.; Regan, J.M.; Logan, B.E. Electricity generation by Rhodopseudomonas palustris DX-1. Environ. Sci. Technol. 2008, 42, 4146–4151. [Google Scholar] [CrossRef]
  50. Bond, D.R.; Lovley, D.R. Evidence for involvement of an electron shuttle in electricity generation by Geothrix fermentans. Appl. Environ. Microbiol. 2005, 71, 2186–2189. [Google Scholar] [CrossRef] [Green Version]
  51. Strycharz, S.M.; Woodard, T.L.; Johnson, J.P.; Nevin, K.P.; Sanford, R.A.; Löffler, F.E.; Lovley, D.R. Graphite electrode as a sole electron donor for reductive dechlorination of tetrachlorethene by Geobacter lovleyi. Appl. Environ. Microbiol. 2008, 74, 5943–5947. [Google Scholar] [CrossRef] [Green Version]
  52. Virdis, B.; Rabaey, K.; Yuan, Z.; Keller, J. Microbial fuel cells for simultaneous carbon and nitrogen removal. Water Res. 2008, 42, 3013–3024. [Google Scholar] [CrossRef]
  53. Rabaey, K.; Boon, N.; Höfte, M.; Verstraete, W. Microbial phenazine production enhances electron transfer in biofuel cells. Environ. Sci. Technol. 2005, 39, 3401–3408. [Google Scholar] [CrossRef]
  54. Wrighton, K.C.; Thrash, J.C.; Melnyk, R.A.; Bigi, J.P.; Byrne-Bailey, K.G.; Remis, J.P.; Schichnes, D.; Auer, M.; Chang, C.J.; Coates, J.D. Evidence for direct electron transfer by a Gram-positive bacterium isolated from a microbial fuel cell. Appl. Environ. Microbiol. 2011, 77, 7633–7639. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Baron, D.; LaBelle, E.; Coursolle, D.; Gralnick, J.A.; Bond, D.R. Electrochemical measurement of electron transfer kinetics by Shewanella oneidensis MR-1. J. Biol. Chem. 2009, 284, 28865–28873. [Google Scholar] [CrossRef] [Green Version]
  56. Titov, D.V.; Cracan, V.; Goodman, R.P.; Peng, J.; Grabarek, Z.; Mootha, V.K. Complementation of mitochondrial electron transport chain by manipulation of the NAD+/NADH ratio. Science 2016, 352, 231–235. [Google Scholar] [CrossRef] [Green Version]
  57. Mathis, B.J.; Marshall, C.W.; Milliken, C.E.; Makkar, R.S.; Creager, S.E.; May, H.D. Electricity generation by thermophilic microorganisms from marine sediment. Appl. Microbiol. Biotecnol. 2008, 78, 147–155. [Google Scholar] [CrossRef] [PubMed]
  58. Lovley, D.R. The microbe electric: Conversion of organic matter to electricity. Curr. Opin. Biotechnol. 2008, 19, 564–571. [Google Scholar] [CrossRef]
  59. Zhang, T.; Cui, C.; Chen, S.; Yang, H.; Shen, P. The direct electrocatalysis of Escherichia coli through electroactivated excretion in microbial fuel cell. Electrochem. Commun. 2008, 10, 293–297. [Google Scholar] [CrossRef]
  60. Sydow, A.; Krieg, T.; Mayer, F.; Schrader, J.; Holtmann, D. Electroactive bacteria—Molecular mechanisms and genetic tools. Appl. Microbiol. Biotechnol. 2014, 98, 8481–8495. [Google Scholar] [CrossRef]
  61. Wang, Z.; Leary, D.H.; Malanoski, A.P.; Li, R.W.; Hervey, W.J.; Eddie, B.J.; Tender, G.S.; Yanosky, S.G.; Vora, G.J.; Tender, L.M. A previously uncharacterized, nonphotosynthetic member of the Chromatiaceae is the primary CO2-fixing constituent in a self-regenerating biocathode. Appl. Environ. Microbiol. 2015, 81, 699–712. [Google Scholar] [CrossRef] [Green Version]
  62. Xu, D.; Gu, T. Bioenergetics explains when and why more severe MIC pitting by SRB can occur. In Proceedings of the NACE International Corrosion 2011 Conference and Expo, Houston, TX, USA, 13–17 March 2011. [Google Scholar]
  63. Cordas, C.M.; Moura, J.J.G. Sulphate reducing bacteria-electroactive biofilm formation. Int. J. Med. Biol. Front. 2011, 17, 295–312. [Google Scholar]
  64. Arends, J.B.; Patil, S.A.; Roume, H.; Rabaey, K. Continuous long-term electricity-driven bioproduction of carboxylates and isopropanol from CO2 with a mixed microbial community. J. CO2 Util. 2017, 20, 141–149. [Google Scholar] [CrossRef] [Green Version]
  65. de Campos Rodrigues, T.; Rosenbaum, M.A. Microbial electroreduction: Screening for new cathodic biocatalysts. ChemElectroChem 2014, 1, 1916–1922. [Google Scholar] [CrossRef]
  66. Qian, Z.; Tianwei, H.; Mackey, H.R.; van Loosdrecht, M.C.; Guanghao, C. Recent advances in dissimilatory sulfate reduction: From metabolic study to application. Water Res. 2019, 150, 162–181. [Google Scholar] [CrossRef] [PubMed]
  67. Berlendis, S.; Ranchou-Peyruse, M.; Fardeau, M.-L.; Lascourreges, J.-F.; Joseph, M.; Ollivier, B.; Aüllo, T.; Dequidt, D.; Magot, M.; Ranchou-Peyruse, A. Desulfotomaculum aquiferis sp. nov. and Desulfotomaculum profundi sp. nov., isolated from a deep natural gas storage aquifer. Int. J. Syst. Evol. Microbiol. 2016, 66, 4329–4338. [Google Scholar] [CrossRef] [PubMed]
  68. Jiang, Y.; May, H.D.; Lu, L.; Liang, P.; Huang, X.; Ren, Z.J. Carbon dioxide and organic waste valorization by microbial electrosynthesis and electro-fermentation. Water Res. 2019, 149, 42–55. [Google Scholar] [CrossRef]
  69. Jiang, Y.; Zeng, R.J. Bidirectional extracellular electron transfers of electrode-biofilm: Mechanism and application. Bioresour. Technol. 2019, 271, 439–448. [Google Scholar] [CrossRef] [PubMed]
  70. He, L.; Du, P.; Chen, Y.; Lu, H.; Cheng, X.; Chang, B.; Wang, Z. Advances in microbial fuel cells for wastewater treatment. Renew. Sustain. Energy Rev. 2017, 71, 388–403. [Google Scholar] [CrossRef]
  71. Bilal, M.; Wang, S.; Iqbal, H.M.N.; Zhao, Y.; Hu, H.; Wang, W.; Zhang, X. Metabolic engineering strategies for enhanced shikimate biosynthesis: Current scenario and future developments. Appl. Microbiol. Biotecnol. 2018, 102, 7759–7773. [Google Scholar] [CrossRef]
  72. Li, Y.; Marshall, A.; Gostomski, P.A. Gaseous pollutant treatment and electricity generation in microbial fuel cells (MFCs) utilising redox mediators. Rev. Environ. Sci. Bio/Technol. 2014, 13, 35–51. [Google Scholar] [CrossRef]
  73. Rahimnejad, M.; Najafpour, G.D.; Ghoreyshi, A.A.; Talebnia, F.; Premier, G.C.; Bakeri, G.; Kim, J.R.; Oh, S.-E. Thionine increases electricity generation from microbial fuel cell using Saccharomyces cerevisiae and exoelectrogenic mixed culture. J. Microbiol. 2012, 50, 575–580. [Google Scholar] [CrossRef] [PubMed]
  74. Thygesen, A.; Poulsen, F.W.; Min, B.; Angelidaki, I.; Thomsen, A.B. The effect of different substrates and humic acid on power generation in microbial fuel cell operation. Bioresour. Technol. 2009, 100, 1186–1191. [Google Scholar] [CrossRef] [PubMed]
  75. Rahimnejad, M.; Najafpour, G.D.; Ghoreyshi, A.A.; Shakeri, M.; Zare, H. Methylene blue as electron promoters in microbial fuel cell. Int. J. Hydrogen Energy 2011, 36, 13335–13341. [Google Scholar] [CrossRef]
  76. Fernando, E.; Keshavarz, T.; Kyazze, G. Enhanced bio-decolourisation of acid orange 7 by Shewanella oneidensis through co-metabolism in a microbial fuel cell. Int. Biodeterior. Biodegrad. 2012, 72, 1–9. [Google Scholar] [CrossRef]
  77. Ringeisen, B.R.; Henderson, E.; Wu, P.K.; Pietron, J.; Ray, R.; Little, B.; Biffinger, J.C.; Jones-Meehan, J.M. High Power Density from a Miniature Microbial Fuel Cell Using Shewanella oneidensis DSP10. Environ. Sci. Technol. 2006, 40, 2629–2634. [Google Scholar] [CrossRef] [Green Version]
  78. Menicucci, J.; Beyenal, H.; Marsili, E.; Veluchamy, R.A.; Demir, G.; Lewandowski, Z. Procedure for Determining Maximum Sustainable Power Generated by Microbial Fuel Cells. Environ. Sci. Technol. 2006, 40, 1062–1068. [Google Scholar] [CrossRef]
  79. Boon, N.; De Maeyer, K.; Höfte, M.; Rabaey, K.; Verstraete, W. Use of Pseudomonas species producing phenazine-based metabolites in the anodes of microbial fuel cells to improve electricity generation. Appl. Microbiol. Biotechnol. 2008, 80, 985–993. [Google Scholar]
  80. Zhang, M.; Ma, Z.; Zhao, N.; Zhang, K.; Song, H. Increased power generation from cylindrical microbial fuel cell inoculated with P. aeruginosa. Biosens. Bioelectron. 2019, 141, 111394. [Google Scholar] [CrossRef]
  81. Takeuchi, Y.; Khawdas, W.; Aso, Y.; Ohara, H. Microbial fuel cells using Cellulomonas spp. with cellulose as fuel. J. Biosci. Bioeng. 2017, 123, 358–363. [Google Scholar] [CrossRef]
  82. Masuda, M.; Freguia, S.; Wang, Y.-F.; Tsujimura, S.; Kano, K. Flavins contained in yeast extract are exploited for anodic electron transfer by Lactococcus lactis. Bioelectrochemistry 2010, 78, 173–175. [Google Scholar] [CrossRef] [Green Version]
  83. 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]
  84. Ringeisen, B.R.; Ray, R.; Little, B. A miniature microbial fuel cell operating with an aerobic anode chamber. J. Power Sources 2007, 165, 591–597. [Google Scholar] [CrossRef]
  85. 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]
  86. Leung, K.M.; Wanger, G.; El-Naggar, M.Y.; Gorby, Y.; Southam, G.; Lau, W.M.; Yang, J. Shewanella oneidensis MR-1 bacterial nanowires exhibit p-type, tunable electronic behavior. Nano Lett. 2013, 13, 2407–2411. [Google Scholar] [CrossRef]
  87. Keller, K.L.; Rapp-Giles, B.J.; Semkiw, E.S.; Porat, I.; Brown, S.D.; Wall, J.D. New model for electron flow for sulfate reduction in Desulfovibrio alaskensis G20. Appl. Environ. Microbiol. 2014, 80, 855–868. [Google Scholar] [CrossRef] [Green Version]
  88. Qiao, Y.; Li, C.M.; Bao, S.-J.; Lu, Z.; Hong, Y. Direct electrochemistry and electrocatalytic mechanism of evolved Escherichia coli cells in microbial fuel cells. Chem. Comm. 2008, 11, 1290–1292. [Google Scholar] [CrossRef]
  89. Marshall, C.W.; May, H.D. Electrochemical evidence of direct electrode reduction by a thermophilic Gram-positive bacterium, Thermincola ferriacetica. Energy. Environ. Sci. 2009, 2, 699–705. [Google Scholar] [CrossRef]
  90. Carmona-Martínez, A.A.; Harnisch, F.; Kuhlicke, U.; Neu, T.R.; Schröder, U. Electron transfer and biofilm formation of Shewanella putrefaciens as function of anode potential. Bioelectrochemistry 2013, 93, 23–29. [Google Scholar] [CrossRef]
  91. Thrash, J.C.; Van Trump, J.I.; Weber, K.A.; Miller, E.; Achenbach, L.A.; Coates, J.D. Electrochemical stimulation of microbial perchlorate reduction. Environ. Sci. Technol. 2007, 41, 1740–1746. [Google Scholar] [CrossRef] [Green Version]
  92. Kalathil, S.; Pant, D. Nanotechnology to rescue bacterial bidirectional extracellular electron transfer in bioelectrochemical systems. RSC Adv. 2016, 6, 30582–30597. [Google Scholar] [CrossRef] [Green Version]
  93. Daniel, D.K.; Mankidy, B.D.; Ambarish, K.; Manogari, R. Construction and operation of a microbial fuel cell for electricity generation from wastewater. Int. J. Hydrogen Energy. 2009, 34, 7555–7560. [Google Scholar] [CrossRef]
  94. Zhi-Dan, L.I.U.; Jing, L.; Zhu-Wei, D.U.; Li, H.-R. Construction of sugar-based microbial fuel cells by dissimilatory metal reduction bacteria. Chin. J. Biotechnol. 2006, 22, 131–137. [Google Scholar]
  95. Liu, L.; Li, F.-b.; Feng, C.-h.; Li, X.-Z. Microbial fuel cell with an azo-dye-feeding cathode. Appl. Microbiol. Biotechnol. 2009, 85, 175–183. [Google Scholar] [CrossRef] [PubMed]
  96. Hassan, S.H.A.; Kim, Y.S.; Oh, S.-E. Power generation from cellulose using mixed and pure cultures of cellulose-degrading bacteria in a microbial fuel cell. Enzyme Microb. Technol. 2012, 51, 269–273. [Google Scholar] [CrossRef]
  97. De Vet, S.J.; Rutgers, R. From waste to energy: First experimental bacterial fuel cells onboard the international space station. Microgravity Sci. Technol. 2007, 19, 225–229. [Google Scholar] [CrossRef]
  98. Zheng, X.; Nirmalakhandan, N. Cattle wastes as substrates for bioelectricity production via microbial fuel cells. Biotechnol. Lett. 2010, 32, 1809–1814. [Google Scholar] [CrossRef]
  99. Marsili, E.; Baron, D.B.; Shikhare, I.D.; Coursolle, D.; Gralnick, J.A.; Bond, D.R. Shewanella secretes flavins that mediate extracellular electron transfer. Proc. Natl. Acad. Sci. USA 2008, 105, 3968–3973. [Google Scholar] [CrossRef] [Green Version]
  100. Ali, N.; Anam, M.; Yousaf, S.; Maleeha, S.; Bangash, Z. Characterization of the electric current generation potential of the pseudomonas aeruginosa using glucose, fructose, and sucrose in double chamber microbial fuel cell. Iran. J. Biotechnol. 2017, 15, 216. [Google Scholar] [CrossRef] [Green Version]
  101. Khawdas, W.; Watanabe, K.; Karatani, H.; Aso, Y.; Tanaka, T.; Ohara, H. Direct electron transfer of Cellulomonas fimi and microbial fuel cells fueled by cellulose. J. Biosci. Bioeng. 2019, 128, 593–598. [Google Scholar] [CrossRef]
  102. Powell, E.E.; Mapiour, M.L.; Evitts, R.W.; Hill, G.A. Growth kinetics of Chlorella vulgaris and its use as a cathodic half cell. Bioresour. Technol. 2009, 100, 269–274. [Google Scholar] [CrossRef]
  103. Cournet, A.; Délia, M.-L.; Bergel, A.; Roques, C.; Bergé, M. Electrochemical reduction of oxygen catalyzed by a wide range of bacteria including Gram-positive. Electrochem. Commun. 2010, 12, 505–508. [Google Scholar] [CrossRef] [Green Version]
  104. Raghavulu, S.V.; Goud, R.K.; Sarma, P.N.; Mohan, S.V. Saccharomyces cerevisiae as anodic biocatalyst for power generation in biofuel cell: Influence of redox condition and substrate load. Bioresour. Technol. 2011, 102, 2751–2757. [Google Scholar] [CrossRef] [PubMed]
  105. Erable, B.; Vandecandelaere, I.; Faimali, M.; Delia, M.-L.; Etcheverry, L.; Vandamme, P.; Bergel, A. Marine aerobic biofilm as biocathode catalyst. Bioelectrochemistry 2010, 78, 51–56. [Google Scholar] [CrossRef] [Green Version]
  106. Nguyen, T.A.; Lu, Y.; Yang, X.; Shi, X. Carbon and Steel Surfaces Modified by Leptothrix discophora SP-6: Characterization and Implications. Environ. Sci. Technol. 2007, 41, 7987–7996. [Google Scholar] [CrossRef] [PubMed]
  107. Wu, S.J.; Hu, Z.H.; Zhang, L.L.; Yu, X.; Chen, J.M. A novel dichloromethane-degrading Lysinibacillus sphaericus strain wh22 and its degradative plasmid. Appl. Mcrobiol. Biotechnol. 2009, 82, 731–740. [Google Scholar] [CrossRef]
  108. Xu, S.; Liu, H. New exoelectrogen Citrobacter sp. SX-1 isolated from a microbial fuel cell. J. App. Microbiol. 2011, 111, 1108–1115. [Google Scholar] [CrossRef] [PubMed]
  109. Inglesby, A.E.; Beatty, D.A.; Fisher, A.C. Rhodopseudomonas palustris purple bacteria fed Arthrospira maxima cyanobacteria: Demonstration of application in microbial fuel cells. RSC. Adv. 2012, 2, 4829–4838. [Google Scholar] [CrossRef]
  110. Wu, C.; Liu, X.-W.; Li, W.-W.; Sheng, G.-P.; Zang, G.-L.; Cheng, Y.-Y.; Shen, N.; Yang, Y.-P.; Yu, H.-Q. A white-rot fungus is used as a biocathode to improve electricity production of a microbial fuel cell. Appl. Energy 2012, 98, 594–596. [Google Scholar] [CrossRef]
  111. Huang, H.; Chai, C.; Yang, S.; Jiang, W.; Gu, Y. Phage serine integrase-mediated genome engineering for efficient expression of chemical biosynthetic pathway in gas-fermenting Clostridium ljungdahlii. Metab. Eng. 2019, 52, 293–302. [Google Scholar] [CrossRef]
  112. Klask, C.-M.; Kliem-Kuster, N.; Molitor, B.; Angenent, L.T. Nitrate Feed Improves Growth and Ethanol Production of Clostridium ljungdahlii with CO2 and H2, but Results in Stochastic Inhibition Events. Front. Microbiol. 2019. [Google Scholar] [CrossRef]
  113. Mateos, R.; Escapa, A.; Vanbroekhoven, K.; Patil, S.A.; Moran, A.; Pant, D. Microbial electrochemical technologies for CO2 and its derived products valorization. In Microbial Electrochemical Technology; Elsevier: Amsterdam, The Netherlands, 2019; pp. 777–796. [Google Scholar]
  114. Kato, S.; Yumoto, I.; Kamagata, Y. Isolation of acetogenic bacteria that induce biocorrosion by utilizing metallic iron as the sole electron donor. Appl. Environ. Microbiol. 2015, 81, 67–73. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Krishna, K.V.; Swathi, K.; Hemalatha, M.; Mohan, S.V. Bioelectrocatalyst in Microbial Electrochemical Systems and Extracellular Electron Transport. In Microbial Electrochemical Technology; Elsevier: Amsterdam, The Netherlands, 2019; pp. 117–141. [Google Scholar]
  116. Reimers, C.E.; Tender, L.M.; Fertig, S.; Wang, W. Harvesting Energy from the Marine Sediment−Water Interface. Environ. Sci. Technol. 2001, 35, 192–195. [Google Scholar] [CrossRef]
  117. Chae, K.-J.; Choi, M.-J.; Lee, J.-W.; Kim, K.-Y.; Kim, I.S. Effect of different substrates on the performance, bacterial diversity, and bacterial viability in microbial fuel cells. Bioresour. Technol. 2009, 100, 3518–3525. [Google Scholar] [CrossRef] [PubMed]
  118. Borole, A.P.; O’Neill, H.; Tsouris, C.; Cesar, S. A microbial fuel cell operating at low pH using the acidophile Acidiphilium cryptum. Biotechnol. Lett. 2008, 30, 1367–1372. [Google Scholar] [CrossRef] [PubMed]
  119. Hassan, S.H.A.; Zohri, A.A.; Kassim, R.M.F. Electricity generation from sugarcane molasses using microbial fuel cell technologies. Energy 2019, 178, 538–543. [Google Scholar] [CrossRef]
  120. Wang, Z.; Lee, T.; Lim, B.; Choi, C.; Park, J. Microbial community structures differentiated in a single-chamber air-cathode microbial fuel cell fueled with rice straw hydrolysate. Biotechnol. Biofuels 2014, 7, 9. [Google Scholar] [CrossRef] [Green Version]
  121. Hassan, S.H.A.; Gad El-Rab, S.M.F.; Rahimnejad, M.; Ghasemi, M.; Joo, J.-H.; Sik-Ok, Y.; Kim, I.S.; Oh, S.-E. Electricity generation from rice straw using a microbial fuel cell. Int. J. Hydrogen Energy 2014, 39, 9490–9496. [Google Scholar] [CrossRef]
  122. Rezaei, F.; Xing, D.; Wagner, R.; Regan, J.M.; Richard, T.L.; Logan, B. E Simultaneous Cellulose Degradation and Electricity Production by Enterobacter cloacae in a Microbial Fuel Cell. Appl. Environ. Microbiol. 2009, 75, 3673. [Google Scholar] [CrossRef] [Green Version]
  123. Manohar, A.K.; Mansfeld, F. The internal resistance of a microbial fuel cell and its dependence on cell design and operating conditions. Electrochim. Acta 2009, 54, 1664–1670. [Google Scholar] [CrossRef]
  124. Jung, S.; Regan, J.M. Comparison of anode bacterial communities and performance in microbial fuel cells with different electron donors. Appl. Microbiol. Biotechnol. 2007, 77, 393–402. [Google Scholar] [CrossRef]
  125. Zou, Y.; Xiang, C.; Yang, L.; Sun, L.-X.; Xu, F.; Cao, Z. A mediatorless microbial fuel cell using polypyrrole coated carbon nanotubes composite as anode material. Int. J. Hydrogen Energy 2008, 33, 4856–4862. [Google Scholar] [CrossRef]
  126. Mohan, Y.; Manoj Muthu Kumar, S.; Das, D. Electricity generation using microbial fuel cells. Int. J. Hydrogen Energy 2008, 33, 423–426. [Google Scholar] [CrossRef]
  127. Ren, Z.; Steinberg, L.M.; Regan, J.M. Electricity production and microbial biofilm characterization in cellulose-fed microbial fuel cells. Water. Sci. Technol. 2008, 58, 617–622. [Google Scholar] [CrossRef] [PubMed]
  128. Hassan, S.H.A.; Van Ginkel, S.W.; Kim, S.-M.; Yoon, S.-H.; Joo, J.-H.; Shin, B.-S.; Jeon, B.-H.; Bae, W.; Oh, S.-E. Isolation and characterization of Acidithiobacillus caldus from a sulfur-oxidizing bacterial biosensor and its role in detection of toxic chemicals. J. Microbiol. Methods 2010, 82, 151–155. [Google Scholar] [CrossRef] [PubMed]
  129. Shen, H.-B.; Yong, X.-Y.; Chen, Y.-L.; Liao, Z.-H.; Si, R.-W.; Zhou, J.; Wang, S.-Y.; Yong, Y.-C.; OuYang, P.-K.; Zheng, T. Enhanced bioelectricity generation by improving pyocyanin production and membrane permeability through sophorolipid addition in Pseudomonas aeruginosa-inoculated microbial fuel cells. Bioresour. Technol. 2014, 167, 490–494. [Google Scholar] [CrossRef] [PubMed]
  130. Jafary, T.; Rahimnejad, M.; Ghoreyshi, A.A.; Najafpour, G.; Hghparast, F.; Daud, W.R.W. Assessment of bioelectricity production in microbial fuel cells through series and parallel connections. Energy Convers. Manag. 2013, 75, 256–262. [Google Scholar] [CrossRef]
  131. Akman, D.; Cirik, K.; Ozdemir, S.; Ozkaya, B.; Cinar, O. Bioelectricity generation in continuously-fed microbial fuel cell: Effects of anode electrode material and hydraulic retention time. Bioresour. Technol. 2013, 149, 459–464. [Google Scholar] [CrossRef]
  132. Mäkinen, A.E.; Lay, C.-H.; Nissilä, M.E.; Puhakka, J.A. Bioelectricity production on xylose with a compost enrichment culture. Int. J. Hydrogen Energy. 2013, 38, 15606–15612. [Google Scholar] [CrossRef]
  133. Aldrovandi, A.; Marsili, E.; Stante, L.; Paganin, P.; Tabacchioni, S.; Giordano, A. Sustainable power production in a membrane-less and mediator-less synthetic wastewater microbial fuel cell. Bioresour. Technol. 2009, 100, 3252–3260. [Google Scholar] [CrossRef] [Green Version]
  134. Dumas, C.; Basseguy, R.; Bergel, A. Microbial electrocatalysis with Geobacter sulfurreducens biofilm on stainless steel cathodes. Electrochim. Acta 2008, 53, 2494–2500. [Google Scholar] [CrossRef] [Green Version]
  135. Catal, T.; Li, K.; Bermek, H.; Liu, H. Electricity production from twelve monosaccharides using microbial fuel cells. J. Power Sources 2008, 175, 196–200. [Google Scholar] [CrossRef]
  136. Rismani-Yazdi, H.; Christy, A.D.; Dehority, B.A.; Morrison, M.; Yu, Z.; Tuovinen, O.H. Electricity generation from cellulose by rumen microorganisms in microbial fuel cells. Biotechnol. Bioeng. 2007, 97, 1398–1407. [Google Scholar] [CrossRef] [PubMed]
  137. Logan, B.; Cheng, S.; Watson, V.; Estadt, G. Graphite fiber brush anodes for increased power production in air-cathode microbial fuel cells. Environ. Sci. Technol. 2007, 41, 3341–3346. [Google Scholar] [CrossRef] [PubMed]
  138. Kim, J.R.; Jung, S.H.; Regan, J.M.; Logan, B.E. Electricity generation and microbial community analysis of alcohol powered microbial fuel cells. Bioresour. Technol. 2007, 98, 2568–2577. [Google Scholar] [CrossRef] [PubMed]
  139. Kargi, F.; Eker, S. Electricity generation with simultaneous wastewater treatment by a microbial fuel cell (MFC) with Cu and Cu–Au electrodes. J. Chem. Technol. Biotechnol. 2007, 82, 658–662. [Google Scholar] [CrossRef]
  140. Logan, B.E.; Murano, C.; Scott, K.; Gray, N.D.; Head, I.M. Electricity generation from cysteine in a microbial fuel cell. Water Res. 2005, 39, 942–952. [Google Scholar] [CrossRef]
  141. Niessen, J.; Schröder, U.; Scholz, F. Exploiting complex carbohydrates for microbial electricity generation–a bacterial fuel cell operating on starch. Electrochem. Commun. 2004, 6, 955–958. [Google Scholar] [CrossRef]
  142. Kalleary, S.; Abbas, F.M.; Ganesan, A.; Meenatchisundaram, S.; Srinivasan, B.; Packirisamy, A.S.B.; krishnan Kesavan, R.; Muthusamy, S. Biodegradation and bioelectricity generation by microbial desalination cell. Int. Biodeterior. Biodegrad. 2014, 92, 20–25. [Google Scholar] [CrossRef]
  143. Jayashree, C.; Arulazhagan, P.; Kumar, S.A.; Kaliappan, S.; Yeom, I.T.; Banu, J.R. Bioelectricity generation from coconut husk retting wastewater in fed batch operating microbial fuel cell by phenol degrading microorganism. Biomass. Bioenerg. 2014, 69, 249–254. [Google Scholar] [CrossRef]
  144. Nimje, V.R.; Chen, C.-Y.; Chen, H.-R.; Chen, C.-C.; Huang, Y.M.; Tseng, M.-J.; Cheng, K.-C.; Chang, Y.-F. Comparative bioelectricity production from various wastewaters in microbial fuel cells using mixed cultures and a pure strain of Shewanella oneidensis. Bioresour. Technol. 2012, 104, 315–323. [Google Scholar] [CrossRef]
  145. Kouzuma, A.; Kasai, T.; Nakagawa, G.; Yamamuro, A.; Abe, T.; Watanabe, K. Comparative metagenomics of anode-associated microbiomes developed in rice paddy-field microbial fuel cells. PLoS ONE 2013, 8, e77443. [Google Scholar] [CrossRef] [PubMed]
  146. Gurav, R.; Bhatia, S.K.; Choi, T.-R.; Jung, H.-R.; Yang, S.-Y.; Song, H.-S.; Park, Y.-L.; Han, Y.-H.; Park, J.-Y.; Kim, Y.-G. Chitin biomass powered microbial fuel cell for electricity production using halophilic Bacillus circulans BBL03 isolated from sea salt harvesting area. Bioelectrochemistry 2019, 130, 107329. [Google Scholar] [CrossRef] [PubMed]
  147. Pushkar, P.; Mungray, A.K. Exploring the use of 3 dimensional low-cost sugar-urea carbon foam electrode in the benthic microbial fuel cell. Renew. Energy 2020, 147, 2032–2042. [Google Scholar] [CrossRef]
  148. Juan, J.G.; Keegan, G.C.; Marcus, O.G.; Sage, E.R.; Peter, R.G.; Michael, A.C. Benthic microbial fuel cells: Long-term power sources for wireless marine sensor networks. In Proceedings of the SPIE 7666, Sensors, and Command, Control, Communications, and Intelligence (C3I) Technologies for Homeland Security and Homeland Defense IX, Orlando, FL, USA, 5 May 2010. [Google Scholar] [CrossRef]
Figure 1. General prototype scheme of a benthic microbial fuel cell.
Figure 1. General prototype scheme of a benthic microbial fuel cell.
Membranes 10 00205 g001
Figure 2. Overview of organic pollutants removal by benthic microbial fuel cell.
Figure 2. Overview of organic pollutants removal by benthic microbial fuel cell.
Membranes 10 00205 g002
Figure 3. Proposed electron transfer mechanisms utilized direct electrons (a,b) and indirect electrons transfer (c) through electrogens using benthic microbial fuel cell.
Figure 3. Proposed electron transfer mechanisms utilized direct electrons (a,b) and indirect electrons transfer (c) through electrogens using benthic microbial fuel cell.
Membranes 10 00205 g003
Figure 4. Representation of power generation by using organic contents as substrates by electrogens in a benthic microbial fuel cell.
Figure 4. Representation of power generation by using organic contents as substrates by electrogens in a benthic microbial fuel cell.
Membranes 10 00205 g004
Table 1. Performance of BMFC configuration through exoelectrogens and endoelectrogens with respect to power density.
Table 1. Performance of BMFC configuration through exoelectrogens and endoelectrogens with respect to power density.
MicroorganismsExternal MediatorPower Density
(mW m−2)
ConfigurationsType of Electrons Transfer MechanismsReferences
Exoelectrogens microorganisms
Shewanella oneidensis strain 140631–amino–2–Napthol>40Single chamberDirect transfer[76]
Shewanella oneidensisAnthraquinone–2,6–disulfonate (AQDS)24Double chamberDirect transfer[77]
Klebsiella pneumoniaeHNQ as mediator biomineralized manganese as electron acceptor__Direct transfer[78]
Pseudomonas speciesphenazine–1–carboxamide__Indirect transfer[79]
Pseudomonas aeruginosaphenazine compounds3322 ± 38Single chamberDirect transfer[80]
Cellulomonas fimianthraquinone–2,6–disulfonate38.7Double chamberDirect transfer[81]
Lactococcus lactisRiboflavin, flavins_Double chamberDirect transfer[82]
Geobacter sulfurreducenscCytochrome z, type IV pili3147Double chamberDirect transfer[83]
Shewanella oneidensis DsP10Anthraquinone–2,6–disulfonate (AQDS)5000Double chamberDirect transfer[77]
Rhodopseudomonas palustris DX-1cType cytochromes2720Single chamberIndirect transfer[49]
Desulfovibrio desulfuricans ATTCcType cytochromes1580Single chamberIndirect transfer[84]
Geobacter metallireducenscType cytochromes, OmcE and OmcB450Single chamberIndirect transfer[85]
Desulfuromonas acetoxidanscType cytochromes2000_Indirect transfer[13]
Klebsiella pneumonia2,6–Ditertbutylpbenzoquinone199__[86]
Desulfovibrio alaskensisTransmembrane complexes, tetraheme cytochrome C3___[87]
Pseudomonas aeruginosaPhenazine–1–carboxamide, pyocyanin4300__[88]
Thermincola ferriaceticaAnthraquinone–2,6–disulfonate12,000Single chamber_[89]
Shewanella putrefaciensc–Type cytochromes including OmcA, MtrC, FAD transporter492Double chamberIndirect transfer[90]
Dechlorospirillum anomalous strain WDAnthraquinone2,6disulfonate hydrogen30_Direct transfer[91]
Geobacter lovleyiMethyl viologen480_Indirect transfer[92]
Chlorella vulgarisMethyl viologen, methylene blue30Single chamberIndirect transfer[91]
Pseudomonas sp.Methylene blue979Single chamberIndirect transfer[93]
Endoelectrogens microorganism
Rhodoferax ferrireducens_158Double chamberDirect transfer[94]
Klebsiela pneumoniae strain L17_34.77Double chamberDirect transfer[95]
Nocardiopsis sp. KNU (strain), Streptomyces enissocaesilis
KNU (K strains)
_162
145
Double chamberDirect transfer[96]
Rhodoferax ferrireducens__Double chamberDirect transfer[97]
Escherichia coli
strain K-12
_215Single chamber_[98]
Shewanella oneidensis__Single chamber_[99]
Pseudomonas aeruginosa_136 ± 87Single chamber_[100]
Cellulomonas fimi_0.74 ± 0.07Single chamberIndirect transfer[101]
Leptothrix discophora SP-6_70_Indirect transfer[102]
Acinetobacter calcoaceticus_110_Indirect transfer[50]
Escherichia coli_3390_ [103]
Winogradskyella poriferorum_40_Indirect transfer[104]
Pseudomonas fluorescens_210Double chamberDirect transfer[105]
Citrobacter sp._205Double chamberIndirect transfer[106]
Lysinibacillus sphaericus_850Double chamberDirect transfer[107]
Dechloromonas sp._300Double chamberIndirect transfer[108]
Arthrospira maxima_100Double chamberDirect transfer[109]
Coriolus versicolor_3200Single chamberIndirect transfer[110]
Table 2. Different substrates used in the BMFCs with corresponding power densities.
Table 2. Different substrates used in the BMFCs with corresponding power densities.
Waste SubstrateElectircigensPower Density
(mW/m2)
ConfigurationsType of Electrons Transfer MechanismsReferences
GlucoseAcidiphilium cryptum12.7Single chamberDirect transfer[120]
CelluloseEnterobacter cloacae5.4 ± 0.3Double chamberDirect transfer[122]
LactateShewanella oneidensis MR-10.3 × 10−2Single chamberIndirect transfer[123]
LactateGeobacter sulfurreducens52 ± 4.7-Indirect transfer[124]
GlucoseEscherichia coli228-Indirect transfer[125]
Malt extractEnterobacter cloacae9.3-Indirect transfer[126]
CelluloseG. sulfurreducens and C. cellulolyticum83Single chamberIndirect transfer[127]
Wheat strawAcidithiobacillus caldus123Single chamber-[128]
MolassesB. borstelensis STRI1185.5Single chamber-[119]
Sophorolipid with glucose and PBSPseudomonas aeruginosa15.29Single chamber-[129]
Glucose, fructose, and sucroseSaccharomyces cerevisiae72.77Single chamber-[130]
Glucose in synthetic wastewater_1313Double chamberDirect transfer[131]
xyloseGeobacter sulfurreducens Escherichia coli,590Double chamberDirect transfer[132]
Synthetic wastewaterαProteobacteria, βProteobacteria, γProteobacteria70Double chamber-[133]
Sodium FumarateGeobacter sulfurreducens_Single chamber-[134]
Glucuronic acidRhodococcus sp. and Paracoccus sp.2770Double chamber-[135]
XyloseClostridium spp. and Comamonas spp.1241_Direct transfer[136]
Acetate_1430_ [137]
EthanolProteobacterium sp., Azoarcus sp. and Desulfuromonas sp.40_Indirect transfer[138]
Synthetic wastewater with molasses and urea_2.9Single chamber [139]
CysteineShewanella affinis39__[140]
StarchClostridium butyricum or Clostridium beijerinckii___[141]
Dye-containing wastewater in microbial desalinationBacillus subtilis, Aeromonas hydrophila subsp. hydrophila2.86__[142]
Rice strawCellulose-degrading bacteria146__[121]
Coconut husk rettingOchrobactrum sp.362Double chamberIndirect transfer[143]
Agriculture wastewaterShewanella oneidensis13Double chamberIndirect transfer[144]
Rice paddyGeobacteraceae_Double chamberIndirect transfer[145]
ChitinBacillus circulans1.742Double chamberIndirect transfer[146]

Share and Cite

MDPI and ACS Style

Umar, M.F.; Abbas, S.Z.; Mohamad Ibrahim, M.N.; Ismail, N.; Rafatullah, M. Insights into Advancements and Electrons Transfer Mechanisms of Electrogens in Benthic Microbial Fuel Cells. Membranes 2020, 10, 205. https://doi.org/10.3390/membranes10090205

AMA Style

Umar MF, Abbas SZ, Mohamad Ibrahim MN, Ismail N, Rafatullah M. Insights into Advancements and Electrons Transfer Mechanisms of Electrogens in Benthic Microbial Fuel Cells. Membranes. 2020; 10(9):205. https://doi.org/10.3390/membranes10090205

Chicago/Turabian Style

Umar, Mohammad Faisal, Syed Zaghum Abbas, Mohamad Nasir Mohamad Ibrahim, Norli Ismail, and Mohd Rafatullah. 2020. "Insights into Advancements and Electrons Transfer Mechanisms of Electrogens in Benthic Microbial Fuel Cells" Membranes 10, no. 9: 205. https://doi.org/10.3390/membranes10090205

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