Open access peer-reviewed chapter

Microbial Community

Written By

Hajar Rajaei Litkohi and Hosein Yazdi Dehnavi

Submitted: 01 December 2023 Reviewed: 02 December 2023 Published: 16 April 2024

DOI: 10.5772/intechopen.1004001

From the Edited Volume

Revolutionizing Energy Conversion - Photoelectrochemical Technologies and Their Role in Sustainability

Mahmoud Zendehdel, Narges Yaghoobi Nia and Mohamed Samer

Chapter metrics overview

34 Chapter Downloads

View Full Metrics

Abstract

The microbial community employed as biocatalyst in microbial fuel cells (MFC) play a crucial role in degradation of organic substances and bioelectricity generation. While degradation of organic matters and electrical current generation in MFC technology is predominantly depend on metabolic activities of electroactive bacteria such as Geobacter and Proteobacteria, these bacteria engage in mutual interactions with non-electroactive counterparts within the microbial community. These mutual interactions can modify system performance, which is widely depended on operational conditions, the source of the initial microbial inoculum, substrate diversity and system’s components. Consequently, it is essential to gain a comprehensive understanding of the ecological behavior of microbial communities under diverse conditions to optimize system efficiency. Numerous research studies have delved into the microbial communities under varying circumstances, and the objective of this research is to elucidate the distinctions among microbial communities and investigate the factors that impact their composition.

Keywords

  • microbial fuel cells
  • microbial consortium
  • biocatalyst
  • bioelectricity generatione
  • electroactive bacteria

1. Introduction

The growing needs of our modern world, such as energy, water, and food, have caused significant harm to our environment. As our demand for electricity continues to rise sharply, there’s a pressing need to explore renewable energy sources [1]. One of the emerging technologies garnering substantial interest as a renewable energy source is microbial Fuel Cells (MFC) that use biological catalysts, primarily bacteria, to generate electric energy from organic matter, whether in the environment or waste [2]. MFC work with microorganisms as bio-catalysts that can extract energy from various waste materials and organic compounds [3]. The microorganisms oxidize organic and inorganic substrates at the anode, leading to the generation of electrons. Efficient transfer of these electrons from within the cells to the anode, in anoxic conditions, is vital for electric current production and involves mechanisms such as electron shuttles, mediators, and microbial nanowires [4, 5].

Diverse microorganisms, primarily bacteria from different phylogenetic groups, exhibit the capacity to generate electricity in MFCs without necessitating a mediator. Known as exoelectrogens or electrogenic microorganisms, several bacterial species like Geobacter spp., Shewanella spp., Rhodoferax ferrireducens, among others, have been identified for their electricity-producing abilities within MFCs. Additionally, microalgae, yeast, and fungi have also been reported in these systems [5].

Various factors significantly impact the operational efficiency of MFCs, including reactor configuration, anodic and cathodic materials, membrane type, substrate characteristics, internal resistance, loading rate, pH, buffer solutions, anodophilic populations, temperature, and salinity [5, 6, 7, 8, 9, 10]. Moreover, biological parameters such as inoculation source, pure culture versus consortium, Extracellular Electron Transfer (EET) rate, biofilm growth rate, and composition/activity of exoelectrogen populations play crucial roles since biocatalytic activity drives the MFC process [11, 12, 13, 14, 15, 16, 17].

Mixed microbial communities, particularly in bacterial MFCs, offer several advantages of adaptability and robustness to changing operating conditions [18, 19]. These communities demonstrate resilience against alterations in operating conditions, enabling detailed investigations of individual factors and fostering synergistic interactions among diverse microbes to enhance MFC stability and performance.

The utilization of mixed microbial communities within MFCs presents numerous advantages due to their ability to adapt and endure fluctuations in operating and environmental conditions. Conversely, the initiation phase of MFCs heavily relies on the establishment of biofilms and an enrichment process. The composition and structure of the resultant biofilm and microbial community formation significantly hinge upon various influencing factors. Nevertheless, comprehending the precise impact of these factors on the microbial community’s composition during optimal MFC performance poses a challenging endeavor. This study seeks to investigate and unravel the intricate influence of these factors on the behavior and dynamics of the microbial community.

Advertisement

2. Source of inoculum

Exoelectrogenic communities can be sourced from a variety of origins, including forest soil, anaerobic and aerobic sludge derived from wastewater treatment facilities, sediments, and compost [20, 21, 22, 23, 24]. In addition, isolated strains of model exoelectrogenes like Geobacter spp. and Shewanella spp. can also be utilized [25]. The impact of these diverse bacterial sources on the performance of such systems has been the subject of investigation in numerous studies. The presence of higher concentrations of exoelectrogenes and bacteria that promote biofilm formation can accelerate the development of biofilms on the anode [26]. However, the presence of electron acceptors, such as metal oxides and methanogenic microorganisms within the initial sludge can significantly influence the composition of bacterial communities on the anode and subsequently affect the performance of microbial fuel cells [27, 28]. Therefore, the careful selection of an inoculum source with a high concentration of active exoelectrogenes is of utmost importance.

The diversity of the initial microbial population and the concentrations of exoelectrogenic and non-exoelectrogenic bacteria are significantly influenced by the choice of inoculum source. Nonetheless, it is feasible to improve the inoculum to mitigate potential negative impacts. In a study conducted by Ishii et al., the inoculum was obtained from three distinct sources: (1) Anaerobic sludge from a wastewater treatment plant, (2) Paddy field soil, and (3) Coastal wetland sediments. The bioelectrochemical behavior of these three inocula was examined in a single-chamber fuel cell, and microbial population analysis was carried out using the 16S rRNA method. Paddy field soil displayed the most diverse initial microbial population in comparison to coastal wetland sediments and anaerobic sludge. In this study, an initial external resistance of 510 Ohms was applied, followed by a stepwise reduction to 100 and 25 Ohms to promote the enrichment of exoelectrogenic microorganisms. After 30 days of operation, the highest electric current, measuring 5.4 mA, was produced by the inoculum from coastal lagoon sediments. The analysis of the microbial population revealed that the stepwise enrichment process resulted in a reduction in microbial diversity within these three reactors. As a result, seven dominant bacterial groups emerged, comprising 90% of the microbial population in these cells. Consequently, by enriching the microbial population, it becomes possible to reduce the dependency of microbial fuel cell performance on the specific source of inoculum preparation [29].

While the source of inoculum can initially influence microbial community, over time and with consistent operational conditions, dominant microbial populations tend to converge. In a study conducted by Gao et al., two distinct inocula were examined. One was derived from the aerobic tank of a wastewater treatment plant, while the other originated from the anaerobic treatment reactor handling high sulfur wastewater. Both inocula underwent anaerobic treatment before being inoculated into the system. Bioelectrochemical analyses revealed that the inoculum from aerobic sludge reached maximum voltage more rapidly and exhibited higher voltage and current generation compared to the anaerobic sludge. Examination of the biofilm morphology on the anode biofilm indicated a higher presence of spherical bacteria in the aerobic sludge. Microbial population analysis in the initial stages showed that the aerobic sludge had greater diversity than the anaerobic sludge. However, as time passed and the system operated continuously, the microbial populations in both cells underwent progressive changes, ultimately leading to a convergence in dominant bacterial groups. In both inocula, dominant microbial populations became similar. The microbial population of the aerobic sludge inoculum was characterized by greater complexity and diversity, whereas the microbial population of the anaerobic sludge had a simpler composition. Notably, the anaerobic inoculum contained bromate-reducing bacteria, which are associated with the degradation of complex organic matter and sulfur reduction, reflecting the high sulfur content in the wastewater source. The similarity in dominant microbial populations in both inocula was attributed to their shared substrate. Specifically, Geobacter exoelectrogen species, biofilm-forming Zoogloea, and Acinetobacter species were prominent in both inocula [30].

Nutrient availability in the primary inoculum source significantly influences the diversity and microbial composition of the initial inoculum. In sedimentary Microbial Fuel Cells (SMFC), the sediments serve a dual role as both the substrate and the source of the inoculum, making sediment properties a crucial factor in SMFC performance. In a research study, seven lakes (denoted as S1 to S7) were assessed using SMFC technology. Of these, S1, S2, and S3 were large lakes, while S4 to S7 were smaller local lakes. The research aimed to investigate the geochemical characteristics and the initial microbial communities of each lake and their correlation with the biochemical performance, microbial composition, and microbial population in the SMFC systems. The findings revealed that the highest power output, 25.2 and 23.9 mW/m2, was observed in S4 and S5, respectively, and these two systems exhibited the quickest startup times, at 7 and 9 days, respectively. The analysis of sediment characteristics indicated that the high relative yield in S4 and S5 could be attributed to the abundance of carbon and nitrogen sources, signifying higher nutrient levels in these sediments. Furthermore, the removal and degradation of organic matter in S4 and S5 were approximately 1.5 times higher than in the other systems. The initial microbial population in S4 and S5 displayed greater diversity compared to the other systems, and the types of bacteria comprising the populations in S4 and S5 differed from those in the other lakes. All seven SMFC systems exhibited similar voltage levels, ranging from 500 to 600 mV. After 60 days, an examination of the microbial populations on the anode indicated that Proteobacteria were predominant in all samples. However, in S4, only 43% of the population consisted of Proteobacteria, with Pseudomonas bacteria playing a significant role in the microbial population of S4 and S5 [31].

In the selection of the inoculum source, it is essential to consider the specific objectives of employing a fuel cell. For instance, when the aim is the remediation or elimination of a particular pollutant, it becomes possible to isolate microorganisms from the contaminated environment containing the target substance and employ them within a fuel cell system. In a study conducted by Sharma et al., samples were obtained from soil contaminated with hydrocarbons for the purpose of addressing phenanthrene pollution through the use of fuel cells. Following the preparation of the soil sample, it was dissolved in heptamethylnonane, and after an incubation period of 300 days, the mixture underwent filtration and subsequent centrifugation. The centrifuged material was utilized as the primary source of inoculum and introduced into a two-chamber fuel cell. Within this cell, phenanthrene was immobilized on the anode electrode, serving as a substrate for the formation of a biofilm by bacteria capable of decomposing organic matter. The investigation revealed that an increase in the loading concentration of this carbon material on the anode led to a reduction in current production. The maximum power output achieved was 37 mW/m2 at a concentration of 2 mg/cm2. Upon the analysis of the microbial population, it was observed that electroactive bacteria, notably Pseudomonas, were prevalent within this system, and these bacteria exhibited the capability to degrade phenanthrene, contributing to the pollutant’s removal process [32].

Advertisement

3. Anolyte as substrate

The electrical energy generated by MFC significantly relies on the composition of the organic matter within the anolyte. Microorganisms exhibit the ability to metabolize an array of substrates, spanning from inorganic compounds to intricate organic compounds, utilizing them as sources of energy or carbon. MFC studies have explored various substrates, including simple organic compounds like glucose, ethanol, and acetate, which are easily metabolized, as well as complex organic compounds like starch, chitin, and cellulose. However, while these organic materials are effective, they might not be economically efficient. Alternatively, wastewaters present a viable and cost-effective option. Nevertheless, wastewaters encompass a diverse range of complex organic materials, influencing the microbial population based on the specific organic constituents present within them [33, 34, 35, 36, 37].

In research conducted by sotres et al., a comparison was made between synthetic wastewater and pig slurry as an anolyte. The synthetic wastewater exhibited a power output of 2138 mW/m3, while pig slurry showed 5623 mW/m3. Analysis of microbial diversity revealed that the type of substrate significantly influenced the microbial population. In the synthetic wastewater where acetate was utilized, bacterial groups like Alcaligenaceae, Pseudomonadaceae and Clostridiaceae were among the dominant families. Conversely, in the pig slurry, the Flavobacteriaceae, Chitinophagaceae, Comamonadaceae and Nitrosomonadaceae families stood out as dominant among the microbial groups [38]. Indeed, the composition of the anolyte exerts a significant influence on the microbial community, regardless of the source of inoculum. The specific constituents and composition of the anolyte play a crucial role in shaping and determining the microbial population within the system. Table 1 provides an overview of various anolytes and their influence on the microbial community.

InoculumAnolytePower outputMicrobial communityRef.
Wastewater sludgeThe organic fraction of municipal solid waste (OFMSW)116 mW/m2[16]
moist manure123 mW/m2
sediment samplebrewery wastewater (inlet)8.001 μW/cm2Geobacter
Shewanella
Clostridium
[39]
brewery wastewater (outlet)1.843 μW/cm2
sludgeglucose,0.11 W/m2Clostridium
Azospirillum
[40]
sodium
acetate
0.098 W/m2Moheibacter
Azospirillum
food waste hydrolysate0.173 W/m2Rummeliibacillus
Burkholderia, Enterococcus Clostridium
anaerobic consortiaorange peel waste358.8 mW/firmicutes[41]
sludgeFood waste422 mW/m2Firmicutes
Bacteroidetes
Proteobacteria
[42]
sludgesea food processing wastewater05 mW/m2Stenotrophomonas[43]
Swine wastewaterswine
wastewater
1.92 kWh/m3Arcobacter
Pseudomonas
Acinetobacter
[44]

Table 1.

Anolytes and their effect on microbial community.

Advertisement

4. Additives

Modifying the anolyte by introducing a range of additives can alter its environmental conditions, consequently enhancing cell performance. This manipulation not only impacts bacterial metabolism but also contributes to improving overall cell efficiency. However, these alterations also exert an influence on the microbial population. Additives such as Cu2+, CD2+, S2−, and surfactants (TWEEN80) have been employed to manipulate the environmental conditions [45, 46, 47].

In a study conducted by Chen et al., they incorporated sodium citrate as an additive, serving dual roles in pH adjustment and electron transfer. This combined addition led to a notable enhancement of 58% compared to the control group, with the output power peaking at a maximum of 785 mW/m2. Investigation of the microbial population revealed an increase in microbial diversity following the introduction of citrate, attributed to two primary reasons. Firstly, maintaining a neutral pH balance promotes the growth of a wider array of microorganisms. Secondly, the citrate additive itself serves as a carbon source for microorganisms, stimulating their growth. In the control condition without citrate, Proteobacteria and Bacteroidetes were the dominant species. However, with the inclusion of citrate, the relative abundance of Proteobacteria decreased, while Bacteroidetes and Firmicutes exhibited higher prevalence. Among the Proteobacteria, gammaproteobacteria were predominant in the control condition, yet the addition of citrate diminished their dominance and elevated the relative abundance of alphaproteobacterial [48].

Exoelectrogens transfer the produced electrons to the anode using two methods: direct transfer and transfer facilitated by electron mediators. Additionally, the introduction of artificial electron mediators can enhance electron transfer within exoelectrogenes [49]. These mediators affect the bioelectrochemical properties of MFC and thus affect the microbial population as well. In Lin’s study, they compared two electron mediators, neutral red (NR) and potassium ferricyanide (PFC), in research involving a two-chamber fuel cell fueled by toluene. The introduction of these electron mediators resulted in an increase in voltage, elevating it from 5.53 mV to 109 mV and 2.88 mV for NR and PFC, respectively. Analysis of microbial diversity indicated that these electron mediators play a role in shaping the microbial community’s development. Seven bacterial groups were consistently found in all microbial fuel cells (MFCs). However, the reduced variability in the presence of these bacterial groups in MFCs utilizing NR and PFC respectively could be attributed to potential toxicity linked with these mediators [50].

Advertisement

5. Components

In general, the prevailing belief is that the current generation in an MFC is contingent on both the quantity of biofilm formation on the anode electrode and the level of bioelectrical activity within this biofilm [51]. Furthermore, the anode electrode stands as a pivotal component in determining the overall cost of fuel cell technology, typically constituting about 10% of the total expenses for constructing an MFC [52]. Consequently, the selection of the anode material plays a critical role in influencing both the performance and cost considerations of the battery. Carbon-based materials, such as carbon paper, carbon cloth, graphite, carbon brush, and similar options, are commonly employed in research due to their advantageous electrochemical properties and cost-effectiveness [53]. Variations among these materials pertain to their surface area, porosity, and conductivity, factors that directly impact the absorption rate and growth of electroactive bacteria.

In this investigation, four distinct carbon materials—granular activated carbon (GAC), granular semicoke (GS), carbon felt cube (CFC), and granular graphite (GG)—each possessing respective surface areas of 686 m2/g, 236 m2/g, 0.845 m2/g, and 0.623 m2/g, were assessed within a dual-chambered MFC. The power output was determined as 31.5, 27.9, 26.1, and 23.8 W/m2 for GAC, CFC, GG, and GS, respectively. The researchers scrutinized the microbial community present on each electrode using the 16S rRNA method. Notably, GAC demonstrated the highest power output and displayed a dominant presence of 95% Geobacter in its microbial composition. Similarly, the CFC anode exhibited a dominance of Geobacter (43%) and Ruminobacillus (16%). On the GG electrode, the most prevalent phylum was Azospira, constituting 65% of the community. Conversely, the GS anode displayed an abundance of Bacteroidetes (19%), Sinorhizobium (15%), and Synergistes (13%). The outcomes underscored that the greater porosity and improved bioelectrochemical behavior of GAC resulted in a higher power output and an enrichment of electroactive bacteria [54].

The physical and chemical properties of the anode material significantly impact the composition of the electroactive biofilm formed by microorganisms. Altering these properties directly influences the microbial community, its diversity, and enhances the enrichment of microbial populations. Numerous studies have demonstrated that stabilizing metals or metal oxides on carbon materials leads to a notable increase in conductivity, often by two to threefold. This heightened conductivity plays a pivotal role in improving the enrichment of microbial populations [55, 56].

In a study by Xu et al., nanoparticles of MNO2, Fe3O4, and Pd were loaded on carbon cloth treated with 5% Nafion. These modified carbon cloth materials served as anodes in a two-chamber cell. They were then compared against a control anode consisting of bare carbon cloth. The initial inoculum for the MFC was obtained from activated sludge sourced from a wastewater treatment plant. Synthetic wastewater was used as the substrate for the MFC. Electrochemical indicated that Pd, MNO2, and Fe3O4 exhibited power output of 824, 782, and 728 mW/m2, respectively. These values surpassed the power output of the control, which utilized bare carbon cloth with a power output of 680 mW/m2. Anodes treated with MNO2 and Fe3O4 particularly facilitated the enrichment of Geobacter, constituting for 73% of the microbial coverage. On the Pd-treated anode, Geobacter and sphaerochaeta bacteria were the predominant species, comprising 38% and 16% of the microbial composition, respectively [57].

In a different study, bio-synthesized gold nanoparticles (bio-Au) were utilized independently as well as in combination with multi-walled carbon nanotubes (MWCNTs) to modify carbon cloth electrodes. Overall, the incorporation of gold nanoparticles significantly enhanced the bioelectric behavior of the MFC. Specifically, BioAu/MWCNT showed the best performance compared to unmodified carbon cloth. The BioAu/MWCNT modification resulted in a 140% improvement in system startup time and a 56% increase in system output power, with the maximum power output reaching 178 mw. In this anode modified with gold nanoparticles and MWCNT, the bacterial groups Gammaproteobacteria and Negativicutes were among the dominant species [58].

In conclusion, the modification of the anode improves its physiochemical characteristics, creating a significant preference for EAB over fermentative bacteria. Despite not notably impacting bacterial diversity, the increased presence of electroactive bacteria on the modified electrodes resulted in better MFC system performance. Table 2 provides a summary of the modifications made to the anodes.

Anode modificationInoculumPower outputMicrobial communityRef.
Graphene oxide (GO)active sludge124.58 mW/m2increase in the relative ratio of
Proteobacteria, but a decrease of Firmicutes
[59]
tungsten carbideanaerobic sludge3.26 W/m2Geobacter
Geothrix
Pseudomonas
[60]
CF/N-CNT/PANI3.8 W/m 3Desulfuromonas
Geobacter
[61]
poly(3,4-ethylenedioxythiophene)sludgeGeobacter[62]
Graphene oxide (GO) and carbon nanotubes (CNTs)Pseudomonas Thauera
Diaphorobacter
Tumebacillus Lysobacter
[63]

Table 2.

Anode modifications and dominant microorganism.

Certain environmental conditions, such as elevated COD levels, high salinity, and the toxicity of wastewater, can hinder the performance of microbial biofilms. To enhance performance under such challenging circumstances, bacteria can be securely immobilized on the anode surface to mitigate the adverse effects of these harsh conditions and prevent a decline in MFC performance. Various methods for immobilizing bacteria on the anode have been reported, including adsorption, covalent binding, and entrapping. Through these methods, a substantial concentration of microorganisms can be firmly anchored to the anode’s surface. Among these different stabilization techniques, the entrapping method stands out due to its stability and enhanced strength [64].

In a study conducted by Leo et al., they employed the entrapping method to immobilize inoculum obtained from a municipal wastewater treatment plant onto carbon cloth using agarose gel. They examined the impact of varying concentrations of sodium acetate (ranging from 1 to 20 g/l) and compared it to non-immobilized carbon cloth as control. The stabilization of bacteria led to a 60% and 70% increase in output power for concentrations of 5 and 10 g/liter, respectively. Through an analysis of the microbial community, they demonstrated that stabilizing the inoculum on carbon cloth resulted in a 60% prevalence of Geobacteria, whereas the control anode exhibited a Geobacteria prevalence of 40% [65].

The membrane constitutes a pivotal factor that influences both the bioelectrochemical behavior and the microbial community within the developed biofilm on the anode electrode of a MFC [66, 67]. Serving as a barrier between the electrodes and compartments of the MFC, it plays a crucial role in upholding the anaerobic condition on the anode and preventing the undesired leakage of organic substances from the anolyte to the catholyte [68, 69]. Any leakage of anolyte substrate to the catholyte adversely impacts the kinetics of oxygen reduction. Furthermore, oxygen leakage into the anolyte promotes the proliferation of aerobic or facultative anaerobic bacteria, leading to the wastage of substrate and influencing the kinetics of electron production [70]. Additionally, the adhesion of microorganisms to the membrane’s surface results in biofouling affecting overall performance of the system [71]. Consequently, the type of membrane emerges as a key determinant affecting the overall performance of the system, consequently influencing the microbial population dynamics.

They modified a polymeric membrane (SPEEK) with nano silver, ranging from weight percentages of 2.5 to 10 mg/cm-2, and assessed its performance in a tubular Microbial Fuel Cell (MFC). The treatment with nano silver showed no substantial difference in water absorption levels, yet it notably enhanced proton ion transfer rates and reduced internal resistance and oxygen transfer rates. Examining various weight percentages for the treated membranes, it was found that the membrane treated with 7.5% by weight of silver exhibited the highest output power compared to other treatments. Subsequently, a comparison was made between the output power and microbial population of SPEEK7.5 with Nafion 117 and the non-treated SPEEK membrane, where the maximum output power recorded was 156 mw/m2. The presence of silver nanoparticles exhibited an antibacterial property, resulting in reduced biofouling on the membrane, with only silver-resistant bacteria thriving in the environment. The analysis identified 33 bacterial groups, with over 6000 bacterial species falling under Proteobacteria and Firmicutes [72].

Advertisement

6. Operational conditions

The oxygen present in the anolyte can significantly impact the functioning of MFC. As the performance of MFCs relies on exoelectrogenic bacteria, the presence of oxygen and aeration can alter their performance by influencing these microorganisms. Exoelectrogens respond differently to the presence of oxygen. For instance, Geobacter, being an obligate anaerobe, is highly susceptible to oxygen exposure. Even minimal contact with oxygen can inhibit its respiration and decrease power generation [73]. Conversely, facultative anaerobes like Shewanella can generate electricity under both aerobic and anaerobic conditions, although the quantity produced varies [74]. While several studies have demonstrated that even minimal oxygen ingress to the anode can diminish MFC performance, higher concentrations of oxygen can actually enhance performance [75].

In an investigation by Quan et al., a MFC was employed, and a combination of aerobic and anaerobic sludge served as the inoculum. The researchers noted that the output voltage was greater under anaerobic conditions compared to aerobic conditions. Additionally, they observed prolonged cycles in the anaerobic setup. These findings suggest that the presence of aerobic conditions encouraged the proliferation of bacteria detrimental to the system’s performance. Analysis of the microbial population revealed that aeration led to a decline in diversity, with species such as Bacteroidetes and others becoming dominant [76].

External resistance is one of the factors influencing the performance of MFC. In some studies, it has been shown that the use of external resistance has increased the removal of COD [77]. Given that the anode functions as the final electron acceptor, adjusting the external resistance can impact system performance by directing electron flow towards the cathode and facilitating the regeneration of the electron acceptor. This electron flow alteration can influence the diversity and composition of the microbial community [78]. Understanding the effect of external resistance on the microbial community provides a holistic understanding of exoelectrogen performance. External resistance prompts bacterial growth, with higher resistance leading to increased biomass. However, as resistance intensifies, there’s a reduction in current generation [79]. Consequently, it’s crucial to explore the appropriate amount of external resistance to optimize system performance.

Katuri et al., conducted a study exploring the influence of external resistance on the composition of the anode biofilm and microbial population. They employed a two-chamber MFC, utilizing glucose as a substrate, and varied the resistances from 100 to 50 kΩ. Their observations revealed that as resistance increased, the current production in the system decreased, and the system reached peak performance at a slower rate. Interestingly, they found that the highest production of biomass was achieved with 25 and 50 kΩ resistors, surpassing even the conditions of open circuit voltage. The investigation of microbial diversity showed that with the increase in resistance from 100 ohms to 50 kΩ, bacterial groups increased from 15 to 27. 100 Ohm resistance showed the simplest microbial diversity [80].

Continuous operation stands as a vital requirement for effective and economical wastewater treatment processes. In this context, the Organic Loading Rate (OLR) holds importance, and its control is influenced by the Hydraulic Retention Time (HRT). HRT, a critical operational parameter, directly affects OLR, subsequently impacting the production of electricity in MFCs [81, 82].

The study conducted by Haavisto et al. utilized a continuous two-chamber system supplied with xylose. Their observations revealed a significant impact of HRT variations on electricity generation. Decreasing the HRT from 3.5 days to 1 day resulted in an increase in system voltage from 344 mV to 480 mV. However, further reduction in HRT to 0.75 days and 0.17 days led to decreased voltages of 218 mV and 156 mV, respectively. Moreover, when adjusting the external resistance, the highest current production of 2460 mA/m2 was achieved with an HRT of 1 day. Additionally, HRTs exceeding 1 day displayed Chemical Oxygen Demand (COD) levels below 0.4 g. L−1/day, indicating insufficient substrate availability for the metabolism of microorganisms in the system. The decrease in Hydraulic Retention Time (HRT) led to the elimination of bacteria that were not attached to the biofilm. Consequently, there were notable shifts in microbial population diversity throughout the experiment. When the HRT was reduced to 0.5 days, there was an increase in the xylose content within the anolyte. This increase correlated with a significant rise in Christensenella, a type of bacteria adept at fermenting xylose. During the fermentation process of xylose, acetate, propionate, and butyrate are generated as byproducts, serving as carbon sources for exoelectrogenic bacteria. However, the heightened concentration of the substrate led to excessive proliferation of xylose-fermenting bacteria. This overgrowth, in turn, resulted in decreased current production within the system [83].

Studies have demonstrated that fluctuations in temperature significantly affect the initiation and continuous operation of MFCs. For instance, MFCs operating below 15°C experienced prolonged startup times and limited electricity production within a month of operation. However, recent research has highlighted successful startup and sustained functioning of both MFCs and Microbial Electrolysis Cells (MECs) even at extremely low temperatures, such as 4°C. These studies uncovered the existence of distinct psychrophilic populations within the anode biofilms, supporting extracellular electron transfer in MECs at lower temperatures [84, 85].

In the study conducted by Mei et al., MFC reactors were initially operated at 25°C to establish a biofilm on the anode surface. Following biofilm formation, the performance of the system was evaluated across temperatures of 10, 20, and 30°C over six cycles. Their findings demonstrated a correlation between decreasing temperature and increased cycle duration, as well as a decrease in power output. The highest output power, reaching 894 mW/m2, was achieved at 30°C. Moreover, the maximum COD removal of 93% was observed at 30°C, with reduced COD removal at lower temperatures.

Throughout all tested temperatures, Proteobacteria, Bacteroidetes, and Firmicutes were the dominant bacterial groups. Specifically, Proteobacteria accounted for 61%, 65%, and 45% at temperatures of 10, 20, and 30°C, respectively. Bacteroidetes prevalence was 17%, 20%, and 25.5% at 10, 20, and 30°C, while Firmicutes constituted 6% at 10 and 20°C, rising to 11% at 30°C [86]. The study indicated that biofilm formation, metabolic activity, and extracellular electron transfer in MFCs are influenced by temperature.

Advertisement

7. Biodegradation

The potential of MFC technology extends beyond bioenergy production, delving into bioremediation, marking a promising avenue for efficient and sustainable environmental cleanup. Biodegradation serves as a sustainable resolution to combat persistent pollutant contamination. While physical-chemical methods exhibit rapid treatment efficacy, they often entail high costs and the potential generation of harmful by-products, diverging from the principles of Green Remediation. Bioremediation, on the other hand, harnesses the inherent capability of natural microbial communities to break down contaminants. This approach aligns with environmental sustainability, offering economic and social advantages while restoring ecosystems [86].

The composition of the microbial community is impacted by the substances that need to be biologically degraded. The recognition of resistance genes and clarification of degradation mechanisms, such as efflux pumps and modifying factors, emphasize the importance of biodegradation within MFCs. It underscores the complex interplay between the primary microorganisms present on the anode and the particular substrates utilized in MFCs. Grasping these associations is crucial for enhancing the efficiency of MFCs in breaking down pollutants, highlighting the essential role of biodegradation in the sustainable remediation of polluted water [87, 88, 89]. Table 3 summarizes various research studies conducted in the field of biodegradation, focusing on the dominant microbial populations responsible for reducing pollution related to specific pollutants.

PollutantEfficiency (%)Power OutputDominant BacteriaRef.
Food wastes95.956 mW/m2Geobacter
Bacteroides
[90]
vanadium76.819 ± 11 mW/mDeltaproteobacteria
Bacteroidetes Spirochaete
[91]
Copper98.3%10.2 W m-3Proteobacteria
Bacteroidetes
Actinobacteria
Acidobacteria
[92]
sulfamethoxazole87.521186.2 mW/m2Proteobacteria
Bacteroidetes
[93]
nitrogen1006.3 W/m3Nitrosomonas
Nitratireductor
Acidovorax
[94]
p-nitrophenol81Corynebacterium
Comamonas
Chryseobacterium
Rhodococcus
[95]
bisphenol and ibuprofen78.7–63.2, respectively28.63 mW m2Chlorobi
Proteobacteria
Firmicutes
Bacteroidetes
[96]
Congo red decolorization97%41.1 mW/m2Proteobacteria
Bacteroidetes
Actinomycetes and Firmicutes
[97]
polychlorinated biphenyls(37.5108.89 mW/m2Geobacter and Pseudomonas Longilinea Desulfofustis)[98]

Table 3.

Pollutant removal and dominant bacterial community.

Advertisement

8. Conclusion

Microbial fuel cells have a dual function of electrical power generation and treating a variety of wastewater that contain both organic and inorganic pollutants. As a form of biological technology, these cells utilize microbial catalysts to drive bioelectrochemical reactions. Mixed microbial populations are predominantly employed due to their exceptional adaptability to diverse environmental conditions. While certain bacterial groups may enriched and dominated based on system function and the type of wastewater utilized, this is contingent upon various factors. The primary source of the initial microbial community significantly influences microbial community development and can be selected based on the intended application of the MFCs. Additionally, organic and inorganic substances within the anolyte create conditions favoring specific species that metabolize these substances, influencing bacterial metabolism and their dominance. Alterations in MFC components, such as electrodes and membrane modifications can influence electrochemical properties, which impact biological properties and the enrichment of the microbial community. Environmental parameters like temperature, light, aeration and external resistance also exert considerable influence over the microbial community. Thus, it is essential to perceive and explore MFCs as a form of biological technology in order to enhance their performance and target specific applications.

Advertisement

Conflict of interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. 1. Himri Y, Malik AS, Boudghene Stambouli A, Himri S, Draoui B. Review and use of the Algerian renewable energy for sustainable development. Renewable and Sustainable Energy Reviews. 2009;13:1584-1591. DOI: 10.1016/j.rser.2008.09.007
  2. 2. Mendes FB, Bordignon SE. Renewable Energy and the Role of Biofuels in the Current World. In: Recent Developments in Bioenergy Research. Amsterdam, The Netherlands: Elsevier; 2020. pp. 65-84. DOI: 10.1016/B978-0-12-819597-0.00003-9
  3. 3. Cai T, Meng L, Chen G, Xi Y, Jiang N, Song J, Zheng S, et al. Application of advanced anodes in microbial fuel cells for power generation: A review. Chemosphere. 2020;248:5. DOI: 10.1016/j.chemosphere.2020.125985
  4. 4. Kouzuma A, Kasai T, Hirose A, Watanabe K. Catabolic and regulatory systems in Shewanella oneidensis MR-1 involved in electricity generation in microbial fuel cells. Frontiers in Microbiology. 2015;6:1-11. DOI: 10.3389/fmicb.2015.00609
  5. 5. Kumar R, Singh L, Zularisam AW. Exoelectrogens: Recent advances in molecular drivers involved in extracellular electron transfer and strategies used to improve it for microbial fuel cell applications. Renewable and Sustainable Energy Reviews. 2016;56:1322-1336. DOI: 10.1016/j.rser.2015.12.029
  6. 6. Lefebvre O, Tan Z, Kharkwal S, Ng HY. Effect of increasing anodic NaCl concentration on microbial fuel cell performance. Bioresource Technology. 2012;112:336-340. DOI: 10.1016/j.biortech.2012.02.048
  7. 7. Borole AP, Reguera G, Ringeisen B, Wang Z-W, Feng Y, Kim BH. Electroactive biofilms: Current status and future research needs. Energy & Environmental Science. 2011;4:4813. DOI: 10.1039/c1ee02511b
  8. 8. Pant D, Van Bogaert G, Diels L, Vanbroekhoven K. A review of the substrates used in microbial fuel cells (MFCs) for sustainable energy production. Bioresource Technology. 2010;101:1533-1543. DOI: 10.1016/j.biortech.2009.10.017
  9. 9. Logan BE, Rabaey K. Conversion of wastes into bioelectricity and chemicals by using microbial electrochemical technologies. Science. 1979;337(2012):686-690. DOI: 10.1126/science.1217412
  10. 10. Jung S, Regan JM. Influence of external resistance on Electrogenesis, Methanogenesis, and anode prokaryotic communities in microbial fuel cells. Applied and Environmental Microbiology. 2011;77:564-571. DOI: 10.1128/AEM.01392-10
  11. 11. Kiely PD, Regan JM, Logan BE. The electric picnic: Synergistic requirements for exoelectrogenic microbial communities. Current Opinion in Biotechnology. 2011;22:378-385. DOI: 10.1016/j.copbio.2011.03.003
  12. 12. White HK, Reimers CE, Cordes EE, Dilly GF, Girguis PR. Quantitative population dynamics of microbial communities in plankton-fed microbial fuel cells. The ISME Journal. 2009;3:635-646. DOI: 10.1038/ismej.2009.12
  13. 13. Freguia S, Teh EH, Boon N, Leung KM, Keller J, Rabaey K. Microbial fuel cells operating on mixed fatty acids. Bioresource Technology. 2010;101:1233-1238. DOI: 10.1016/j.biortech.2009.09.054
  14. 14. Ishii S, Suzuki S, Norden-Krichmar TM, Phan T, Wanger G, Nealson KH, et al. Microbial population and functional dynamics associated with surface potential and carbon metabolism. The ISME Journal. 2014;8:963-978. DOI: 10.1038/ismej.2013.217
  15. 15. Miceli JF, Parameswaran P, Kang D-W, Krajmalnik-Brown R, Torres CI. Enrichment and analysis of anode-respiring bacteria from diverse anaerobic Inocula. Environmental Science & Technology. 2012;46:10349-10355. DOI: 10.1021/es301902h
  16. 16. El-Chakhtoura J, El-Fadel M, Rao HA, Li D, Ghanimeh S, Saikaly PE. Electricity generation and microbial community structure of air-cathode microbial fuel cells powered with the organic fraction of municipal solid waste and inoculated with different seeds. Biomass and Bioenergy. 2014;67:24-31. DOI: 10.1016/j.biombioe.2014.04.020
  17. 17. Yates MD, Kiely PD, Call DF, Rismani-Yazdi H, Bibby K, Peccia J, et al. Convergent development of anodic bacterial communities in microbial fuel cells. The ISME Journal. 2012;6:2002-2013. DOI: 10.1038/ismej.2012.42
  18. 18. Yan X, Lee H-S, Li N, Wang X. The micro-niche of exoelectrogens influences bioelectricity generation in bioelectrochemical systems. Renewable and Sustainable Energy Reviews. 2020;134:110184. DOI: 10.1016/j.rser.2020.110184
  19. 19. Nevin KP, Richter H, Covalla SF, Johnson JP, Woodard TL, Orloff AL, et al. Power output and columbic efficiencies from biofilms of Geobacter sulfurreducens comparable to mixed community microbial fuel cells. Environmental Microbiology. 2008;10:2505-2514. DOI: 10.1111/j.1462-2920.2008.01675.x
  20. 20. Ishii S, Shimoyama T, Hotta Y, Watanabe K. Characterization of a filamentous biofilm community established in a cellulose-fed microbial fuel cell. BMC Microbiology. 2008;8:6. DOI: 10.1186/1471-2180-8-6
  21. 21. Ishii S, Suzuki S, Norden-Krichmar TM, Nealson KH, Sekiguchi Y, Gorby YA, et al. Functionally stable and phylogenetically diverse microbial enrichments from microbial fuel cells during wastewater treatment. PLoS One. 2012;7:e30495. DOI: 10.1371/journal.pone.0030495
  22. 22. 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. DOI: 10.1371/journal.pone.0077443
  23. 23. Dunaj SJ, Vallino JJ, Hines ME, Gay M, Kobyljanec C, Rooney-Varga JN. Relationships between soil organic matter, nutrients, bacterial community structure, and the performance of microbial fuel cells. Environmental Science & Technology. 2012;46:1914-1922. DOI: 10.1021/es2032532
  24. 24. Cercado B, Byrne N, Bertrand M, Pocaznoi D, Rimboud M, Achouak W, et al. Garden compost inoculum leads to microbial bioanodes with potential-independent characteristics. Bioresource Technology. 2013;134:276-284. DOI: 10.1016/j.biortech.2013.01.123
  25. 25. Madjarov J, Prokhorova A, Messinger T, Gescher J, Kerzenmacher S. The performance of microbial anodes in municipal wastewater: Pre-grown multispecies biofilm vs. natural inocula. Bioresource Technology. 2016;221:165-171. DOI: 10.1016/j.biortech.2016.09.004
  26. 26. Logan BE. Exoelectrogenic bacteria that power microbial fuel cells. Nature Reviews. Microbiology. 2009;7:375-381. DOI: 10.1038/nrmicro2113
  27. 27. An J, Kim B, Nam J, Ng HY, Chang IS. Comparison in performance of sediment microbial fuel cells according to depth of embedded anode. Bioresource Technology. 2013;127:138-142. DOI: 10.1016/j.biortech.2012.09.095
  28. 28. Song T-S, Cai H-Y, Yan Z-S, Zhao Z-W, Jiang H-L. Various voltage productions by microbial fuel cells with sedimentary inocula taken from different sites in one freshwater lake. Bioresource Technology. 2012;108:68-75. DOI: 10.1016/j.biortech.2011.11.136
  29. 29. Ishii S, Suzuki S, Yamanaka Y, Wu A, Nealson KH, Bretschger O. Population dynamics of electrogenic microbial communities in microbial fuel cells started with three different inoculum sources. Bioelectrochemistry. 2017;117:74-82. DOI: 10.1016/j.bioelechem.2017.06.003
  30. 30. Gao C, Wang A, Wu W-M, Yin Y, Zhao Y-G. Enrichment of anodic biofilm inoculated with anaerobic or aerobic sludge in single chambered air-cathode microbial fuel cells. Bioresource Technology. 2014;167:124-132. DOI: 10.1016/j.biortech.2014.05.120
  31. 31. Song N, Jiang H-L. Effects of initial sediment properties on start-up times for sediment microbial fuel cells. International Journal of Hydrogen Energy. 2018;43:10082-10093. DOI: 10.1016/j.ijhydene.2018.04.082
  32. 32. Sharma M, Nandy A, Taylor N, Venkatesan SV, Ozhukil Kollath V, Karan K, et al. Bioelectrochemical remediation of phenanthrene in a microbial fuel cell using an anaerobic consortium enriched from a hydrocarbon-contaminated site. Journal of Hazardous Materials. 2020;389:121845. DOI: 10.1016/j.jhazmat.2019.121845
  33. 33. Ren Z, Ward TE, Regan JM. Electricity production from cellulose in a microbial fuel cell using a defined binary culture. Environmental Science & Technology. 2007;41:4781-4786. DOI: 10.1021/es070577h
  34. 34. Rezaei F, Richard TL, Logan BE. Analysis of chitin particle size on maximum power generation, power longevity, and Coulombic efficiency in solid–substrate microbial fuel cells. Journal of Power Sources. 2009;192:304-309. DOI: 10.1016/j.jpowsour.2009.03.023
  35. 35. Logan B, Cheng S, Watson V, Estadt G. Graphite Fiber brush anodes for increased power production in air-cathode microbial fuel cells. Environmental Science & Technology. 2007;41:3341-3346. DOI: 10.1021/es062644y
  36. 36. Cheng S, Liu H, Logan BE. Increased performance of single-chamber microbial fuel cells using an improved cathode structure. Electrochemistry Communications. 2006;8:489-494. DOI: 10.1016/j.elecom.2006.01.010
  37. 37. Liu H, Cheng S, Logan BE. Production of electricity from acetate or butyrate using a single-chamber microbial fuel cell. Environmental Science & Technology. 2005;39:658-662. DOI: 10.1021/es048927c
  38. 38. Sotres A, Tey L, Bonmatí A, Viñas M. Microbial community dynamics in continuous microbial fuel cells fed with synthetic wastewater and pig slurry. Bioelectrochemistry. 2016;111:70-82. DOI: 10.1016/j.bioelechem.2016.04.007
  39. 39. Çetinkaya AY, Köroğlu EO, Demir NM, Baysoy DY, Özkaya B, Çakmakçı M. Electricity production by a microbial fuel cell fueled by brewery wastewater and the factors in its membrane deterioration. Chinese Journal of Catalysis. 2015;36:1068-1076. DOI: 10.1016/S1872-2067(15)60833-6
  40. 40. Xin X, Hong J, Liu Y. Insights into microbial community profiles associated with electric energy production in microbial fuel cells fed with food waste hydrolysate. Science of The Total Environment. 2019;670:50-58. DOI: 10.1016/j.scitotenv.2019.03.213
  41. 41. Miran W, Nawaz M, Jang J, Lee DS. Conversion of orange peel waste biomass to bioelectricity using a mediator-less microbial fuel cell. Science of The Total Environment. 2016;547:197-205. DOI: 10.1016/j.scitotenv.2016.01.004
  42. 42. Asefi B, Li S-L, Moreno HA, Sanchez-Torres V, Hu A, Li J, et al. Characterization of electricity production and microbial community of food waste-fed microbial fuel cells. Process Safety and Environmental Protection. 2019;125:83-91. DOI: 10.1016/j.psep.2019.03.016
  43. 43. Jayashree C, Tamilarasan K, Rajkumar M, Arulazhagan P, Yogalakshmi KN, Srikanth M, et al. Treatment of seafood processing wastewater using upflow microbial fuel cell for power generation and identification of bacterial community in anodic biofilm. Journal of Environmental Management. 2016;180:351-358. DOI: 10.1016/j.jenvman.2016.05.050
  44. 44. Li M, Zhou M, Tian X, Tan C, Gu T. Enhanced bioenergy recovery and nutrient removal from swine wastewater using an airlift-type photosynthetic microbial fuel cell. Energy. 2021;226:120422. DOI: 10.1016/j.energy.2021.120422
  45. 45. Xu Y-S, Zheng T, Yong X-Y, Zhai D-D, Si R-W, Li B, et al. Trace heavy metal ions promoted extracellular electron transfer and power generation by Shewanella in microbial fuel cells. Bioresource Technology. 2016;211:542-547. DOI: 10.1016/j.biortech.2016.03.144
  46. 46. Ieropoulos I, Gálvez A, Greenman J. Effects of sulphate addition and sulphide inhibition on microbial fuel cells. Enzyme and Microbial Technology. 2013;52:32-37. DOI: 10.1016/j.enzmictec.2012.10.002
  47. 47. Wen Q , Kong F, Ma F, Ren Y, Pan Z. Improved performance of air-cathode microbial fuel cell through additional tween 80. Journal of Power Sources. 2011;196:899-904. DOI: 10.1016/j.jpowsour.2010.09.009
  48. 48. Chen W, Liu Z, Li Y, Xing X, Liao Q , Zhu X. Improved electricity generation, coulombic efficiency and microbial community structure of microbial fuel cells using sodium citrate as an effective additive. Journal of Power Sources. 2021;482:228947. DOI: 10.1016/j.jpowsour.2020.228947
  49. 49. Matsena MT, Nkhalambayausi Chirwa EM. Advances in microbial fuel cell technology for zero carbon emission energy generation from waste. In: Biofuels and Bioenergy, Elsevier: Biofuels and Bioenergy. Amsterdam, The Netherlands: Elsevier; 2022. pp. 321-358. DOI: 10.1016/B978-0-323-85269-2.00013-7
  50. 50. Lin C-W, Wu C-H, Chiu Y-H, Tsai S-L. Effects of different mediators on electricity generation and microbial structure of a toluene powered microbial fuel cell. Fuel. 2014;125:30-35. DOI: 10.1016/j.fuel.2014.02.018
  51. 51. Rabaey K, Rozendal RA. Microbial electrosynthesis — Revisiting the electrical route for microbial production. Nature Reviews. Microbiology. 2010;8:706-716. DOI: 10.1038/nrmicro2422
  52. 52. Rozendal RA, Hamelers HVM, Rabaey K, Keller J, Buisman CJN. Towards practical implementation of bioelectrochemical wastewater treatment. Trends in Biotechnology. 2008;26:450-459. DOI: 10.1016/j.tibtech.2008.04.008
  53. 53. Fan X, Zhou Y, Jin X, Song R, Li Z, Zhang Q. Carbon material-based anodes in the microbial fuel cells, carbon. Energy. 2021;3:449-472. DOI: 10.1002/cey2.113
  54. 54. Sun Y, Wei J, Liang P, Huang X. Electricity generation and microbial community changes in microbial fuel cells packed with different anodic materials. Bioresource Technology. 2011;102:10886-10891. DOI: 10.1016/j.biortech.2011.09.038
  55. 55. Jung K-W, Cho S-K, Yun Y-M, Shin H-S, Kim D-H. Rapid formation of hydrogen-producing granules in an up-flow anaerobic sludge blanket reactor coupled with high-rate recirculation. International Journal of Hydrogen Energy. 2013;38:9097-9103. DOI: 10.1016/j.ijhydene.2013.05.059
  56. 56. Khilari S, Pandit S, Ghangrekar MM, Das D, Pradhan D. Graphene supported α-MnO2 nanotubes as a cathode catalyst for improved power generation and wastewater treatment in single-chambered microbial fuel cells. RSC Advances. 2013;3:7902. DOI: 10.1039/c3ra22569k
  57. 57. Xu H, Quan X, Xiao Z, Chen L. Effect of anodes decoration with metal and metal oxides nanoparticles on pharmaceutically active compounds removal and power generation in microbial fuel cells. Chemical Engineering Journal. 2018;335:539-547. DOI: 10.1016/j.cej.2017.10.159
  58. 58. Wu X, Xiong X, Owens G, Brunetti G, Zhou J, Yong X, et al. Anode modification by biogenic gold nanoparticles for the improved performance of microbial fuel cells and microbial community shift. Bioresource Technology. 2018;270:11-19. DOI: 10.1016/j.biortech.2018.08.092
  59. 59. Chen J, Hu Y, Zhang L, Huang W, Sun J. Bacterial community shift and improved performance induced by in situ preparing dual graphene modified bioelectrode in microbial fuel cell. Bioresource Technology. 2017;238:273-280. DOI: 10.1016/j.biortech.2017.04.044
  60. 60. Liu D, Chang Q , Gao Y, Huang W, Sun Z, Yan M, et al. High performance of microbial fuel cell afforded by metallic tungsten carbide decorated carbon cloth anode. Electrochimica Acta. 2020;330:135243. DOI: 10.1016/j.electacta.2019.135243
  61. 61. Wang Y, Chen Y, Wen Q , Zheng H, Xu H, Qi L. Electricity generation, energy storage, and microbial-community analysis in microbial fuel cells with multilayer capacitive anodes. Energy. 2019;189:116342. DOI: 10.1016/j.energy.2019.116342
  62. 62. Kang YL, Pichiah S, Ibrahim S. Facile reconstruction of microbial fuel cell (MFC) anode with enhanced exoelectrogens selection for intensified electricity generation. International Journal of Hydrogen Energy. 2017;42:1661-1671. DOI: 10.1016/j.ijhydene.2016.09.059
  63. 63. Liang Y, Zhai H, Liu B, Ji M, Li J. Carbon nanomaterial-modified graphite felt as an anode enhanced the power production and polycyclic aromatic hydrocarbon removal in sediment microbial fuel cells. Science of The Total Environment. 2020;713:136483. DOI: 10.1016/j.scitotenv.2019.136483
  64. 64. Yang X-Y, Tian G, Jiang N, Su B-L. Immobilization technology: A sustainable solution for biofuel cell design. Energy & Environmental Science. 2012;5:5540-5563. DOI: 10.1039/C1EE02391H
  65. 65. Luo H, Yu S, Liu G, Zhang R, Teng W. Effect of in-situ immobilized anode on performance of the microbial fuel cell with high concentration of sodium acetate. Fuel. 2016;182:732-739. DOI: 10.1016/j.fuel.2016.06.032
  66. 66. Jang JK, Pham TH, Chang IS, Kang KH, Moon H, Cho KS, et al. Construction and operation of a novel mediator- and membrane-less microbial fuel cell. Process Biochemistry. 2004;39:1007-1012. DOI: 10.1016/S0032-9592(03)00203-6
  67. 67. Ghangrekar MM, Shinde VB. Performance of membrane-less microbial fuel cell treating wastewater and effect of electrode distance and area on electricity production. Bioresource Technology. 2007;98:2879-2885. DOI: 10.1016/j.biortech.2006.09.050
  68. 68. Modestra JA, Reddy CN, Krishna KV, Min B, Mohan SV. Regulated surface potential impacts bioelectrogenic activity, interfacial electron transfer and microbial dynamics in microbial fuel cell, renew. Energy. 2020;149:424-434. DOI: 10.1016/j.renene.2019.12.018
  69. 69. Behera M, Jana PS, More TT, Ghangrekar MM. Rice mill wastewater treatment in microbial fuel cells fabricated using proton exchange membrane and earthen pot at different pH. Bioelectrochemistry. 2010;79:228-233. DOI: 10.1016/j.bioelechem.2010.06.002
  70. 70. Erable B, Duţeanu NM, Ghangrekar MM, Dumas C, Scott K. Application of electro-active biofilms. Biofouling. 2010;26:57-71. DOI: 10.1080/08927010903161281
  71. 71. Noori MT, Ghangrekar MM, Mukherjee CK, Min B. Biofouling effects on the performance of microbial fuel cells and recent advances in biotechnological and chemical strategies for mitigation. Biotechnology Advances. 2019;37:107420. DOI: 10.1016/j.biotechadv.2019.107420
  72. 72. Kugarajah V, Dharmalingam S. Effect of silver incorporated sulphonated poly ether ether ketone membranes on microbial fuel cell performance and microbial community analysis. Chemical Engineering Journal. 2021;415:128961. DOI: 10.1016/j.cej.2021.128961
  73. 73. Lovley DR, Ueki T, Zhang T, Malvankar NS, Shrestha PM, Flanagan KA, et al. Geobacter: The microbe electric’s physiology, Ecology, and practical applications. In: Advances in Microbial Physiology. Cambridge, Massachusetts: Academic Press; 2011. pp. 1-100. doi: 10.1016/B978-0-12-387661-4.00004-5
  74. 74. Biffinger JC, Byrd JN, Dudley BL, Ringeisen BR. Oxygen exposure promotes fuel diversity for Shewanella oneidensis microbial fuel cells. Biosensors & Bioelectronics. 2008;23:820-826. DOI: 10.1016/j.bios.2007.08.021
  75. 75. Biffinger JC, Ringeisen BR. Air-exposed microbial fuel cells and screening techniques. Journal of Biotechnology. 2010;150:24-25. DOI: 10.1016/j.jbiotec.2010.08.074
  76. 76. Quan X, Quan Y, Tao K. Effect of anode aeration on the performance and microbial community of an air–cathode microbial fuel cell. Chemical Engineering Journal. 2012;210:150-156. DOI: 10.1016/j.cej.2012.09.009
  77. 77. Aelterman P, Versichele M, Marzorati M, Boon N, Verstraete W. Loading rate and external resistance control the electricity generation of microbial fuel cells with different three-dimensional anodes. Bioresource Technology. 2008;99:8895-8902. DOI: 10.1016/j.biortech.2008.04.061
  78. 78. Lyon DY, Buret F, Vogel TM, Monier J-M. Is resistance futile? Changing external resistance does not improve microbial fuel cell performance. Bioelectrochemistry. 2010;78:2-7. DOI: 10.1016/j.bioelechem.2009.09.001
  79. 79. Picioreanu C, Head IM, Katuri KP, van Loosdrecht MCM, Scott K. A computational model for biofilm-based microbial fuel cells. Water Research. 2007;41:2921-2940. DOI: 10.1016/j.watres.2007.04.009
  80. 80. Katuri KP, Scott K, Head IM, Picioreanu C, Curtis TP. Microbial fuel cells meet with external resistance. Bioresource Technology. 2011;102:2758-2766. DOI: 10.1016/j.biortech.2010.10.147
  81. 81. Hashemi J, Samimi A. Steady state electric power generation in up-flow microbial fuel cell using the estimated time span method for bacteria growth domestic wastewater. Biomass and Bioenergy. 2012;45:65-76. DOI: 10.1016/j.biombioe.2012.05.011
  82. 82. Lay C-H, Kokko ME, Puhakka JA. Power generation in fed-batch and continuous up-flow microbial fuel cell from synthetic wastewater. Energy. 2015;91:235-241. DOI: 10.1016/j.energy.2015.08.029
  83. 83. Haavisto JM, Kokko ME, Lay C-H, Puhakka JA. Effect of hydraulic retention time on continuous electricity production from xylose in up-flow microbial fuel cell. International Journal of Hydrogen Energy. 2017;42:27494-27501. DOI: 10.1016/j.ijhydene.2017.05.068
  84. 84. Lu L, Ren N, Zhao X, Wang H, Wu D, Xing D. Hydrogen production, methanogen inhibition and microbial community structures in psychrophilic single-chamber microbial electrolysis cells. Energy & Environmental Science. 2011;4:1329. DOI: 10.1039/c0ee00588f
  85. 85. Cheng S, Xing D, Logan BE. Electricity generation of single-chamber microbial fuel cells at low temperatures. Biosensors & Bioelectronics. 2011;26:1913-1917. DOI: 10.1016/j.bios.2010.05.016
  86. 86. Boas JV, Oliveira VB, Simões M, Pinto AMFR. Review on microbial fuel cells applications, developments and costs. Journal of Environmental Management. 2022;307:114525. DOI: 10.1016/j.jenvman.2022.114525
  87. 87. Xue W, Li F, Zhou Q. Degradation mechanisms of sulfamethoxazole and its induction of bacterial community changes and antibiotic resistance genes in a microbial fuel cell. Bioresource Technology. 2019;289:121632. DOI: 10.1016/j.biortech.2019.121632
  88. 88. Chen J, Yang Y, Liu Y, Tang M, Wang R, Tian Y, et al. Bacterial community shift and antibiotics resistant genes analysis in response to biodegradation of oxytetracycline in dual graphene modified bioelectrode microbial fuel cell. Bioresource Technology. 2019;276:236-243. DOI: 10.1016/j.biortech.2019.01.006
  89. 89. Liu R, Gao C, Zhao Y-G, Wang A, Lu S, Wang M, et al. Biological treatment of steroidal drug industrial effluent and electricity generation in the microbial fuel cells. Bioresource Technology. 2012;123:86-91. DOI: 10.1016/j.biortech.2012.07.094
  90. 90. Jia J, Tang Y, Liu B, Wu D, Ren N, Xing D. Electricity generation from food wastes and microbial community structure in microbial fuel cells. Bioresource Technology. 2013;144:94-99. DOI: 10.1016/j.biortech.2013.06.072
  91. 91. Zhang B, Tian C, Liu Y, Hao L, Liu Y, Feng C, et al. Simultaneous microbial and electrochemical reductions of vanadium (V) with bioelectricity generation in microbial fuel cells. Bioresource Technology. 2015;179:91-97. DOI: 10.1016/j.biortech.2014.12.010
  92. 92. Wu Y, Zhao X, Jin M, Li Y, Li S, Kong F, et al. Copper removal and microbial community analysis in single-chamber microbial fuel cell. Bioresource Technology. 2018;253:372-377. DOI: 10.1016/j.biortech.2018.01.046
  93. 93. Jiang J, Wang H, Zhang S, Li S, Zeng W, Li F. The influence of external resistance on the performance of microbial fuel cell and the removal of sulfamethoxazole wastewater. Bioresource Technology. 2021;336:125308. DOI: 10.1016/j.biortech.2021.125308
  94. 94. Park Y, Park S, Nguyen VK, Yu J, Torres CI, Rittmann BE, et al. Complete nitrogen removal by simultaneous nitrification and denitrification in flat-panel air-cathode microbial fuel cells treating domestic wastewater. Chemical Engineering Journal. 2017;316:673-679. DOI: 10.1016/j.cej.2017.02.005
  95. 95. Zhao H, Kong C-H. Enhanced removal of p-nitrophenol in a microbial fuel cell after long-term operation and the catabolic versatility of its microbial community. Chemical Engineering Journal. 2018;339:424-431. DOI: 10.1016/j.cej.2018.01.158
  96. 96. Li H, Zhang S, Yang X-L, Yang Y-L, Xu H, Li X-N, et al. Enhanced degradation of bisphenol and ibuprofen by an up-flow microbial fuel cell-coupled constructed wetland and analysis of bacterial community structure. Chemosphere. 2019;217:599-608. DOI: 10.1016/j.chemosphere.2018.11.022
  97. 97. Li C, Luo M, Zhou S, He H, Cao J, Luo J, et al. Study on synergistic mechanism of PANDAN modification, current and electroactive biofilms on Congo red decolorization in microbial fuel cells. International Journal of Hydrogen Energy. 2020;45:29417-29429. DOI: 10.1016/j.ijhydene.2020.07.246
  98. 98. 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. Science of The Total Environment. 2019;656:39-44. DOI: 10.1016/j.scitotenv.2018.11.326

Written By

Hajar Rajaei Litkohi and Hosein Yazdi Dehnavi

Submitted: 01 December 2023 Reviewed: 02 December 2023 Published: 16 April 2024