Vous êtes sur la page 1sur 17

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/309278331

Microbial Fuel Cells

Chapter · October 2016


DOI: 10.1016/B978-0-12-409548-9.10122-8

CITATIONS READS
3 2,106

4 authors:

Shantonu Roy Stefania Marzorati


Indian Institute of Technology Kharagpur Università degli Studi di Milano
37 PUBLICATIONS   312 CITATIONS    21 PUBLICATIONS   121 CITATIONS   

SEE PROFILE SEE PROFILE

Andrea Schievano Deepak Pant


University of Milan Flemish Institute for Technological Research
63 PUBLICATIONS   1,174 CITATIONS    157 PUBLICATIONS   5,039 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

CO2ElectroRefinery View project

Artificial photosynthesis View project

All content following this page was uploaded by Deepak Pant on 25 September 2017.

The user has requested enhancement of the downloaded file.


Author's personal copy

Provided for non-commercial research and educational use.


Not for reproduction, distribution or commercial use.

This article was originally published in the online Encyclopedia of Sustainable


Technologies, published by Elsevier, and the attached copy is provided by Elsevier for
the author's benefit and for the benefit of the author's institution, for non-commercial
research and educational use including without limitation use in instruction at your
institution, sending it to specific colleagues who you know, and providing a copy to your
institution's administrator.

All other uses, reproduction and distribution, including


without limitation commercial reprints, selling or
licensing copies or access, or posting on open
internet sites, your personal or institution’s website or
repository, are prohibited. For exceptions, permission
may be sought for such use through Elsevier’s
permissions site at:
https://www.elsevier.com/about/our-business/policies/copyright/permissions

From Roy, S., Marzorati, S., Schievano, A., Pant, D., 2017. Microbial Fuel Cells. In:
Abraham, M.A. (Ed.), Encyclopedia of Sustainable Technologies. Elsevier, pp.
245–259.
ISBN: 9780128046777
Copyright © 2017 Elsevier Inc. All rights reserved.
Elsevier
Author's personal copy

Microbial Fuel Cells


S Roy, Indian Institute of Technology, Kharagpur, India
S Marzorati and A Schievano, University of Milan, Milan, Italy
D Pant, VITO – Flemish Institute for Technological Research, Mol, Belgium
Ó 2017 Elsevier Inc. All rights reserved.

Microbial Fuel Cells: Harvesters of Chemical Energy

The detrimental effects of carbon emissions to the environment and natural resource depletion have recently caused a general awak-
ening, promoting the design and use of greener technologies. Environmental concerns and energy security are expected to be major
drivers of future markets. Microbial fuel cell (MFC) could offer carbon-neutral alternatives in this field. MFCs utilize organic
substrates and subsequently convert their chemical energy to electricity in a single step using microorganisms. There has been an
increasing interest for the production of nongrid electricity from decentralized diffused sources (sunlight, biomass, etc.), which
can be produced at the site of requirement and cater to the local needs. In this respect, in the future, MFCs may find potential appli-
cations such as decentralized wastewater treatment, bioremediation, and renewable energy harvesting to supply off-grid and remote
devices (Bentley, 2002).
It is possible to directly generate electricity using bacteria while accomplishing wastewater treatment or natural water bioreme-
diation in processes based on MFC technologies.
MFCs are bioelectrochemical transducers that convert organic matter into electricity, using bacteria as catalyst. In a broader sense,
MFCs are devices capable of directly transforming chemical energy into electrical energy via electrochemical reactions involving
biochemical pathways and biological enzymatic catalysis. MFCs offer many advantages over the traditional fuel cells and enzymatic
fuel cells. Use of bacteria or a consortium of bacteria as catalyst provides the possibility of using a wide range of microbially degrad-
able organic or inorganic matter such as organic waste and soil sediments. The direct conversion of substrate energy to electricity
enables high conversion efficiency. Moreover, they are operated at ambient temperature and atmospheric pressure in contrast to
the fuel cells that need high temperature and pressure (Rabaey et al., 2009).
MFCs have operational and functional advantages over the technologies currently used for generating energy from organic
matter (Virdis, 2011). First, the direct conversion of substrate energy to electricity enables high conversion efficiency. Second,
MFCs operate efficiently at ambient and even at low temperatures, distinguishing them from all current bioenergy processes. Third,
an MFC does not require gas treatment because the off-gases of MFCs are enriched in carbon dioxide and normally have no useful
energy content. Fourth, MFCs do not need energy input for aeration provided the cathode is passively aerated. Fifth, MFCs have
potential for widespread application in locations lacking electrical infrastructures and also to expand the diversity of fuels we
use to satisfy our energy requirements.

Fundamental Principles of Voltage/Current Generation in MFC


Working Principle of MFC
An MFC can be divided into three major components, namely, an anaerobic anode chamber, an aerobic cathode chamber, and sepa-
rator connecting the two chambers, which has to guarantee proton exchange. Anode chamber provides all the necessary conditions
for the growth and the electron extraction from microorganisms. The chamber is fed with growth media named as anolyte,
microorganisms, and an electrode that acts as the anode. The oxidative microbial metabolism in this chamber produces protons
and electrons. The metabolic reactions are not allowed to proceed to completion and the intermediate electrons are drawn from
the cell to produce electrical work. Electrons are transferred by microbes to the extracellular acceptor (anode) and flow to the
cathode through a resistor, producing electricity. The protons migrate through the electrolyte to the cathode and participate to
the oxygen reduction reaction (ORR) to OH or H2O, accepting electrons from the cathode. Typical electrode reactions are shown
in the succeeding text using acetate as a model substrate:
Microbes
Anodic reaction : CH3 COO þ 2H2 O / 2CO2 þ 7Hþ þ 8e (1)

Cathodic reaction : 2O2 þ 8Hþ þ 8e / 4H2 O (2)

DG ¼ 847:60 kJ mol1 ; emf ¼ 1:10 V (3)


Anaerobic metabolism must be promoted at the anode in order to convert organic matter to electricity in an effective manner.
The key difference in microbial electricity production versus natural biogeochemical processes, such as Fe3 þ reduction, is that the
electrons are transferred to a conductive but chemically stable material, rather than to a natural electron acceptor. Electricity is gener-
ated since the overall reaction is thermodynamically favorable (Eq. 3).

Encyclopedia of Sustainable Technologies, Volume 3 http://dx.doi.org/10.1016/B978-0-12-409548-9.10122-8 245

Encyclopedia of Sustainable Technologies, First Edition, 2017, 245–259


Author's personal copy
246 Microbial Fuel Cells

Performance Evaluation for MFC


MFC performance and efficiency are generally evaluated in terms of electrical parameters (power density, current density, potential
difference, and internal resistance) and biodegradation efficiency (removal of organics). Coulombic efficiency (CE) is the parameter
that combines the two previous ones: the ratio between the amount of electrons harvested as current and total equivalents con-
tained in the organic matter available to microbial oxidative metabolism (generally measured as Chemical oxygen demand
(COD)) (Logan, 2009).

Polarization and Power Density Curves


Polarization (i.e., voltage vs. current) curves have become the standard method of presenting MFC performances (as for chemical
fuel cells). Not only does this show the maximum current and open-circuit voltage of the cell, but also it shows characteristics of the
voltage–current behavior, which give predominant indications on internal resistances. There is considerable variation in methods of
reporting these curves. In few studies, a short potential sweep is done that does not allow the system to reach steady state at any one
voltage or current (often leading to overreporting of performance), while others use a fixed voltage or resistance until steady state is
reached.
The voltage measured on contact with electron donor (U) differs from the electromotive force of the MFC operating (E). This is
due to the MFC polarization and ohmic power failure (due to internal resistance) and represented as Eq. 4:
U ¼ E  DE  Ir (4)
Polarization of an element consists of polarization of the anode and the cathode, that is, DE ¼ DEa þ DEc.
For efficient MFC performance, the magnitude of the polarization element DE needs to be minimized. The polarization
elements can be reduced by the following steps: (a) the use of effective catalysts that facilitate electromechanical reactions,
(b) increment of the operating temperatures, (c) increasing the true surface of the electrode, and (d) increasing the concentration
of the reactants.
Typically, to initiate electron transport by electron donor oxidation to the anode (electron acceptor), it must overcome an energy
barrier. This energy barrier comprises an additional potential known as activation overpotential. The activation overpotential is
responsible to drive the transfer of electrons from source to microbial shuttles and eventually to the anode. In Fig. 1, the zone
of activation losses is shown. A sharp decrease in MFC voltage (low polarization) at the initial low current densities was observed.
A subsequent ohmic and concentration loss was observed at intermediate or high current densities. Moreover, it was evident that
activation losses occur at both anode and the cathode. It is important to note that the cathodic overpotentials are much larger than
anodic overpotentials.

Fig. 1 Typical anodic and cathodic polarization curves in MFCs, as compared with theoretical trends.

Encyclopedia of Sustainable Technologies, First Edition, 2017, 245–259


Author's personal copy
Microbial Fuel Cells 247

The Current Interrupt Method


This method is widely used for ohmic drop and ohmic resistance evaluation in electrochemical systems, including batteries,
corrosion, and fuel cells. The internal resistance of MFCs, also, can be calculated by the current interrupt method (Hosseini and
Ahadzadeh, 2013). While in closed circuit mode, once the MFC produces a stable current output (I) and potential (V), the circuit
is opened causing a steep initial rise in the cell voltage (VR) followed by gradual further increment. The steep rise is attributed to
ohmic losses caused by the internal resistance (Rint) of the MFC and can be hence calculated as (Eq. 5)
VR
Rint ¼ (5)
I
However, this method precludes the differentiation of the charge transfer resistance, the double-layer capacitance at the elec-
trodes, and the mass transfer resistance. A complete understanding of electrochemical constraints requires a detailed investigation
of ohmic, kinetic, and mass transport properties. Internal resistance (Rint) mainly consists of three parts: charge transfer/polariza-
tion/activation resistance (Rp), ohmic/solution resistance (Rohm), and concentration/diffusion resistance (Rd) (Zhao et al., 2009).
The internal resistance in MFCs originates from electrode materials, catalyst coatings on electrodes, poor biofilm development, and
electrochemical reactions on the anodes and the cathodes. Electrochemical impedance spectroscopy (EIS) is a powerful nondestruc-
tive technique that has been used for analyzing the distinct contributions to internal resistance in MFCs.

Electrochemical Impedance Spectroscopy


EIS is a useful method for quantifying the different components of internal resistance (solution resistance, charge transfer resis-
tances, and diffusion resistance) in electrochemical systems (Aelterman et al., 2006). Ohm’s law defines resistance (R) in terms
of voltage (E) and current (I). While Ohm’s law is limited to an ideal resistor, circuit elements in the real world exhibit much
more complex behavior. Impedance is a measure of the ability of a circuit to resist the flow of electrical current. Impedance is
measured in EIS by applying a small excitation signal that is an AC (alternating current) potential to an electrochemical system
and reading the AC response. Impedance (Z) is expressed analogously to Ohm’s law as follows (Eq. 6):
Et E0 sinðut Þ sinðut Þ
Z¼ ¼ ¼ Z0 (6)
It I0 sinðut þ 4Þ sinðut þ 4Þ

where Et is the potential at time t, E0 is the amplitude of the signal, u is the radial frequency, 4 is the phase shift, It is the response
signal at time t, and I0 is the amplitude of the response. With Euler’s relationship, it becomes simplified to (Eq. 7)
z ¼ z0 ðcos4 þ jsin4Þ ¼ ReZ þ ImZ (7)
Impedance is composed of a real part (ReZ ¼ Z0cos4) and an imaginary part (ImZ ¼ Z0jsin4), where j is the imaginary unit. Each
point on a Nyquist plot is the imaginary impedance versus the real impedance at each frequency. Higher frequencies are on the left
side and low frequencies are on the right side in the Nyquist plot.
Information about the electrochemical reactions that occur on electrodes and the surface and material properties of electrodes
can be obtained in the EIS measurement of an individual electrode (He and Mansfeld, 2009). Manohar et al. used EIS to evaluate the
electrochemical behavior of the anode and the cathode of a mediator-less MFC (Manohar and Mansfeld, 2009). They considered the
internal resistance can be determined through data of the impedance spectra at different cell voltages.

Cyclic Voltammetry
Cyclic voltammetry (CV) is an important method to determine the mechanisms of electrode reactions underlying oxidation or
reduction reactions on the electrode surface by measuring the current response at an electrode surface to a specific range of applied
potentials. When the voltage at the electrode surface reaches a species’ standard reduction or oxidation potential, an increase in
current can be seen as peaks in the voltammogram (current vs. applied voltage curve). Depending on whether the electrode is being
reduced or oxidized, electrons will be transferred either to or from the electrode surface, respectively (Fricke et al., 2008).
It requires a three-electrode (working electrode, reference electrode, and counter electrode) configuration to obtain accurate
results. The characteristics of CV depend on several factors, such as the electrode surface pretreatment, the rate of the electron transfer
reactions, the chemical and biological species present and their thermodynamic properties, the concentration of electroactive
species and their rates of diffusion, and the potential scan rate (Zhao et al., 2009).
In MFCs, CV experiments were carried out extensively to (i) elucidate the exocellular electron transfer mechanisms of anode reac-
tions involving both direct and indirect electron transfer between the biofilm and the electrode, (ii) determine the redox potentials
of the chemical or biological species involved at the anode, and (iii) assess the performance of the catalysts being studied for ORR or
anode surface modification (Zhao et al., 2009).
MFC studies employing CV generally use forward and backward voltage sweeps with rates in the range of 1–20 mV s 1. Due to
the presence of several different redox species in MFC, multiple peaks in the cyclic voltammograms may be observed (Harnisch and
Freguia, 2012).

Encyclopedia of Sustainable Technologies, First Edition, 2017, 245–259


Author's personal copy
248 Microbial Fuel Cells

Charge represents the number of electrons present at particular instance on the anode surface during oxidation. Charge separa-
tion can also be represented by capacitance, which is denoted as the ability of fuel cell to hold an electrical charge. The amount of
electrical energy stored at an applied electric potential is given by the following relation (Eq. 8):
C¼QV (8)

where Q is the charge (coulomb), C is the capacitance (farad), and V is the maximum voltage applied (volt). The biofilm developed
on the anode accumulates higher charge where the electrons are held on the electrode and slowly get transferred into the circuit. As
charge is directly proportional to capacitance, increase in charge also concurrently increases the capacitance. High capacitance
suggests the higher availability of electrons due to the conversion of metabolite substrate at anode:
W ¼ Vmax  Q (9)

The energy (W) generated during potential sweep is calculated using Eq. 9. The energy conversion is also directly proportional to
the charge obtained during operation. The charge, capacitance, and energy profile increased linearly with the increment in organic
load of wastewater.

CE and Energy Efficiency


The CE is defined as the ratio of the total coulombs actually transferred to the anode from the substrate. The maximum possible
coulomb value is reached if all substrate removal produces current. The total coulombs obtained are determined by integrating
the current over time, so that the CE for an MFC runs in fed-batch mode (Logan et al., 2006). The CE is diminished by utilization
of alternate electron acceptors by the bacteria, either those present in the medium (or wastewater) or those diffusing such as oxygen.
Other factors that reduce CE are competitive processes and bacterial growth. Bacteria unable to utilize the electrode as electron
acceptor are likely to use substrate for fermentation and/or methanogenesis. It has been observed that fermentative patterns
diminish through time during enrichment of the microbial consortium in the MFC. As long as the anode remains attractive enough
for the bacteria due to its potential, alternative electron acceptors will not be used. The most important factor for evaluating the
performance of an MFC for making electricity, compared with more traditional techniques, is to evaluate the system in terms of
the energy recovery (Rabaey et al., 2009). The energy efficiency of an MFC can be defined as the energy recovered from the organic
matter to the total energy content of the organic material. In MFCs, energy efficiencies range from 2% to more than 10% depending
on the type of substrate (Virdis, 2011).

Exoelectrogens and the Mechanism of Electron Transfer


Mediator-Based Electron Transfer Mechanism
Low-molecular-weight redox species may assist the shuttling of electrons between the intracellular bacterial species and an elec-
trode. However, there are many important requirements that such a mediator should satisfy in order to provide an efficient electron
transport from the bacterial metabolites to the anode:
(a) The oxidized state of the mediator should easily penetrate the bacterial membrane to reach the reductive species inside the
bacterium.
(b) The redox potential of the mediator should fit the potential of the reductive metabolite (the mediator potential should be
positive enough to provide fast electron transfer from the metabolite, but it should not be so positive as to prevent significant
loss of potential).
(c) Oxidation state of the mediator should interfere with other metabolic processes (should not inhibit them or be decomposed by
them).
(d) The reduced state of the mediator should easily escape from the cell through the bacterial membrane.
(e) Both oxidation states of the mediator should be chemically stable in the electrolyte solution; they should be well soluble and
not get adsorbed on the bacterial cells or electrode surface.
The mediators can be coupled to the microorganisms in three different ways:
(i) A diffusional mediator shuttling between the microbial suspension and the anode surface
(ii) A diffusional mediator shuttling between the anode and microbial cells covalently linked to the electrode
(iii) Mediator adsorbed on the microbial cells providing electron transport from the cells to the anode
The microbial cells can be covalently linked to the electrode surface having eCOOH groups and amino groups of the microbial
membrane resulting in the formation of amide bond. Standard organic reagents like carbodiimide and acetyl chloride are used
to link the microbial cells to the surface.

Microorganisms That Produce Their Own Mediators


In some instances, microorganisms might produce their own mediators to promote extracellular electron transfer. It was first
proposed as a mechanism to facilitate electron transfer to Fe3 þ in Shewanella oneidensis and later in organisms such as Geothrix

Encyclopedia of Sustainable Technologies, First Edition, 2017, 245–259


Author's personal copy
Microbial Fuel Cells 249

Fig. 2 (A) An indirect microbial fuel cell where nonelectrogenic fermentative microorganisms can be used along with external mediators that
partially recover the electrons available. (B) Microbes can produce soluble mediators in a mediator-driven microbial fuel cell. In both cases, mediators
get reduced by accepting electrons from cell and oxidized by donating electrons to anode.

fermentans and Pseudomonas species. Biosynthesizing an electron shuttle is energetically expensive, and therefore, an electron shuttle
must be recycled many times in order to utilize this energy investment. For this reason, microorganisms that produce electron shut-
tles were expected to be at a competitive disadvantage in open environments in which the shuttle will rapidly be lost from the site of
release (Fig. 2). The same considerations are likely to apply to MFCs. Electron shuttles were produced in an MFC using glucose as the
substrate. Pseudomonas species and Pseudomonas aeruginosa isolated from the fuel cell produced phenazine electron shuttles that
could aid in electron transfer to electrodes. Releasing an electron shuttle might be an adaptive strategy under these conditions
because the shuttle is not lost from the system. In pure culture studies, a one-time replacement of the medium from a Geothrix
fuel cell decreased power production by 50% (Bond and Lovely, 2005). If an electron shuttle was continually being flushed
from the system, this would result in a net energetic loss to the organism.

Electron Transfer in Mediator-Less MFC


In mediator-less MFC, microbes are “anode-respiring”; therefore, they do not require any mediators to transfer electrons to anode
during respiration. Biochemical and genetic characterizations indicated that outer-membrane cytochrome can be involved in exog-
enous electron transfer. Microorganisms such as members of the Geobacter family, Shewanella putrefaciens and Shewanella oneidensis
(Kim et al., 2002), Rhodoferax ferrireducens (Liu et al., 2007), Pseudomonas aeruginosa (Dantas et al., 2013), Clostridium butyrium (Park
et al., 2001), and Aeromonas hydrophila (Pham et al., 2003) have been reported to oxidize organic matter by donating electrons
directly to the anode to complete their metabolism process.
The initial findings of power generation in MFC were limited to bacteria capable of dissimilatory iron reduction. However, it was
proved that MFCs consisted of diverse microbial communities. Microorganisms belonging to Proteobacteria, as well as the Firmi-
cutes and Acidobacteria phyla, were well known for their current generation ability. The yeast Pichia anomala was reported to have
redox enzymes on its outer membrane that helped in current production in an MFC (Prasad et al. 2007). The oxygenic phototrophic
cyanobacterium Synechocystis sp. PCC 6803 was discovered to produce electrically conductive appendages called nanowires (Gorby
et al., 2006).
The most important mechanism is the direct electron transfer of the microorganisms to the electrodes. When cultures of Shewa-
nella putrefaciens produced electricity while metabolizing lactate, it was first proposed that microorganisms might be able to transfer
electrons directly to an electrode surface. The metal-reducing bacterium Shewanella putrefaciens MR-1 was reported to have cyto-
chromes in its outer membrane (Harris et al., 2010). Under anaerobic conditions, these electron carriers (i.e., cytochromes) were

Encyclopedia of Sustainable Technologies, First Edition, 2017, 245–259


Author's personal copy
250 Microbial Fuel Cells

Fig. 3 Intracellular electron transfer mechanisms. Components involved in the electron transport from cells to the anode in MFC, in metal-reducing
microorganisms (e.g., Shewanella putrefaciens).

able to generate anodic current in the absence of terminal electron acceptors. Further, it was also recognized that outer-membrane
cytochromes were important in electron shuttle reduction in Shewanella (Meitl et al., 2009).
The mediator-less electron transfer in an MFC is also known as direct electron transfer. This is found in certain microorganisms
(termed “electricigen” or “exoelectrogen” or “electrogen” or “anodophiles” or “anode-respiring bacteria” or “electrochemically
active bacteria, EAB”) that can transfer the final electrons coming from the oxidized organic matter directly to the electrode of
the MFC. This step of electron transfer to the electrode is the key point of an MFC technology. Similarities between microbial reduc-
tion of insoluble metal oxides and current production in MFCs have been found in both the cases; the final electrons are transported
to an extracellular solid substrate. This transport may occur either in direct contact between the cell surface and the solid substrate or
indirectly by endogenous mediators.
Thus, direct electron transfer from microorganisms to the electrode is again of two types:
(1) Involving endogenous mediator secretion by the exoelectrogen, which mediates the electron transfer between the bacteria and
the electrode.
(2) Involving direct electron transfer to the electrode via pili or “nanowire” after the formation of a biofilm on the electrode by the
bacteria.
(3) Bacteria grow by catalyzing chemical reactions and harnessing and storing energy in the form of adenosine triphosphate (ATP).
(4) In some bacteria, reduced substrates are oxidized, and electrons are transferred to respiratory enzymes by NADH, the reduced
form of nicotinamide adenine dinucleotide (NAD).
Fig. 3 illustrates the chemical components proposed to be involved in electron transportation from electron carriers in the intra-
cellular matrix to the solid-state final electron acceptor (anode) in the dissimilatory metal-reducing microorganisms (Holmes
et al., 2004). Electron transfer from the inner membrane to outer membrane in EAB appears to involve an electron transport chain
of inner-membrane, periplasmic, and outer-membrane c-type cytochrome (Lovley et al., 2004).
Electrically conductive bacterial appendages known as nanowire have only recently been discovered so their structures are there-
fore not well studied or understood (Gorby et al., 2006). Pili produced by some bacteria have so far been shown to be electrically
conductive using scanning tunneling electron microscopy. However, more research is required to establish the role of nanowire in
active electron transfer.

Configurations and Materials Used in MFCs

MFC architecture has become recently considerably flexible in order to meet the large number of possible in-lab and on-field appli-
cations. Due to the possible upcoming applications, many designs are under investigation, and infinite different cell architectures
could in principle be proposed, with different combinations of MFC constituents. In this section, a possible overview of the state-of-
the-art MFC devices is given.
MFC devices have been extensively studied in a range of different configurations (Verstraete and Rabaey, 2006; Logan, 2008.).
First of all, MFCs can be formed of single, double, or triple chamber (Fig. 4A–C).

Encyclopedia of Sustainable Technologies, First Edition, 2017, 245–259


Author's personal copy
Microbial Fuel Cells 251

Fig. 4 Schematic representation of (A) double-chamber MFC with proton exchange membrane (PEM), (B) single-chamber MFC (open air cathode),
and (C) triple-chamber configuration, with cationic exchange membrane (CEM) and anion exchange membrane (AEM), used in microbial desalination
cells (MDCs). (D) Examples of electrode materials used for MFC: (i) carbon cloth, (ii) graphite felt, (iii) carbon brush, (iv) granular graphite, and
(v) stainless steel mesh.

Mostly used in labs are double-chamber MFCs (Du et al., 2007). They include the anodic and the cathodic compartments and
a separator (Fig. 4). Proton (or ion) exchange membranes (PEMs), salt bridges, or septa with different ranges of porosity have the
function of allowing protons to move across to reach the abiotic aqueous cathodic compartment where dissolved ORR takes place
(Min and Logan, 2004). The cathodic chamber (an aqueous solution) is kept at constant and maximum oxygen concentrations by
continuous air bubbling or by the presence of photosynthetic organisms (Gajda et al., 2013).
In membrane-less MFC (single chamber), the cost of the separating membrane is avoided and oxygen supply at cathode is guar-
anteed by direct air diffusion to the cathode (Liu and Logan, 2004; Logan and Regan, 2006). Air cathodes expose their outer side
directly to air and their inner side to the solution, as presented in Fig. 4. Anode and cathode are integrated in a unique compartment,
in contact with the same solution. This contributes in reducing the internal resistance and the overpotentials, as compared with
double-chamber systems (Cristiani et al., 2013). Microbial biofilms grow on both the anode and the inner side of the cathode.
The cathodic biofilm helps in preserving anaerobic conditions in the anodic chamber. Oxygen diffuses only in a relatively thin layer
and forms a decreasing gradient toward the anaerobic environment of the anodic compartment (Cristiani et al., 2013; Daghio et al.,
2015).
Triple-chamber MFCs are used for a specific purpose, water desalination. As shown in Fig. 4, they comprise an anodic compart-
ment, where microbes oxidize organic matter transferring electrons to the anode, a middle desalination chamber limited on one
side by a PEM and on the other side by an anion exchange membrane (AEM), and a cathodic compartment where oxygen is reduced
(Kim and Logan, 2013). The osmotic force generated by the electrical field drags dissolved salts from the central chamber toward the
two membranes.

Anodic Compartment
The anode must be simultaneously an excellent charge collector and solid matrix favoring microbial biofilm attachment. Some
fundamental properties are required:
l Biocompatibility
l Electronic conductivity
l Chemical stability over time and against corrosion/biodegradation
l High volumetric surface area

Encyclopedia of Sustainable Technologies, First Edition, 2017, 245–259


Author's personal copy
252 Microbial Fuel Cells

l High surface porosity to favor biofilm attachment


l Low cost (especially for large-scale applications)
The volumetric surface area must be as high as possible (e.g., 18 200 m2 m 3 (Logan et al., 2007)), to optimize the microbial
biofilm interactions with the anolyte and the oxidation of the organic matter. Anodes should occupy the largest volume of the
anodic chamber; specific shapes are intended to maximize the capillary contact with the organic matter in the solution. Conductive
brushes, felts, or granular materials are typical anode geometries, allowing a high volumetric surface area and porosity to favor elec-
trogenic biofilm activity (Logan et al., 2007).
Anode chemical composition is the second key factor. Meshes or felts based on stable metals (e.g., stainless steel and titanium)
have been tested (Logan, 2010); however, to guarantee chemical stability against biocorrosion, the cost of the material is generally
too high for large-scale applications. Additionally, metals have low porosity and don’t favor biofilm attachment.
Carbonaceous materials are one of the best candidates to balancing costs, chemical stability, and biocompatibility. Carbon or
graphite fibers, cloths, felt, and brush are often used because of their high conductivity, high stability, and relatively acceptable cost.
Carbon cloth and carbon felt are excellent substrates for biofilm anodic growth; however, for their relatively high costs, carbon mesh
is often preferred. Graphite fiber brushes are made of a stainless steel or titanium core as current collector and a dense framework of
carbon fibers; these shapes are one of the most promising, due to their high surface areas (18 200 m2 m 3 (Logan et al., 2007))
(Fig. 4D).
To improve the catalytic characteristics, attempts in treating carbon-based anodes were shown to be very effective; when the
carbon cloth was treated with a 5% ammonia solution, the catalytic performance increased reaching a power density of
1970 mW m 2 as shown by Logan et al. (2007). Finally, a few types of conductive polymers have been investigated as coatings
onto anode materials. However, results in MFC performances were not excellent, reaching a voltage of 55% compared with a carbon
cloth anode (Logan, 2008).

Cathodic Compartment
One of the main reasons of hindrance of real-world applications of MFC is, in particular, the costs and design of the cathodic elec-
trode. The challenge is allowing oxygen to continuously and abundantly reach the electrocatalyst for the ORR. Air cathodes are often
preferred because they allow a direct contact of the solution and the conductor with air (three-phase interface: gas–liquid–solid),
without any need of forced systems to enhance oxygen availability at the catalyst layer. Air cathodes must be porous in order to let
oxygen diffuse through the electrode. This is often achieved by a gas diffusion layer (GDL) or a gas diffusion membrane. GDLs are
generally layers made of the same material of the electrode (carbon fibers and activated carbon powder), mixed at different ratios
with binding materials (e.g., polytetrafluoroethylene and perfluorosulfonic acid) (Guerrini et al., 2015).
Cathodes can catalyze the ORR by both abiotic and biotic mechanisms. When an abiotic mechanism takes place exposing the
electrode to a bacteria-free solution or, more simply, to air, the same issues of fuel cells have to be overcome: ORR is kinetically
hindered and requires very effective catalysts to meet the power requirements to provide enough energy. Pt-based catalysts and
Pt alloys are the most credited for this issue; however, due to precious metal high costs and availability, they need to be substituted.
High power densities (2500 mW m 2) have been achieved in literature by Park et al., with highly efficient catalysts like Pt-loaded
carbon (Kumar et al., 2013). In a comparative study, Santoro et al. demonstrated that Pt-based cathodes had higher power gener-
ation than those of the Pt-free cathodes at the start-up period, but the difference quickly vanished after a few days of operation
(Santoro et al., 2013). The frontiers of MFC cathodes are nowadays expanding in the direction of Pt-free electrodes. Carbon-
based materials, doped with heteroatoms and nonprecious metals, were demonstrated to be good candidates for the purpose
(Marzorati et al., 2015).
Especially in single-chamber MFCs, the efficiency of abiotic catalysts is strongly limited to relatively short-term operation (some
weeks as maximum) of devices, due to microbial growth on cathodic surface. Biofouling tends to limit the catalytic efficiency
(Santoro et al., 2012). After biofilm formation, biotic mechanisms tend to substitute chemical catalysis in the ORR (Zhao et al.,
2006). The cathode becomes a biocathode, where microbial biofilm directly acts as catalyst. In some cases, extreme pH conditions
at the cathode side (caustic formation due to ORR) avoid microbial growth and keep abiotic conditions (Gajda et al., 2014).
In single-chamber MFCs, microbial catalysis takes place after formation of a cathodic biofilm, which is exposed organic
substances and oxygen, diffused from the porous layer (GDL) (Squadrito and Cristiani, 2016). Cathodic biofilms were shown to
behave as really effective catalyzers of the ORR (Cristiani et al., 2013). Carbon, in the form of activated mesoporous granules or
more sophisticated structures (i.e., nanofibers, nanotubes, and powder), offers good biocompatibility, high surface area, and
low costs (Santoro et al., 2014). Sometimes, these carbon-based materials are linked to a metal mesh current collector, such as
Ni or stainless steel mesh (Logan, 2010).

Membranes and Separators


Membranes are used to keep the anodic biotic chamber separated from the cathodic, to allow maintaining a “clean” and bacteria-
free solution at cathode (Liu and Logan, 2004). PEMs are selectively permeable to protons while avoiding other species to be
transferred from a compartment to the other. Cation exchange membranes (CEMs) and AEMs have been extensively used in
double-chamber and triple-chamber MFCs. The most commonly used CEMs are Nafion, Hyflon, Zirfon, and Ultrex CMI 7000

Encyclopedia of Sustainable Technologies, First Edition, 2017, 245–259


Author's personal copy
Microbial Fuel Cells 253

(Leong et al., 2013). Nafion, first developed for hydrogen fuel cells, is not only the most promising but also the most expensive
among the series; it can cost up to $1400 m 2. The main problem of CEM is pH splitting between the anode and cathode chambers
caused by proton accumulation in the anodic compartment. Attempts to overcome this problem led to the development of AEM
that conducts anions from the cathode to the anode chamber (Kim et al., 2007).
At least two important advantages are achieved when the membrane is conveniently removed and a single-chamber MFC is used:
costs and the internal ohmic resistance are decreased. In membrane-less MFCs, the need to keep electrodes close to each other
remains while avoiding electrode contact and short-circuiting. For this reason, other kinds of separator with higher porosity ranges
(porous media like earthware, ceramics, soil, and gravel) are used, to allow close electrode spacing and to improve power density
(Fan et al., 2007). One aspect to be considered in this case is related to biofilm growth on the separator, which might influence MFC
performances over time.

Configurations of MFC Reactors Toward Practical Applications


MFCs differ for reactor geometry (flat plate, tubular, spherical, etc.), the electrode separator (when present, a membrane, a salt
bridge, or a mechanical structure), and the cathode type (water cathode or air cathode, abiotic cathode, or biocathode). Many archi-
tectures and designs are nowadays under investigation, each one with advantages and drawbacks. Depending on the application, the
design should guarantee high power densities together with cheap materials and simple configurations. In view of connecting a high
number of MFC modules in parallel or series for large-scale energy harvesting or wastewater treatment, the unit module system
should be kept as simple as possible, with low-cost materials.
An example of applicable reactor setup, thought to meet low-cost material requirements, was presented by Ieropoulos et al.
(Pasternak et al., 2015). Their system comprises a cylindrical earthware membrane/separator that has become the structural material
of the cell. The terra-cotta core supports externally and internally carbon cloth anode and cathode, respectively. The terra-cotta works
as a porous septum, allowing ion transportation to the cathode and hindering water flow throughout its thickness.
Stacked systems electrically connect a number of single MFCs. In one single MFC, even neglecting the internal losses, the voltage
will never exceed a theoretical open-circuit voltage of 1.14 V (Aelterman et al., 2006). The use of series or parallel stacked MFCs is
therefore essential to increase voltages and currents. To achieve higher power output, the connection of multiple MFCs in stacked
configurations, either in series or in parallel, was recently studied in various tests (Feng et al., 2014; Gurung and Oh, 2012;
Pasupuleti et al., 2015a).
Pasupuleti et al. connected a number of MFC in a stack series, showing a gradual increment in voltage reaching a maximum of
2.12 V, with a power density of 3163 mW m 3 (Pasupuleti et al., 2015a). In a similar experiment, Rahimnejad et al. obtained
a maximum current and power generation in the stack of MFC, which were 6447 mA m 2 and 2003 mW m 2, respectively
(Rahimnejad et al., 2012).

Applications of MFC Beyond Power Generation


Wastewater Treatment
MFCs can be thought as self-powered wastewater treatment devices. While oxidizing organic matter dissolved in wastewaters from
different sources (domestic, industrial, agricultural, and agro-industrial), MFC reactors can recover up to ninefold the electrical
power required to treat the same wastewaters with traditional biological systems (Borjas et al., 2015). Even without recovering extra
electrical power, the favorable thermodynamics of the overall redox reaction system in MFCs drive the spontaneous oxidation of
organics. In parallel to COD abattoir, wastewater treatment requires nitrogen removal or recovery. Clauwaert et al. (2007) were
the first to study an MFC in which, simultaneously, organic substrate removal, power production, and complete denitrification
were achieved. Now, several research groups are studying nitrogen removal mechanisms, such as simultaneous nitrification and
denitrification (Virdis et al., 2010), cathodic denitrification (Puig et al., 2010), and ammonia stripping due to locally increased
pH at air cathodes (Kim et al., 2008).
The challenge of an efficient MFC-based wastewater treatment reactor is mainly linked to the capacity of bringing abundant
oxygen in contact with the cathodic surface while at the same time the organic-rich water in contact with the anodic surface.
When these two conditions are maximized, the efficiency of the treatment is proportional to the electrode surface area (both anodic
and cathodic) per reactor volume and inversely proportional to internal resistance of the system (Logan, 2010). Several stackable/
repeatable geometries were proposed in literature, with a high number of modular single cells connected in series or parallel
(Gurung and Oh, 2012; Pasupuleti et al., 2015a) or also working as stand-alone MFCs, if electricity recovery is not a priority (Borjas
et al., 2015).
Despite these applications having been studied for more than a decade, very few pilot prototypes at acceptably scaled-up dimen-
sions were developed worldwide. Here, we mention the tubular air cathode MFCs tested by the Advanced Water Management
Centre at the University of Queensland. The reactor consisted of 12 cylindrical modules, each 3 m high, with a total volume of
 1 m3, made of carbon fiber brush anodes and graphite fiber brush cathodes, with a GDL (Logan et al., 2006). The recent iMETland
project (http://imetland.eu/), University of Alcalá de Henares, Spain, proposed a microbial electrochemical wetland based on a bed
of activated carbon granules or other conductive materials. Here, the electrochemical activity of microbes is claimed to allow a 10-
fold improvement of biodegradation rate on domestic wastewaters, as compared with a wetland with a nonconductive bed.

Encyclopedia of Sustainable Technologies, First Edition, 2017, 245–259


Author's personal copy
254 Microbial Fuel Cells

To date, however, commercial applications of MFC wastewater treatment reactors are still missing, mainly due to the unit cost of
the materials of electrodes and separators. To achieve competitive water treatment capacities in a stacked MFC, electrode volumetric
surface area should be at least of the order of 100 m2 m 3 (Logan, 2010). In this case, to keep the cost of the m3 of reactor below
100 V (or $, roughly), the electrode (and membrane) materials should cost < 0.3 V m 2. This is true in the hypothesis that the
anodic surface area is equal to the cathodic and to the GDL (or membrane). To date, the most studied materials (carbon or graphite
fibers, cloths, felts, etc.), even without any metal catalyzer or treatment with gas-diffusing materials or microporous layers, don’t
cost < 10 V m 2 (Logan, 2010).

Microbial Solar Cells and Plant Microbial Fuel Cells


Microbial solar cells are devices that use photosynthetic organisms to harvest solar energy and synthesize organic matter (Fig. 5A).
This organic matter is oxidized in the anodic compartment by EAB (Elmekawy et al., 2014; Strik et al., 2008a; Strik et al., 2011).
Examples of these systems were proposed by Strik et al. suggesting a plant microbial fuel cell (P-MFC) by integrating the roots
of a living plant in the anode compartment of an MFC. The living plant produces and releases different types of organic compounds

Fig. 5 Schematic examples of MFC applicable setups (M ¼ microorganisms): (A) plant MFC (Helder et al., 2012), (B) sediment/benthic MFC (Li and
Yu, 2015), (C) saliva-fed miniaturized MFC (25 mL) (Mink et al., 2014), (D) floating MFC (Martinucci et al., 2015), and (E) bioremediation, nutrient,
and metal recovery in MFC (Wang and Ren, 2014).

Encyclopedia of Sustainable Technologies, First Edition, 2017, 245–259


Author's personal copy
Microbial Fuel Cells 255

into the soil (sugars, organic acids, enzymes, gases, etc.). Plant rhizodeposits are used by bacteria to generate electricity (Strik et al.,
2008b). Power densities in P-MFC depend on several aspects: solar radiation, plant photosynthetic efficiency, organic matter allo-
cation from plant to soil, and efficiency of MFC. A possible way to increase power output of the P-MFC is to reduce the internal
resistances. Strik’s P-MFC design was based on a tubular anode and the cathode was at the bottom of the cylinder. The flat-plate
P-MFC design proposed by Helder, with anode and cathode placed next to each other, separated by a membrane, minimized
internal resistances as compared with the tubular P-MFC design (Helder et al., 2012).
Similar types of MFCs are sometimes called microbial carbon capture cells (MCCs). Since the cathode chamber is sparged with
CO2, sequestration of this greenhouse gas can take place through biological conversion to organic matter by photosynthetic micro-
organisms such as microalgae and cyanobacteria. This organic matter can be transformed into products such as ethanol, bio-
fertilizer, hydrogen, and amino acids (Pandit et al., 2012). As an example, González Del Campo et al. (2013) developed an
MCC by introducing gas generated through bacterial respiration and metabolism at the anode into a cathode in which a photosyn-
thetic microorganism (Chlorella vulgaris) was growing.

Sediment/Benthic MFCs
Sediment microbial fuel cells (S-MFCs) represent a different concept of MFC, allowing in situ energy production from sustainable
sources (Fig. 5B). By burying the anode in an anaerobic sediment, with the cathode immersed in the overlying oxygen-rich water
(river, sea, lake, and groundwater), power harvest for prolonged periods can supply remote environmental sensors (Ewing et al.,
2014; Lowy and Tender, 2008). Water salinity facilitates conductivity between electrodes, and the organic matter needed by bacteria
to produce electricity is already present in the sediments. Reimers et al. noted that by S-MFC power generation, a level of
50 mW m 2 essentially indefinitely could be sustained (Reimers et al., 2001). The power produced by S-MFCs can be increased
by using the same principles developed in membrane-less MFC. Very large cathodes have been designed with the aim of relying
on dissolved oxygen, and anode performances have been modified to contain mediators, increasing from 1.5 to 1.7 times the power
than systems with plain graphite anodes. The main limiting factor in these systems is the electrode spacing, resulting in high internal
resistances (An et al., 2010).

Bioremediation: Electrodeposition, Electrosorption, and Electrocoagulation


Sustainable water management and soil bioremediation should aim not only at degrading organic pollutants but also at recovering/
reusing nutrients. MFCs can work as self-powered bioelectrochemical devices used for bioremediation of inorganic-contaminated
water in natural environments (Fig. 5E). The osmotic force created between anode and cathode of an MFC drags cations from the
anodic chamber to the cathode surface. Here, metal oxidized forms (Men þ) can accept electrons at cathode (by both abiotic and
biotic mechanisms) and be reduced to less toxic states (e.g., Cr6 þ to Cr3 þ) or elemental (Me0) or insoluble forms. A useful review
of the wide spectrum of metal deposition reactions that can be favored by MFCs was recently published by Wang and Ren (2014).
Other different mechanisms can lead to precipitation of insoluble complexes that include metals, organic substances, and nutri-
ents (P, Mg, Ca, K, etc.). Locally extreme alkaline pH conditions at cathode interface can induce insoluble salt precipitations (e.g.,
CaCO3, struvite, and various metal hydroxides) (Kelly and He, 2014; Li et al., 2012).

Microbial Desalination Cells


Microbial desalination cells (MDCs) (Fig. 4C) are another example of new, energy-sustainable methods for using organic matter in
wastewater as the power source for water desalination. Global water shortages have increased the need for desalination of waste-
water. The high energy requirement of conventional desalination techniques is demanding for the alternative treatment technology
that will require less energy for its efficient operation and recover useful energy to make this operation sustainable. MDCs use the
chemical energy of waste organic matter as driving force for desalination (Kim and Logan, 2013). The electric potential gradient
created by exoelectrogenic bacteria desalinates water by driving ion transport through a series of ion exchange membranes.

Environmental Sensors
Monitoring environmental parameters in natural contexts can help in understanding ecosystems and modeling their responses to
pollution stress. Complex networks of sensors, distributed in remote locations, need cheap power sources for operation and must
minimize maintenance operations. MFCs are nice candidates to comply this task, particularly in rivers, lakes, sea, sediments, etc.
where periodic battery replacement could be onerous. Sediment (as presented in the preceding text) or floating MFCs (Fig. 5D)
are applied to power monitoring devices for water quality in natural waterbodies (An et al., 2010; Huang et al., 2012). The anode
is generally embedded in organic-rich sediment, either natural or artificial (for the floating versions). Power densities are generally
limited (< 30 mW m 2) due to high internal resistances, the use of cheap electrode materials (plain carbon clothes or felts, activated
carbon granules, etc.), and the low concentrations of available organics at the anode. However, continuous energy storage devices
can collect the power of MFC stacks, to achieve the required voltage and current outputs.
Besides power harvesting, the MFC itself can be used as a sensor of different parameters. The current produced can be related in
real time to the organic matter concentration in water or sediments. Early studies correlated the electricity production to COD and

Encyclopedia of Sustainable Technologies, First Edition, 2017, 245–259


Author's personal copy
256 Microbial Fuel Cells

Biochemical oxygen demand (BOD) in batch-fed and flow MFCs (Kumlanghan et al., 2007; Min and Logan, 2004). Later, acetate
concentration was correlated with current production up to 2.3 mM in an MFC inoculated with a pure culture of Geobacter sulfurre-
ducens (Tront et al., 2008). The same relationship was observed in an MFC inoculated with Shewanella oneidensis and fed with lactate
in the range of 0–41 mM (Tront et al., 2008).

Miniaturizing/Microfluidic MFC
The objective of the microsized MFC (Fig. 5C) is to maximize power production for use as viable liquid-fueled miniature energy
harvesters. These devices are suitable for applications demanding miniaturized carbon-neutral and renewable energy sources to
power electronics such as implantable medical devices and microsensors (Mink et al., 2014). Key factors in designing such
systems are the single-chamber integrated cathode, which lowers the internal resistance, the high surface area of the anode
(with respect to the volume of the chamber) in order to enhance electron transfer mechanisms, and the possibility of using
very concentrated fuels, allowing higher power output (Ren et al., 2012). Miniaturized MFCs have been shown to embed
many attractive features such as faster mass transfer and reaction kinetics and short start-up time, which render this field
extremely attractive for future studies.

Conclusion

As a technology, MFCs have come a long way with rapid strides made in the last decade. These developments have taken in all
aspects including increase in current and power density (Logan et al., 2015; Patil et al., 2015), materials (electrodes and
membranes), design, scale, and range of biocatalysts used (Logan, 2009). In terms of substrates, over the years, various types
of wastewaters have been explored as substrates in MFCs (ElMekawy et al., 2015; Pandey et al., 2016; Pant et al., 2010).
Initially, the focus was mainly on substrates rich in mixture of carbon sources such as simple substrates like acetate and glucose
(Pant et al., 2010). However, in recent years, more complex and unconventional substrates like agricultural wastes, distilleries,
food, and dairy wastewaters have also been explored as substrates in MFCs (ElMekawy et al., 2015; Pandey et al., 2016). Simi-
larly, a wide range of materials have been explored as both anode and cathode in MFCs with increasing performance
(Pasupuleti et al., 2015a,b; Wei et al., 2011; Zhou et al., 2011). Same is the case for the membranes and/or separators (Leong
et al., 2013; Pasupuleti et al., 2015a). However, somehow, despite all these advances, the MFCs as a wastewater treatment tech-
nology or energy recovery devices are yet to come of age, and only in the last few years, efforts to commercialize them are being
explored, and first reports of real-world practical applications are coming out now (Pandey et al., 2016; cambrianinnovation,
n.d.). Besides energy recovery and wastewater treatment, offshoots of MFC technology are being explored for hydrogen produc-
tion in microbial electrolysis cells (Pasupuleti et al., 2015a), chemicals and fuel production in microbial electrosynthesis
systems (Bajracharya et al., 2015), water desalination in MDC, and solar energy recovery from living plants in P-MFCs
(ElMekawy et al., 2014).

References

Aelterman, P., Rabaey, K., Pham, H.T., Boon, N., Verstraete, W., 2006. Continuous electricity generation at high voltages and currents using stacked microbial fuel cells.
Environmental Science and Technology 40, 3388–3394. http://dx.doi.org/10.1021/es0525511.
An, J., Moon, H., Chang, I.S., 2010. Multiphase electrode microbial fuel cell system that simultaneously converts organics coexisting in water and sediment phases into electricity.
Environmental Science and Technology 44, 7145–7150. http://dx.doi.org/10.1021/es100498g.
Bajracharya, S., ter Heijne, A., Dominguez Benetton, X., Vanbroekhoven, K., Buisman, C.J.N., Strik, D., Pant, D., 2015. Carbon dioxide reduction by mixed and pure cultures in
microbial electrosynthesis using an assembly of graphite felt and stainless steel as a cathode. Bioresource Technology 195, 14–24. http://dx.doi.org/10.1016/
j.biortech.2015.05.081.
Bentley, R.W., 2002. Global oil & gas depletion: An overview. Energy Policy 30, 189–205. http://dx.doi.org/10.1016/S0301-4215(01)00144-6.
Borjas, Z., Ortiz, J., Aldaz, A., Feliu, J., Esteve-Núñez, A., 2015. Strategies for reducing the start-up operation of microbial electrochemical treatments of urban wastewater. Energies
8, 14064–14077. http://dx.doi.org/10.3390/en81212416.
Bond, D.R., Lovley, D.R., 2005. Evidence for involvement of an electron shuttle in electricity generation by Geothrix fermentans. Appl. Environ. Microbiol. 71, 2186–2189. http://
dx.doi.org/10.1128/AEM.71.4.2186-2189.2005.
http://cambrianinnovation.com/ecovolt-now-commercially-available/ (Accessed on 10.04.16).
Clauwaert, P., Rabaey, K., Aelterman, P., De Schamphelaire, L., Pham, T.H., Boeckx, P., et al., 2007. Biological denitrification in microbial fuel cells. Environmental Science and
Technology 41, 3354–3360. http://dx.doi.org/10.1021/es062580r.
Cristiani, P., Carvalho, M.L., Guerrini, E., Daghio, M., Santoro, C., Li, B., 2013. Cathodic and anodic biofilms in single chamber microbial fuel cells. Bioelectrochemistry 92, 6–13.
http://dx.doi.org/10.1016/j.bioelechem.2013.01.005.
Daghio, M., Gandolfi, I., Bestetti, G., Franzetti, A., Guerrini, E., Cristiani, P., 2015. Anodic and cathodic microbial communities in single chamber microbial fuel cells. New
Biotechnology 32, 79–84. http://dx.doi.org/10.1016/j.nbt.2014.09.005.
Dantas, P.V., Peres, S., Campos-Takaki, G.M., La Rotta, C.E., 2013. Utilization of raw glycerol for pyocyanin production from pseudomonas aeruginosa in half-microbial fuel cells:
Evaluation of two electrochemical approaches. Journal of the Electrochemical Society 160, G142–G148.
Du, Z., Li, H., Gu, T., 2007. A state of the art review on microbial fuel cells: A promising technology for wastewater treatment and bioenergy. Biotechnology Advances 25, 464–482.
http://dx.doi.org/10.1016/j.biotechadv.2007.05.004.
Elmekawy, A., Hegab, H.M., Vanbroekhoven, K., Pant, D., 2014. Techno-productive potential of photosynthetic microbial fuel cells through different configurations. Renewable and
Sustainable Energy Reviews 39, 617–627. http://dx.doi.org/10.1016/j.rser.2014.07.116.

Encyclopedia of Sustainable Technologies, First Edition, 2017, 245–259


Author's personal copy
Microbial Fuel Cells 257

ElMekawy, A., Sandipam, S., Bajracharya, S., Hegab, H.M., Nigam, P.S., Singh, A., Venkata Mohan, S., Pant, D., 2015. Food and agricultural wastes as substrates for bio-
electrochemical system (BES): The synchronized recovery of sustainable energy and waste treatment. Food Research International 73C, 213–225. http://dx.doi.org/10.1016/
j.foodres.2014.11.045.
Ewing, T., Ha, P.T., Babauta, J.T., Tang, N.T., Heo, D., Beyenal, H., 2014. Scale-up of sediment microbial fuel cells. Journal of Power Sources 272, 311–319. http://dx.doi.org/
10.1016/j.jpowsour.2014.08.070.
Fan, Y., Hu, H., Liu, H., 2007. Enhanced Coulombic efficiency and power density of air-cathode microbial fuel cells with an improved cell configuration. Journal of Power Sources
171, 348–354. http://dx.doi.org/10.1016/j.jpowsour.2007.06.220.
Feng, Y., He, W., Liu, J., Wang, X., Qu, Y., Ren, N., 2014. A horizontal plug flow and stackable pilot microbial fuel cell for municipal wastewater treatment. Bioresource Technology
156, 132–138. http://dx.doi.org/10.1016/j.biortech.2013.12.104.
Fricke, K., Harnisch, F., Schröder, U., 2008. On the use of cyclic voltammetry for the study of anodic electron transfer in microbial fuel cells. Energy and Environmental Science 1,
144–147. http://dx.doi.org/10.1039/B802363H.
Gajda, I., Greenman, J., Melhuish, C., Ieropoulos, I., 2013. Photosynthetic cathodes for microbial fuel cells. International Journal of Hydrogen Energy 38, 11559–11564. http://
dx.doi.org/10.1016/j.ijhydene.2013.02.111.
Gajda, I., Greenman, J., Melhuish, C., Santoro, C., Li, B., Cristiani, P., et al., 2014. Water formation at the cathode and sodium recovery using microbial fuel cells (MFCs).
Sustainable Energy Technologies and Assessments 7, 187–194. http://dx.doi.org/10.1016/j.seta.2014.05.001.
González Del Campo, A., Cañizares, P., Rodrigo, M.A., Fernández, F.J., Lobato, J., 2013. Microbial fuel cell with an algae-assisted cathode: A preliminary assessment. Journal of
Power Sources 242, 638–645. http://dx.doi.org/10.1016/j.jpowsour.2013.05.110.
Gorby, Y.A., Yanina, S., McLean, J.S., Rosso, K.M., Moyles, D., Dohnalkova, A., et al., 2006. Electrically conductive bacterial nanowires produced by Shewanella oneidensis strain
MR-1 and other microorganisms. Proceedings of the National Academy of Science 103, 11358–11363. http://dx.doi.org/10.1073/pnas.0604517103.
Guerrini, E., Grattieri, M., Faggianelli, A., Cristiani, P., Trasatti, S., 2015. PTFE effect on the electrocatalysis of the oxygen reduction reaction in membraneless microbial fuel cells.
Bioelectrochemistry 106, 240–247. http://dx.doi.org/10.1016/j.bioelechem.2015.05.008.
Gurung, A., Oh, S.-E., 2012. The performance of serially and parallelly connected microbial fuel cells. Energy Sources, Part A Recover Utilization, and Environmental Effects 34,
1591–1598. http://dx.doi.org/10.1080/15567036.2011.629277.
Harnisch, F., Freguia, S., 2012. A basic tutorial on cyclic voltammetry for the investigation of electroactive microbial biofilms. Chemistry, An Asian Journal 7, 466–475. http://
dx.doi.org/10.1002/asia.201100740.
Harris, H.W., El-Naggar, M.Y., Bretschger, O., Ward, M.J., Romine, M.F., Obraztsova, A.Y., et al., 2010. Electrokinesis is a microbial behavior that requires extracellular electron
transport. Proceedings of the National Academy of Science of the United States of America 107, 326–331. http://dx.doi.org/10.1073/pnas.0907468107.
He, Z., Mansfeld, F., 2009. Exploring the use of electrochemical impedance spectroscopy (EIS) in microbial fuel cell studies. Energy and Environmental Science 2, 215–219. http://
dx.doi.org/10.1039/B814914C.
Helder, M., Strik, D.P., Hamelers, H.V., Buisman, C.J., 2012. The flat-plate plant-microbial fuel cell: The effect of a new design on internal resistances. Biotechnology for Biofuels 5,
70. http://dx.doi.org/10.1186/1754-6834-5-70.
Holmes, D.E., Bond, D.R., O’Neil, R.A., Reimers, C.E., Tender, L.R., Lovley, D.R., 2004. Microbial communities associated with electrodes harvesting electricity from a variety of
aquatic sediments. Microbial Ecology 48, 178–190. http://dx.doi.org/10.1007/s00248-003-0004-4.
Hosseini, M.G., Ahadzadeh, I., 2013. Application and comparison of current interruption and electrochemical impedance spectroscopy methods to study a microbial fuel cell.
Instrumentation Science and Technology 41, 72–81. http://dx.doi.org/10.1080/10739149.2012.717331.
Huang Y, He Z, and Mansfeld F (2012) Floating microbial fuel cells with cation exchange membrane. U.S. Pat. Appl. Publ. US 20120082868 A1 20120405.
Kelly, P.T., He, Z., 2014. Nutrients removal and recovery in bioelectrochemical systems: A review. Bioresource Technology 153, 351–360. http://dx.doi.org/10.1016/
j.biortech.2013.12.046.
Kim, Y., Logan, B.E., 2013. Microbial desalination cells for energy production and desalination. Desalination 308, 122–130. http://dx.doi.org/10.1016/j.desal.2012.07.022.
Kim, H.J., Park, H.S., Hyun, M.S., Chang, I.S., Kim, M., Kim, B.H., 2002. A mediator-less microbial fuel cell using a metal reducing bacterium, Shewanella putrefaciens. Enzyme and
Microbial Technology 30, 145–152. http://dx.doi.org/10.1016/S0141-0229(01)00478-1.
Kim, J.R., Cheng, S., Oh, S.-E., Logan, B.E., 2007. Power generation using different cation, anion, and ultrafiltration membranes in microbial fuel cells. Environmental Science and
Technology 41, 1004–1009. http://dx.doi.org/10.1021/es062202m.
Kim, J.R., Zuo, Y., Regan, J.M., Logan, B.E., 2008. Analysis of ammonia loss mechanisms in microbial fuel cells treating animal wastewater. Biotechnology and Bioengineering 99,
1120–1127. http://dx.doi.org/10.1002/bit.21687.
Kumar, G.G., Sarathi, V.G.S., Nahm, K.S., 2013. Recent advances and challenges in the anode architecture and their modifications for the applications of microbial fuel cells.
Biosensors and Bioelectronics 43, 461–475. http://dx.doi.org/10.1016/j.bios.2012.12.048.
Kumlanghan, A., Liu, J., Thavarungkul, P., Kanatharana, P., Mattiasson, B., 2007. Microbial fuel cell-based biosensor for fast analysis of biodegradable organic matter. Biosensors
and Bioelectronics 22, 2939–2944. http://dx.doi.org/10.1016/j.bios.2006.12.014.
Leong, J.X., Daud, W.R.W., Ghasemi, M., Liew, K.B., Ismail, M., 2013. Ion exchange membranes as separators in microbial fuel cells for bioenergy conversion: A comprehensive
review. Renewable and Sustainable Energy Reviews 28, 575–587. http://dx.doi.org/10.1016/j.rser.2013.08.052.
Li, W.W., Yu, H.Q., 2015. Stimulating sediment bioremediation with benthic microbial fuel cells. Biotechnology Advances 33, 1–12. http://dx.doi.org/10.1016/
j.biotechadv.2014.12.011.
Li, Z., Ren, X., Zuo, J., Liu, Y., Duan, E., Yang, J., et al., 2012. Struvite precipitation for ammonia nitrogen removal in 7-aminocephalosporanic acid wastewater. Molecules 17,
2126–2139. http://dx.doi.org/10.3390/molecules17022126.
Liu, H., Logan, B.E., 2004. Electricity generation using an air-cathode single chamber microbial fuel cell in the presence and absence of a proton exchange membrane. Envi-
ronmental Science and Technology 38, 4040–4046. http://dx.doi.org/10.1021/es0499344.
Liu, Z.D., Du, Z.W., Lian, J., Zhu, X.Y., Li, S.H., Li, H.R., 2007. Improving energy accumulation of microbial fuel cells by metabolism regulation using Rhodoferax ferrireducens as
biocatalyst. Letters in Applied Microbiology 44, 393–398. http://dx.doi.org/10.1111/j.1472-765X.2006.02088.x.
Logan, B.E., 2008. Microbial Fuel Cells. John Wiley & Sons.
Logan, B.E., 2009. Exoelectrogenic bacteria that power microbial fuel cells. Nature Reviews Microbiology 7 (5), 375–381. http://dx.doi.org/10.1038/nrmicro2113.
Logan, B.E., 2010. Scaling up microbial fuel cells and other bioelectrochemical systems. Applied Microbiology and Biotechnology 85, 1665–1671. http://dx.doi.org/10.1007/
s00253-009-2378-9.
Logan, B.E., Regan, J.M., 2006. Electricity-producing bacterial communities in microbial fuel cells. Trends in Microbiology 14, 512–518. http://dx.doi.org/10.1016/
j.tim.2006.10.003.
Logan, B.E., Hamelers, B., Rozendal, R., Schröder, U., Keller, J., Freguia, S., et al., 2006. Microbial fuel cells: Methodology and technology. Environmental Science and Technology
40, 5181–5192. http://dx.doi.org/10.1021/es0605016.
Logan, B., Cheng, S., Watson, V., Estadt, G., 2007. Graphite fiber brush anodes for increased power production in air-cathode microbial fuel cells. Environmental Science and
Technology 41, 3341–3346. http://dx.doi.org/10.1021/es062644y.
Logan, B.E., Wallack, M.J., Kim, K.Y., He, W., Feng, Y., Saikaly, P.E., 2015. Assessment of microbial fuel cell configurations and power densities. Environmental Science and
Technology Letters 2 (8), 206–214. http://dx.doi.org/10.1021/acs.estlett.5b00180.
Lovley, D.R., Holmes, D.E., Nevin, K.P., 2004. Dissimilatory Fe(III) and Mn(IV) reduction. Advances in Microbial Physiology 49, 219–286. http://dx.doi.org/10.1016/S0065-2911(04)
49005-5.

Encyclopedia of Sustainable Technologies, First Edition, 2017, 245–259


Author's personal copy
258 Microbial Fuel Cells

Lowy, D.A., Tender, L.M., 2008. Harvesting energy from the marine sediment–water interface. Journal of Power Sources 185, 70–75. http://dx.doi.org/10.1016/
j.jpowsour.2008.06.079.
Manohar, A.K., Mansfeld, F., 2009. The internal resistance of a microbial fuel cell and its dependence on cell design and operating conditions. Electrochimica Acta 54, 1664–1670.
http://dx.doi.org/10.1016/j.electacta.2008.06.047.
Martinucci, E., Pizza, F., Perrino, D., Colombo, A., Trasatti, S.P.M., Lazzarini Barnabei, A., et al., 2015. Energy balance and microbial fuel cells experimentation at wastewater
treatment plant Milano-Nosedo. International Journal of Hydrogen Energy 40, 14683–14689. http://dx.doi.org/10.1016/j.ijhydene.2015.08.100.
Marzorati, S., Vasconcelos, J.M., Ding, J., Longhi, M., Colavita, P.E., 2015. Template-free ultraspray pyrolysis synthesis of N/Fe-doped carbon microspheres for oxygen reduction
electrocatalysis. Journal of Materials Chemistry A 3, 18920–18927. http://dx.doi.org/10.1039/C5TA02570B.
Meitl, L.A., Eggleston, C.M., Colberg, P.J.S., Khare, N., Reardon, C.L., Shi, L., 2009. Electrochemical interaction of Shewanella oneidensis MR-1 and its outer membrane
cytochromes OmcA and MtrC with hematite electrodes. Geochimica et Cosmochimica Acta 73, 5292–5307. http://dx.doi.org/10.1016/j.gca.2009.06.021.
Min, B., Logan, B.E., 2004. Continuous electricity generation from domestic wastewater and organic substrates in a flat plate microbial fuel cell. Environmental Science and
Technology 38, 5809–5814. http://dx.doi.org/10.1021/es0491026.
Mink, J.E., Qaisi, R.M., Logan, B.E., Hussain, M.M., 2014. Energy harvesting from organic liquids in micro-sized microbial fuel cells. NPG Asia Materials 6, e89. http://dx.doi.org/
10.1038/am.2014.1.
Pandey, P., Shinde, V.N., Deopurkar, R.L., Kale, S.P., Patil, S.A., Pant, D., 2016. Recent advances in the use of different substrates in microbial fuel cells towards wastewater
treatment and simultaneous energy recovery. Applied Energy 168, 706–723. http://dx.doi.org/10.1016/j.apenergy.2016.01.056.
Pandit, S., Nayak, B.K., Das, D., 2012. Microbial carbon capture cell using cyanobacteria for simultaneous power generation, carbon dioxide sequestration and wastewater
treatment. Bioresource Technology 107, 97–102. http://dx.doi.org/10.1016/j.biortech.2011.12.067.
Pant, D., Van Bogaert, G., Diels, L., Vanbroekhoven, K., 2010. A review of the substrates used in microbial fuel cells (MFCs) for sustainable energy production. Bioresource
Technology 101, 1533–1543. http://dx.doi.org/10.1016/j.biortech.2009.10.017.
Park, H.S., Kim, B.H., Kim, H.S., Kim, H.J., Kim, G.T., Kim, M., et al., 2001. A novel electrochemically active and Fe(III)-reducing bacterium phylogenetically related to clostridium
butyricum isolated from a microbial fuel cell. Anaerobe 7, 297–306. http://dx.doi.org/10.1006/anae.2001.0399.
Pasternak, G., Greenman, J., Ieropoulos, I., 2015. Comprehensive study on ceramic membranes for low-cost microbial fuel cells. ChemSusChem 9, 88–96. http://dx.doi.org/
10.1002/cssc.201501320.
Pasupuleti, S.B., Srikanth, S., Venkata Mohan, S., Pant, D., 2015a. Continuous mode operation of microbial fuel cell (MFC) stack with dual gas diffusion cathode design for the
treatment of dark fermentation effluent. International Journal of Hydrogen Energy 40 (36), 12424–12435. http://dx.doi.org/10.1016/j.ijhydene.2015.07.049.
Pasupuleti, S.B., Srikanth, S., Venkata Mohan, S., Pant, D., 2015b. Development of exoelectrogenic bioanode and study on feasibility of hydrogen production using abiotic VITO-
CoRE™ and VITO-CASE™ electrodes in a single chamber microbial electrolysis cell (MEC) at low current densities. Bioresour Technology 195, 131–138. http://dx.doi.org/
10.1016/j.biortech.2015.06.145.
Patil, S.A., Gildemyn, S., Pant, D., Zengler, K., Logan, B.E., Rabaey, K., 2015. A logical data representation framework for electricity-driven bioproduction processes. Biotechnology
Advances 33 (6), 736–744. http://dx.doi.org/10.1016/j.biotechadv.2015.03.002.
Pham, C.A., Jung, S.J., Phung, N.T., Lee, J., Chang, I.S., Kim, B.H., et al., 2003. A novel electrochemically active and Fe(III)-reducing bacterium phylogenetically related to
Aeromonas hydrophila, isolated from a microbial fuel cell. FEMS Microbiology Letters 223, 129–134. http://dx.doi.org/10.1016/S0378-1097(03)00354-9.
Prasad, D., Arun, S., Murugesan, M., Padmanaban, S., Satyanarayanan, R.S., Berchmans, S., Yegnaraman, V., 2007. Direct electron transfer with yeast cells and construction of
a mediatorless microbial fuel cell. Biosensors and Bioelectronics 22, 2604–2610. http://dx.doi.org/10.1016/j.bios.2006.10.028.
Puig, S., Serra, M., Coma, M., Cabré, M., Balaguer, M.D., Colprim, J., 2010. Effect of pH on nutrient dynamics and electricity production using microbial fuel cells. Bioresource
Technology 101, 9594–9599. http://dx.doi.org/10.1016/j.biortech.2010.07.082.
Rabaey, K., Angenent, L.T., Schröder, U., Keller, J., 2009. Bioelectrochemical systems: From extracellular electron transfer to biotechnological application. Water Intelligence Online
8. http://dx.doi.org/10.2166/9781780401621, 9781780401621.
Rahimnejad, M., Ghoreyshi, A.A., Najafpour, G.D., Younesi, H., Shakeri, M., 2012. A novel microbial fuel cell stack for continuous production of clean energy. International Journal of
Hydrogen Energy 37, 5992–6000. http://dx.doi.org/10.1016/j.ijhydene.2011.12.154.
Reimers, C.E., Tender, L.M., Fertig, S., Wang, W., 2001. Harvesting energy from the marine sediment  water interface. Environmental Science and Technology 35, 192–195.
http://dx.doi.org/10.1021/es001223s.
Ren, H., Lee, H.-S., Chae, J., 2012. Miniaturizing microbial fuel cells for potential portable power sources: Promises and challenges. Microfluidics and Nanofluidics 13, 353–381.
http://dx.doi.org/10.1007/s10404-012-0986-7.
Santoro, C., Lei, Y., Li, B., Cristiani, P., 2012. Power generation from wastewater using single chamber microbial fuel cells (MFCs) with platinum-free cathodes and pre-colonized
anodes. Biochemical Engineering Journal 62, 8–16. http://dx.doi.org/10.1016/j.bej.2011.12.006.
Santoro, C., Ieropoulos, I., Greenman, J., Cristiani, P., Vadas, T., Mackay, A., et al., 2013. Power generation and contaminant removal in single chamber microbial fuel cells
(SCMFCs) treating human urine. International Journal of Hydrogen Energy 38, 11543–11551. http://dx.doi.org/10.1016/j.ijhydene.2013.02.070.
Santoro, C., Artyushkova, K., Babanova, S., Atanassov, P., Ieropoulos, I., Grattieri, M., et al., 2014. Parameters characterization and optimization of activated carbon (AC) cathodes
for microbial fuel cell application. Bioresource Technology 163, 54–63. http://dx.doi.org/10.1016/j.biortech.2014.03.091.
Squadrito, G., Cristiani, P., 2016. Compendium of hydrogen energy. Elsevier. http://dx.doi.org/10.1016/B978-1-78242-363-8.00006-2.
Strik, D.P.B.T.B., Terlouw, H., Hamelers, H.V.M., Buisman, C.J.N., 2008a. Renewable sustainable biocatalyzed electricity production in a photosynthetic algal microbial fuel cell
(PAMFC). Applied Microbiology and Biotechnology 81, 659–668. http://dx.doi.org/10.1007/s00253-008-1679-8.
Strik, D.P.B.T.B., Hamelers (Bert), H.V.M., Snel, J.F.H., Buisman, C.J.N., 2008b. Green electricity production with living plants and bacteria in a fuel cell. International Journal of
Energy Research 32, 870–876. http://dx.doi.org/10.1002/er.1397.
Strik, D.P.B.T.B., Timmers, R.A., Helder, M., Steinbusch, K.J.J., Hamelers, H.V.M., Buisman, C.J.N., 2011. Microbial solar cells: Applying photosynthetic and electrochemically
active organisms. Trends in Biotechnology 29, 41–49. http://dx.doi.org/10.1016/j.tibtech.2010.10.001.
Tront, J.M., Fortner, J.D., Plötze, M., Hughes, J.B., Puzrin, A.M., 2008. Microbial fuel cell biosensor for in situ assessment of microbial activity. Biosensors and Bioelectronics 24,
586–590. http://dx.doi.org/10.1016/j.bios.2008.06.006.
Verstraete, W., Rabaey, K., 2006. Critical review microbial fuel cells: Methodology and technology. Environmental Science and Technology 40, 5181–5192. http://dx.doi.org/
10.1021/es0605016.
Virdis, B., Freguia, S., Rozendal, R.A., Rabaey, K., Yuan, Z., Keller, J., 2011. Microbial fuel cells. In: Wilderer, P. (Ed.), Treatise on Water Science 4, 641–665. http://dx.doi.org/
10.1016/B978-0-444-53199-5.00098-1.
Virdis, B., Rabaey, K., Rozendal, R.A., Yuan, Z., Keller, J., 2010. Simultaneous nitrification, denitrification and carbon removal in microbial fuel cells. Water Research 44, 2970–
2980. http://dx.doi.org/10.1016/j.watres.2010.02.022.
Wang, H., Ren, Z.J., 2014. Bioelectrochemical metal recovery from wastewater: A review. Water Research 66, 219–232. http://dx.doi.org/10.1016/j.watres.2014.08.013.
Wei, J., Liang, P., Huang, X., 2011. Recent progress in electrodes for microbial fuel cells. Bioresource Technology 102 (20), 9335–9344. http://dx.doi.org/10.1016/
j.biortech.2011.07.019.
Zhao, F., Harnisch, F., Schröder, U., Scholz, F., Bogdanoff, P., Herrmann, I., 2006. Challenges and constraints of using oxygen cathodes in microbial fuel cells. Environmental
Science and Technology 40, 5193–5199. http://dx.doi.org/10.1021/es060332p.
Zhao, F., Slade, R.C.T., Varcoe, J.R., 2009. Techniques for the study and development of microbial fuel cells: An electrochemical perspective. Chemical Society Reviews 38, 1926–
1939. http://dx.doi.org/10.1039/b819866g.
Zhou, M., Chi, M., Luo, J., He, H., Jin, T., 2011. An overview of electrode materials in microbial fuel cells. Journal of Power Sources 196, 4427–4435. http://dx.doi.org/10.1016/
j.jpowsour.2011.01.012.

Encyclopedia of Sustainable Technologies, First Edition, 2017, 245–259


Author's personal copy
Microbial Fuel Cells 259

Relevant Websites

http://www.engr.psu.edu/ce/enve/logan/bioenergy/research_mfc.htm – The website from the group of Prof. Bruce Logan. It provides several practical tips on how to build an MFC.
http://www.geobacter.org/Microbial_Fuel_Cells – The website from the group of Prof. Derek Lovley. It provides information about Geobacter-powered MFCs and links to MFCs in
popular media.
http://www.is-met.org/ – This is the website of the flagship organization on MFC technology, the International Society for Microbial Electrochemistry and Technology.
https://www.linkedin.com/groups/1903929 – A group on LinkedIn-connecting MFC researchers.
https://www.mudwatt.com/ – A company selling MFCs for educational purposes.
http://www.plantpower.eu/ – The company working on plant microbial fuel cells (P-MFCs).

Encyclopedia of Sustainable Technologies, First Edition, 2017, 245–259


View publication stats

Vous aimerez peut-être aussi