Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Sensors and Actuators A 177 (2012) 3036

Contents lists available at SciVerse ScienceDirect

Sensors and Actuators A: Physical


journal homepage: www.elsevier.com/locate/sna

Structural optimization of contact electrodes in microbial fuel cells for current


density enhancements
Shogo Inoue a,f, , Erika A. Parra b,f , Adrienne Higa f , Yingqi Jiang f , Pengbo Wang c,f , Cullen R. Buie d,f ,
John D. Coates e , Liwei Lin f
a

Microdevice Research & Development Department, TAIYO YUDEN CO., LTD., Japan
Department of Organismic & Evolutionary Biology, Harvard University, USA
c
College of Engineering, China Agricultural University, China
d
Department of Mechanical Engineering, Massachusetts Institute of Technology, USA
e
Department of Plant and Microbial Biology, University of California at Berkeley, USA
f
Berkeley Sensor and Actuator Center, Department of Mechanical Engineering, University of California at Berkeley, USA
b

a r t i c l e

i n f o

Article history:
Available online 22 September 2011
Keywords:
Microbial fuel cell
Current density
Power density
Carbon nanotube
Geobacter sulfurreducens

a b s t r a c t
More than 200% current density enhancement in miniaturized microbial fuel cells (MFCs) has been successfully demonstrated by optimizing the contact electrode structure using micro and nano features. Two
fundamental issues are addressed in this work: (1) a methodology to enhance current/power density of
MFCs by changing micro and nano structural congurations of contact electrodes and (2) a study on the
effectiveness of charge transfer between living cells with organic nanowire-pili and micro/nano interfacial electrodes. This paper details the fabrication and characterization processes of miniaturized MFCs
with experimental results, and discusses the prospective power density of MFCs using micro/nano processes. Moreover, a hypothesis for the direct electron transport mechanism from living cells to electrodes
is experimentally corroborated. As such, this work represents a step toward higher energy conversion
efciency as well as practical applications of MFCs.
2011 Elsevier B.V. All rights reserved.

1. Introduction
Compact and highly efcient power sources are integral components for the completion of autonomous sensors and microsystems.
A microbial fuel cell (MFC) is a promising candidate for electrical energy harvesting from ubiquitous organic wastes [1]. The
breakdown of organic substances to retrieve energy is a naturally
occurring process in nature and the possibility to extract electrical charges in the form of MFCs holds great potential in practical
applications such as implantable medical sensors and long-term
monitoring systems in remote locations.
Recently, L-scale MFCs fabricated with MEMS technologies
were demonstrated using Geobacteraceae [2] and Shewanellaceae
[3] as the organic catalysts that break down reduced carbon
molecules during the metabolic process. Previously, our group has
also reported miniaturized MFCs which use bakers yeasts [4,5]
and Geobacter sulfurreducens [6] as the bio-catalysts. A photosynthetic electrochemical cell has also been presented that harnesses
the subcellular thylakoid photosystems isolated from spinach cells

Corresponding author at: 64, Nishiwaki, Ohkubo-cho, Akashi, Hyogo 674-8555,


Japan. Tel.: +81 78 937 1270; fax: +81 78 937 1271.
E-mail addresses: inoue-shogo@jty.yuden.co.jp, VZA04031@nifty.com (S. Inoue).
0924-4247/$ see front matter 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.sna.2011.09.023

to convert light energy into electricity [7]. In general, electrons


were extracted during these biological processes to provide electricity. Electron mediators have been used to extract electrons and
protons from cells or sub-cellular components (to be referred to
as sub-cell hereafter) in engineering fuel cell systems. Electrons
are transported externally via a conducting electrode (anode), and
protons are transported through a proton-exchange-membrane
(PEM) and reduce the oxidant in the cathode compartment. The
efciencies of these MFCs are low due to many loss mechanisms,
particularly during the electron transfer processes. Therefore, various efforts have been conducted to reduce the energy losses in
MFCs. One direction to improve the efciencies is to modify the
electrode surfaces with nanostructures such as carbon nanotubes
(CNTs). It has been demonstrated that anode electrodes with CNTdoped polyaniline improved the MFC performance [8] and Pt loaded
CNTs electrodes increased the power density as high as 6-fold as
compared to graphite electrodes [9]. Carbon nanostructures on
stainless steel mesh anodes have been shown to increase the power
density of MFC by a factor of 60 [10] and CNT powder added to the
anode chamber have reduced the anodic resistance [11]. On the
other hand, power densities in MFCs are also dependent on the
absolute and relative size of their components. The anode to cathode surface area ratio can affect power output, where tripling the
surface area of the cathode increased power density by 22% [12].

S. Inoue et al. / Sensors and Actuators A 177 (2012) 3036

e-

e-

Anode
CO2

Cathode

CO2
Metabolism

H2O
e-

Bacterium

Fuel
(e.g. Acetate)

Fuel

31

e-

O2
H+

H2O

H+

Air
(O2)

Proton exchange membrane (PEM)


Fig. 1. Schematic diagram showing the operating principle of the microbial fuel cell.
Acetate (or other fuel) is fed into the anode chamber and bacteria are used to break
down this fuel during metabolism to produce electrical charges.

Furthermore, it has been also demonstrated that the surface area


of both electrodes and the PEM in MFCs could affect power density
[13]. Our previous work [14,15] also studied and characterized the
inuences of sub-cell-to-electrode and cell-to-electrode contact
structures, respectively, to increase living sub-cell-to-electrode and
cell-to-electrode contact areas. These are possible approaches to
reduce the loss of electron transfer by using micro- and nanostructured electrodes and to improve the power density of MFCs.
This paper details the structural optimization of contact electrodes
with experimental characterization to demonstrate current density
enhancements with quantitative data to provide design guidelines
for future MFCs.

Fig. 3. Fabrication processes of microelectrodes with holes and channels. Two kinds
of silicon etching processes were used to make perpendicular and tapered sidewalls.

shown in Fig. 2: (A) at surface as a reference, (B) surface with


micro-holes perpendicularly etched into the substrate, (C) surface
with tapered micro-channels, and (D) vertically aligned carbon
nanotube (CNT) arrays. These contact structures were used as the
basic tools to investigate the current density enhancements and
effectiveness of charge transfer between living cells with organic
nanowire-pili and micro/nano interfacial electrodes, and to optimize the structural congurations of contact electrodes for current
density enhancements.

2. Operation principle

3. Fabrication

Fig. 1 shows the schematic diagram of the operation principle


of a MFC. In the anode compartment, organic fuel such as acetate
is fed into the system and microorganisms act as bio-catalysts to
break down the organic matter. Electrons generated during the
metabolic process are transferred to the anode electrode via electron mediators [4,5]. Fig. 2 illustrates the basic operation principle
of the MFC by using G. sulfurreducens as the catalyst. G. sulfurreducens was chosen in this work because their pili have shown to
function like organic nanowires to directly transfer electrons to
the electrodes [1618]. For our experiments, four different kinds
of modied cell-to-electrode contact structures were fabricated as

3.1. Micro electrodes

Fig. 2. The basic operating principle of an MFC using Geobacter sulfurreducens as


the bio-catalyst. The inset shows four different kinds of modied cell-to-electrode
contact structures: (A) at, (B) hole, (C) channel, and (D) vertically aligned CNTs
(not drawn to scale). Cell-to-electrode contact structures affect the current/power
density of the MFCs.

Micro electrodes with (A) at surfaces, (B) surfaces with


5 m 5 m holes and (C) surfaces with 5-m-wide, 3.6-mdeep channels were fabricated. The fabrication processes of micro
electrodes with holes and channels are shown in Fig. 3. After patterning photoresist, silicon substrates were etched. The hole-shape
patterns were made using DRIE processing, creating 8-m-deep
holes with sidewalls perpendicular to the top surface. Since the
metal deposition process was done by the evaporation of Au/Cr
(400/200 nm), the sidewalls were not completely covered, making only the 2-m-wide gaps between the hole patterns active for
collecting the charge during MFC operation. The channel-shape patterns were fabricated using an isotropic plasma etching process.
The plasma etching process was conducted for 5 min using SF6 and
O2 with a ow rate of 9:1 applying the RF power of 100 W. The
nal angle of the sidewalls was 55 , this enabled the entire surface,
including the channel sidewalls, to be covered with metal to collect
charges. Fig. 4a and b are SEM images of fabricated electrode structures with (a) the hole-pattern and (b) channel-pattern surfaces,
respectively. The right-hand images in Fig. 4 show a close-up of
the electrode surface. G. sulfurreducens have a length of about 1 m
and a diameter of approximately 0.3 m. Therefore, the dimensions
of fabricated holes and channels are large enough to accommodate bacteria. The active metalized areas for the hole and channel
patterns are calculated. For the hole-shape patterns, only the gap
area between the hole-patterns was measured using the SEM image
shown on the right of Fig. 4a. For the channel-shape patterns, the
surface area was measured by considering the cross-section of
channels as triangles with 2 m gaps as shown by red lines on the
right of Fig. 4b. Consequently, the active metalized areas for hole
and channel patterns are calculated as 49% and 154%, respectively,
of the at reference surface area.

32

S. Inoue et al. / Sensors and Actuators A 177 (2012) 3036

Fig. 4. SEM images of fabricated micro electrodes: (A) electrodes with hole-shape patterns, (B) electrode with channel patterns. In (A) and (B), the gure on the right is a
higher magnication image of the electrode surface on the left.

3.2. CNT electrode


CNT electrodes have been constructed on top of silicon substrates with vertically aligned, 10-m-wide, 36-m-tall CNT
forests. The fabrication process is shown in Fig. 5 [19]. After evaporating an 80-nm-thick molybdenum layer on a thermally oxidized
silicon wafer, a lithography process was conducted as shown in
Fig. 5a. Afterwards, Al and Fe layers of 10 nm and 5 nm, respectively,
were deposited in Fig. 5b and unwanted metals were removed by
lift-off as shown in Fig. 5c. In the CNT growth process, Fe acts as the
catalyst for CNT growth. Fig. 5d shows the thermal chemical vapor
deposition (CVD) process to grow the CNTs. The furnace was rst
purged with hydrogen and then heated to 720 C in a hydrogen
environment. Subsequently, the mixture of ethylene and hydrogen with a proportion of about 1:2.5 was owed for 10 min. CNT
patterns were made within 3 mm 3 mm area on at molybdenum as illustrated in Fig. 5e. Fig. 6 shows SEM images of fabricated
CNT electrodes in the top view (Fig. 6a) and cross-sectional view

(Fig. 6b). The height of the CNT electrode was measured as 36 m.


The 10-m-wide gaps are large enough to accommodate bacteria
between CNT walls. The external surface area of the electrode with
CNT patterns was calculated by assuming CNT patterns as walls 10m-wide, 36-m-tall with 10 m gaps. The external surface area
of the electrode with CNT walls was 460% with respect to the at
surface area. However, because there are at molybdenum areas
surrounding the CNT patterns as shown in Fig. 5e, the total effective electrical conducting area in the anode chamber was calculated
to be 262% in comparison to the at electrode area.
3.3. MFC assembly
Miniaturized MFCs were assembled as shown in Fig. 7. The size
of the anode and cathode electrodes were 10 mm 10 mm and both
electrodes have a SiO2 ring structure to limit the active electrode
area to 3 mm 3 mm as shown. The cathode used electrodes with
at surfaces, and the anode had electrodes with one of the various
micro/nano-structured surfaces. Both anode and cathode chambers
were constructed by using polydimethylsiloxane (PDMS) and the
volume of the chamber was 4 mm 5 mm 2 mm (40 L). Inlet and
outlet tubes were embedded in the PDMS layer to supply aqueous
solutions into the anode and cathode chambers. A 25-m-thick
PEM [20] was placed separating the electrodes in the middle of the
device as shown and all components were sandwiched by using
mechanical clips. The total device thickness was about 5 mm.
4. Results
4.1. Reference MFC with at electrodes

Fig. 5. Fabrication processes of CNT electrodes. Vertically aligned CNT forests were
selectively synthesized on molybdenum conductive layer.

Before electrical measurements, the anode and cathode chambers were lled with anolyte and catholyte, respectively. The
anolyte consisted solely of anaerobic minimal media lacking carbon sources. The catholyte consisted of potassium ferricyanide at
high concentration (1 M) to minimize the probability of cathode

S. Inoue et al. / Sensors and Actuators A 177 (2012) 3036

33

Fig. 6. SEM images of vertically aligned CNT electrodes.

Fig. 7. (left) Schematic diagram of the micro MFC where contact electrodes were built on a silicon chip of 10 mm 10 mm. (right) A fully assembled MFC with a one-cent US
coin for size comparison.

4.2. MFCs with micro/nano electrodes


To evaluate the performance enhancement of MFCs with different micro and nano structural congurations, the short circuit
current was monitored in each fabricated MFC. Since the current
production can uctuate due to metabolic conditions and environmental factors, the batch to batch variability was mitigated by

(A) Flat electrode

Pmax=0.32 W

0.4
MFC
Cathode
Anode
Measurement circuit

Voc=390 mV

300

0.3

200

0.2

Isc=6.8 A

100

Power (W)

400

Voltage (mV)

limited reactions. High concentration of ferricyanide might erode


and damage gold electrodes and lower concentration such as
50 mM might be more suitable for long-term operations. Both the
chambers were lled using a vacuum assisted process from outlet tubing to completely remove bubbles in the chambers [21]. The
bacterium, G. sulfurreducens, was initially cultured using anaerobic
minimal media in a PIPES buffer with 10 mM acetate and 40 mM
fumarate as electron acceptors at 30 C. The bacteria were then
added to the anolyte just after initiating electrical measurements.
First, the current-voltage (polarization) response of a reference
MFC with at electrodes was measured. After G. sulfurreducens was
inoculated to the anode chamber and the open circuit potential was
stabilized, a polarization curve was acquired using a Gamry Reference 600 Potentiostat as shown in Fig. 8. The resistance between
electrodes was lowered stepwise, pausing at each resistance setting
for 5 min. The open circuit voltage Voc and short circuit current Isc
were measured as 390 mV and 6.8 A, respectively. The lower Voc
as compared to the values of 0.460.8 V as reported in other MFCs
[22,23] could stem from oxygen penetrating through the PDMS,
ferricyanide leakage through the PEM, or pH shifts in either electrolyte within the L chambers. The maximum output power Pmax
was calculated as 0.32 W at 25 k-load, corresponding to a power
density of 3.6 W/cm2 . This power density is reasonable compared
with the previous L-sized MFCs which also used G. sulfurreducens
as the bio-catalysts (4.6533 W/cm2 ) [22]. Recently, much higher
power density has been achieved in mL-scaled MFCs with G. sulfurreducens by using mixed communities [24] and a specic strain
of G. sulfurreducens [25], to about 190 W/cm2 and 390 W/cm2 ,
respectively.

0.1

0.0
0

Current (A)
Fig. 8. Measurement results of the currentvoltage and currentpower relationship
from a MFC using a at-reference electrode. The inset shows the circuit connection in
currentvoltage measurements. The peak power was 0.32 W at 25 k-load, which
corresponds to a power density of 3.6 W/cm2 .

34

S. Inoue et al. / Sensors and Actuators A 177 (2012) 3036

1500

1000
(A) Flat electrode

Current (nA)

Current (nA)

(A) Flat electrode


(B) Hole-shape electrode
1000

(D) CNT electrode

Bacteria
injection

500

247%

Bacteria

500

205%

injection

60.3%

Background current

0
0

10

Time (Hour)
Fig. 9. Measurement results of short circuit current on (B) a hole-shape electrode.
The peak currents for (A) and (B) were 1037 nA and 625 nA, respectively, such that
the ratio of maximum short circuit current for (B) a hole-shape electrode versus (A)
a at-reference electrode was 60.3%.

simultaneously running MFCs with different contact electrodes and


a reference, at-surface MFC.
The experiments measuring the output current versus time for
the MFC operation using the hole-shape electrode were conducted
concurrently as shown in Fig. 9. G. sulfurreducens was simultaneously inoculated into anode chambers of both at- and hole-shape
electrodes for comparison in parallel. The output current of the
at-shape electrode as the reference MFC rapidly increased after
the injection of bacteria. However, the increase of output current
from the hole-shape electrode was delayed for about 2 h. The peak
current for the MFC using the hole-shape electrode occurred about
1 h later than the occurrence of the peak output current from the
MFC using at-reference electrode. Roughly 4 h after the inoculation of G. sulfurreducens, current production in both MFCs gradually
decreased due to the lack of fuel for the bacteria. If additional fuel
such as acetate is supplied into the anode chamber, the current
production returns to its peak level. The maximum short circuit current for the hole-shape electrode in conguration (B), 625 nA was
60.3% of the maximum short circuit current for the at-reference
electrode in conguration (A), 1037 nA.
In a similar experiment, the measurements of short circuit current from MFCs using the channel-shape and at-shape electrodes
were conducted as shown in Fig. 10. In this case, the short circuit

Fig. 11. Measurement results of short circuit current of (D) a CNT electrode. The
peak currents for (A) and (D) were 349 nA and 714 nA, respectively, such that the
maximum short circuit current for (D) a CNT electrode versus (A) a at-reference
electrode was 205%. If the background current of 100 nA is deducted from the output
current of both MFCs, the enhancement of current output by the CNT electrode
versus a at-reference electrode is 247%.

current of both MFCs increased after the injections of bacteria with


minimum delay. The peak currents for electrode congurations
(A) and (C) in Fig. 10 were 158 nA and 218 nA, respectively, such
that the maximum short circuit current ratio for the channel-shape
electrode conguration (C) against the at-reference electrode conguration (A) was 138%. In contrast to the MFC with hole-shape
electrode (Fig. 9), the peak output current occurred about the same
time for both MFCs using channel-shape and at-reference electrodes. About 2 h after the inoculation of G. sulfurreducens, current
production in both MFCs gradually decreased due to the lack of fuel
for the bacteria.
Concurrent measurements on MFCs using CNT and the atreference electrodes were also conducted as shown in Fig. 11. In
this experiment, the cathode electrodes of both MFCs and the anode
electrode for the reference MFC had used at-shape electrodes that
were covered with molybdenum. The anode and cathode electrodes
for both MFCs had not used the SiO2 ring design as illustrated
in Fig. 7. Experimental results have shown that the background
current was quite high (about 100 nA) as compared with the previous experiments on the hole-shape and channel-shape electrodes,
probably due to the reactivity of the molybdenum on the electrodes
as elemental Mo can oxidize with water as,
MoO2 (s) + 4H+ + 4e  Mo(s) + 2H2 O (E o = 0.15 V)

300

Current (nA)

(C) Channel-shape electrode


200

Bacteria
injection

138%

5. Discussion

0
0

(1)

However, the enhancement of current outputs by the CNT electrode was obvious. The peak currents for electrode congurations
(A) and (D) in Fig. 11 were 349 nA and 714 nA, respectively, such
that the maximum short circuit current ratio for the CNT electrode
conguration (D) against the at-reference electrode conguration (A) was 205%. If the background current of 100 nA is deducted
from output current of both MFCs, the current output by the CNT
electrode is 247% more than the at-reference electrode.

(A) Flat electrode

100

10

Time (Hour)

10

Time (Hour)
Fig. 10. Measurement results of short circuit current on (C) a channel-shape electrode. The peak currents for (A) and (C) were 158 nA and 218 nA, respectively, such
that the maximum short circuit current for (C) a channel-shape electrode versus (A)
a at-reference electrode was 138%.

The enhancement of current density in MFCs and effectiveness


of charge transfer between living cells with pili (G. sulfurreducens)
and micro/nano interfacial electrodes have been characterized by
experimental measurements in this work. The condition of the
cells, including their activities, concentrations, adhesion abilities
to the substrate and other factors all affect the current generation capabilities. Hence, to minimize experimental variations due
to these factors, simultaneous experiments on different electrode

S. Inoue et al. / Sensors and Actuators A 177 (2012) 3036

35

Normalized maximum
short circuit current

(C) Channel
2
(B) Hole

or
op
r
P

tio

n
(D) CNT
deducted
background current

(D) CNT
not deducted
background current

(A) Flat electrode


(Reference)
0
0

Normalized surface area


Fig. 12. Experimental results of MFCs showing maximum output current vs. cellto-electrode contact area of various MFCs. The red circle for (D) CNT indicates the
normalized maximum short circuit current without deducting the background current and the blue square for (D) CNT indicates the normalized maximum short circuit
current after deducting the background current of 100 nA. The linear relationship
supports the hypothesis that only bacteria attached to the electrode surface can
donate electrons and this is likely the key mechanism of charge transfer. (For interpretation of the references to colour in this gure legend, the reader is referred to
the web version of this article.)

congurations were conducted. Nevertheless, the peak currents


being generated by the at-electrode conguration (A) differed in
our experiments likely due to the usage of cells under different
culturing periods. In spite of these differences, current ratios from
the simultaneous experiments of different electrode congurations provide valuable insight. Fig. 12 summarizes the experimental
results of maximum short circuit currents (normalized to current
output of at-reference electrode) from various MFCs with different electrode structures versus the calculated surface areas of
individual corresponding electrodes (normalized to at-reference
electrode). The spot (A) reects the result of simultaneous measurement of two identical reference MFCs which have exactly same
electrode surface area, and the value is close to 1 as expected. The
small discrepancy from 1 indicates the possible experimental variability in concurrent measurements. The rest of the data follows
the expected linear relationship between the current and effective
surface area. The red circle for (D) CNT indicates the normalized
maximum short circuit current without deducting the background
current and the blue square for (D) CNT indicates the normalized
maximum short circuit current after deducting the background
current of 100 nA in the same way as Fig. 11. Although the red circle for (D) CNT is slightly smaller than the expected value from
the effective surface area of the electrode, the blue square for (D)
CNT nearly follows the linear relationship between the current and
effective surface area. This result suggests that the background
current should be deducted for evaluating the enhanced current
outputs.
As no added electron mediator was added in the setup, the result
of the linear relationship supports the hypothesis that the main
charge transfer mechanism is from bacteria attaching to the electrode surface to donate electrons. The CNT electrodes improved the
current density about 200% as the effective surface area increased
about 200%. It has been reported, however, that CNTs have a cellular toxicity that could lead to proliferation inhibition and cell death
[26,27]. In this work, the toxicity of CNTs has not been observed
probably because the run time was short (less than 10 h). For longterm applications, the modications of electrodes with CNTs may
be necessary to reduce the cellular toxicity by coating with some
other materials, such as polyaniline [8]. The CNT forests used in
this work were of high density, hence, via size exclusion, G. sulfurreducens could have penetrated in-between CNT walls but not
within CNT forests. One possible future direction is to synthesize

Fig. 13. Schematic of bacteria fully lled into a CNT forest with the dimension
of 10 mm 10 mm 1 mm. The maximum output power is estimated as high as
12 mW from the measured power density of 3.6 W/cm2 (the diameter of bacteria
is 0.3 m).

CNT forests with larger gaps to accommodate a higher density of


bacteria. For example, if bacteria are fully lled into CNT forests
with the dimension of 10 mm 10 mm 1 mm as shown in Fig. 13,
the maximum output power is roughly estimated as high as 12 mW
from the measured power density of 3.6 W/cm2 in Fig. 8 (the
diameter of bacteria is 0.3 m). This prospective power density is
enough to for powering not only most autonomous microsystems,
but also small portable electronics.
6. Conclusions
More than a 200% current density enhancement in miniaturized microbial fuel cells has been successfully demonstrated by
optimizing the contact electrode structure using micro and nano
features. The data demonstrates a linear relationship between the
current production and effective surface area of electrodes. This
result supports the hypothesis that the charge transfer mechanism
is based on direct attachment of bacteria to the electrode surface
for electron transport. We estimate that the volumetric power density of MFCs using CNT forests with larger gaps could be as high as
12 mW/100 mm3 .
Acknowledgements
This work was performed at the University of California, Berkeley during 20092010 when Shogo Inoue worked for FUJITSU
LABORATORIES LTD. The project was partially supported by the
Sustainable Products and Solutions Program at the University of
California, Berkeley and the University of California Presidents
Postdoctoral Fellowship (Cullen R. Buie).
References
[1] D.R. Bond, D.E. Holmes, L.M. Tender, D.R. Lovley, Electrode-reducing microorganisms that harvest energy from marine sediments, Science 295 (2002)
483485.
[2] S. Choi, H. Lee, Y. Yang, P. Parameswaran, C.I. Torres, B.E. Rittmann, J. Chae, A
L-scale micromachined microbial fuel cell having high power density, Lab.
Chip 11 (2011) 11101117.
[3] Y. Chen, Y. Zhao, K. Qiu, J. Chu, R. Lu, M. Sun, X. Liu, G. Sheng, H. Yu, J. Chen, W. Li,
G. Liu, Y. Tian, Y. Xiong, An Innovative miniature microbial fuel cell fabricated
using photolithography, Biosens. Bioelectron. 26 (2011) 28412846.
[4] M. Chiao, K.B. Lam, L. Lin, Micromachined microbial fuel cells, in: Proc. IEEE
MEMS Conference, Kyoto, 2003, pp. 383386.
[5] M. Chiao, K.B. Lam, L. Lin, Micromachined microbial and photosynthetic fuel
cells, J. Micromech. Microeng. 16 (2006) 25472553.
[6] E. Parra, L. Lin, Microbial fuel cell based on electrode-exoelectrogenic bacterial
interface, in: Proc. IEEE MEMS Conference, Sorrento, 2009, pp. 3134.
[7] K.B. Lam, E. Johnson, L. Lin, A MEMS photosynthetic electrochemical cell
powered by sub-cellular plant photosystems, IEEE/ASME JMEMS 15 (2006)
12431250.
[8] Y. Qiao, C.M. Li, S. Bao, Q. Bao, Carbon nanotube/polyaniline composite as anode
material for microbial fuel cells, J. Power Sources 170 (2007) 7984.

36

S. Inoue et al. / Sensors and Actuators A 177 (2012) 3036

[9] T. Sharma, A.L.M. Reddy, T.S. Chandra, S. Ramaprabhu, Development of carbon


nanotubes and nanouids based microbial fuel cell, Int. J. Hydrogen Energy 33
(2008) 67496754.
[10] J.L. Lamp, J.S. Guest, S. Naha, K.A. Radavich, N.G. Love, M.W. Ellis, I.K. Puri, Flame
synthesis of carbon nanostructures on stainless steel anodes for use in microbial
fuel cells, J. Power Sources 196 (14) (2011) 58295834.
[11] P. Liang, H. Wang, X. Xia, X. Huang, Y. Mo, X. Cao, M. Fan, Carbon nanotube powders as electrode modier to enhance the activity of anodic biolm in microbial
fuel cells, Biosens. Bioelectron. 26 (2011) 30003004.
[12] S. Oh, B. Min, B.E. Logan, Cathode performance as a factor in electricity generation in microbial fuel cells, Environ. Sci. Technol. 38 (2004) 49004904.
[13] S. Oh, B.E. Logan, Proton exchange membrane and electrode surface areas as
factors that affect power generation in microbial fuel cells, Appl. Microbiol.
Biotechnol. 70 (2006) 162169.
[14] K.B. Lam, E.F. Irwin, K.E. Healy, L. Lin, Bioelectroanalytic self-assembled thylakoids for micro power and sensing applications, Sens. Actuators B: Chem.
117 (2006) 480487.
[15] S. Inoue, E.A. Parra, A. Higa, L. Lin, Cell-to-electrode contact structures for power
density enhancements in microbial fuel cells, in: Proc. IEEE MEMS Conference,
Cancun, 2011, pp. 12971300.
[16] G. Reguera, K.D. McCarthy, T. Mehta, J.S. Nicoll, M.T. Tuominen, D.R. Lovley,
Extracellular electron transfer via microbial nanowires, Nature 435 (2005)
10981101.
[17] G. Reguera, K.P. Nevin, J.S. Nicoll, S.F. Covalla, T.L. Woodard, D.R. Lovley, Biolm
and nanowire production leads to increased current in Geobacter sulfurreducens
fuel cells, Appl. Environ. Microbiol. 72 (2006) 73457348.

[18] K.P. Nevin, B. Kim, R.H. Glaven, J.P. Johnson, T.L. Woodard, B.A. Methe, R.J. DiDonato, S.F. Covalla, A.E. Franks, A. Liu, D.R. Lovley, Anode biolm transcriptomics
reveals outer surface components essential for high density current production
in Geobacter sulfurreducens fuel cells, PLoS One 4 (2009) 111.
[19] Y.Q. Jiang, Q. Zhou, L. Lin, Planar MEMS supercapacitor using carbon nanotube
forests, in: Proc. IEEE MEMS Conference, Sorrento, 2009, pp. 587590.
[20] DuPontTM Naon PFSA membrane NR211(1 mil).
[21] Y.Q. Jiang, P. Wang, J. Zhang, W. Li, L. Lin, 3D supercapacitor using nickel electroplated vertical aligned carbon nanotube array electrode, in: Proc. IEEE MEMS
Conference, Hong Kong, 2010, pp. 11711174.
[22] S. Choi, J. Chae, A series array of microliter-sized microbial fuel cell, in: Proc.
IEEE MEMS Conference, Cancun, 2011, pp. 12891292.
[23] D.R. Bond, D.R. Lovley, Electricity production by Geobacter sulfurreducens
attached to electrodes, Appl. Environ. Microbiol. 69 (2003) 15481555.
[24] K.P. Nevin, H. Richter, S.F. Covalla, J.P. Johnson, T.L. Woodard, A.L. Orloff, H. Jia,
M. Zhang, D.R. Lovley, Power output and columbic efciencies from biolms of
Geobacter sulfurreducens comparable to mixed community microbial fuel cells,
Environ. Microbiol. 10 (2008) 25052514.
[25] H. Yi, K.P. Nevin, B. Kim, A.E. Franks, A. Klimes, L.M. Tender, D.R. Lovley, Selection of a variant of Geobacter sulfurreducens with enhanced capacity for current
production in microbial fuel cells, Biosens. Bioelectron. 24 (2009) 34983503.
[26] E. Flahaut, M.C. Durrieu, M. Remy-Zolghadri, R. Bareille, C.H. Baquey, Study of
the cytotoxicity of CCVD carbon nanotubes, J. Mater. Sci. 41 (2006) 24112416.
[27] A. Magrez, S. Kasas, V. Salicio, N. Pasquier, J.W. Seo, M. Celio, S. Catsicas, B.
Schwaller, L. Forro, Cellular toxicity of carbon-based nanomaterials, Nano Lett.
6 (2006) 11211125.

You might also like