You are on page 1of 9

ARTICLE

Energy From Algae Using Microbial Fuel Cells


Sharon B. Velasquez-Orta,1 Tom P. Curtis,1 Bruce E. Logan2
1

School of Civil Engineering and Geosciences, Newcastle University,


Newcastle upon Tyne NE17RU, United Kingdom; telephone: 44-191-222-6415;
fax: 44-191-222-6502; e-mail: s.b.velasquez-orta@ncl.ac.uk
2
Department of Civil and Environmental Engineering, The Pennsylvania State University,
University Park, Pennsylvania 16802
Received 3 December 2008; revision received 19 February 2009; accepted 27 March 2009
Published online 3 April 2009 in Wiley InterScience (www.interscience.wiley.com). DOI 10.1002/bit.22346

ABSTRACT: Bioelectricity production from a phytoplankton,


Chlorella vulgaris, and a macrophyte, Ulva lactuca was
examined in single chamber microbial fuel cells (MFCs).
MFCs were fed with the two algae (as powders), obtaining
differences in energy recovery, degradation efficiency, and
power densities. C. vulgaris produced more energy generation
per substrate mass (2.5 kWh/kg), but U. lactuca was degraded
more completely over a batch cycle (73  1% COD).
Maximum power densities obtained using either single cycle
or multiple cycle methods were 0.98 W/m2 (277 W/m3) using
C. vulgaris, and 0.76 W/m2 (215 W/m3) using U. lactuca.
Polarization curves obtained using a common method of
linear sweep voltammetry (LSV) overestimated maximum
power densities at a scan rate of 1 mV/s. At 0.1 mV/s,
however, the LSV polarization data was in better agreement
with single- and multiple-cycle polarization curves. The
fingerprints of microbial communities developed in reactors
had only 11% similarity to inocula and clustered according to
the type of bioprocess used. These results demonstrate that
algae can in principle, be used as a renewable source of
electricity production in MFCs.
Biotechnol. Bioeng. 2009;103: 10681076.
2009 Wiley Periodicals, Inc.
KEYWORDS: microbial fuel cell; algae; bioenergy; substrate
composition; polarization

Introduction
The concern about global warming effects and fossil fuel
costs has encouraged the search for alternative sources of
energy (Brecha, 2008). Biomass products tested for energy
generation include a wide range of growth plants, crops and
Correspondence to: S.B. Velasquez-Orta
Contract grant sponsor: Consejo Nacional de Ciencia y Technologia (CONACyT)
Contract grant number: 196298
Contract grant sponsor: National Science Foundation
Contract grant number: CBET-0730359
Contract grant sponsor: King Abdullah University of Science and Technology (KAUST)
Global Research Partnership
Contract grant number: KUS-I1-003013

1068

Biotechnology and Bioengineering, Vol. 103, No. 6, August 15, 2009

wastes (Deublein and Steinhauser, 2008). In this context,


algae are seen as an alternative to land based alternatives.
The cultivation of algae has several advantages over
terrestrial plants; they require less space (1/7th less surface
area), have higher growth rates, and do not compete with
food production (Hartman, 2008). Algae are divided into
two groups: phytoplankton (microalgae), or macrophytes
(macroalgae). Microalgae are unicellular green plants rich
in chlorophyll that lack lignin or cellulose, and contain
proteins, carbohydrates and lipids in strain-specific proportions (Schenk et al., 2008). They are abundant in oceans,
suitable for cultivation in rivers, and serve as the primary
source of carbohydrates and proteins for aquatic organisms.
Macroalgae are more resistant to predators and environmental conditions than microalgae. They are abundant
in costal zones (Riegman et al., 1993), lack lignin, and
largely consist of polysaccharides (alginate, laminaran, and
mannitol) and unsaturated fatty acids which are easily
hydrolyzed, and low concentrations of cellulose (VergaraFernandez et al., 2008). Both macrophytes and phytoplankton are suitable for cultivation using river water, sea
water, and some wastewaters (de-Bashan et al., 2004; Walker
et al., 2005).
The generation of energy products from microalgae and
macroalgae has been examined by a number of workers.
Microalgae have been tested as raw material for the
production of bio-oil (Li et al., 2007), methane (Golueke
and Oswald, 1959; Minowa and Sawayama, 1999), methanol
(Hirano et al., 1998) and hydrogen (Kim et al., 2006; Turner
et al., 2008). Macroalgae have primarily been used to
produce methane (Hansson, 1983; Morand and Briand,
1999; Troiano et al., 1976). A disadvantage of all these
technologies is that the fuel produced must be stored,
transported and further processed to produce electricity.
Microbial fuel cells (MFCs) offer an alternative way to
obtain electricity from the hydrolysis and fermentation of
algae in only one process unit. MFCs consist of an anode and
cathode connected by a load (usually a resistor in laboratory
studies). The anode contains mixed or pure cultures of
2009 Wiley Periodicals, Inc.

microorganisms that are used to catalyze the decomposition


of the organic matter into electrons and protons. Power is
produced through the reduction of oxygen or another
chemical at the cathode. A metal is usually used to catalyze
oxygen reduction, although it has recently been shown that
microorganisms can be used for this purpose as well (He and
Angenent, 2006).
In this study, we evaluated the performance of MFCs
using two different types of algae as substrates. Chlorella
vulgaris (a microalgae) and Ulva lactuca (a macroalgae) were
selected because they have been widely tested in other
technologies for energy generation. Moreover, they have
very different organic matter composition, with C. vulgaris
containing more than 50% protein (Becker, 2007) and
U. lactuca having around 60% carbohydrates (Ventura and
Castanon, 1998). The substrate degradation and microbial
composition in MFCs and in anaerobic reactors were
compared. To better understand the degradation process,
we monitored the production of by-products (biogas and
volatile fatty acids) using U. lactuca. To estimate the
maximum power obtainable, we used different methods to
obtain polarization and power density curves. We found
that potentiostat methods usually employed for hydrogen
fuel cells can overestimate maximum power production by
MFCs with algae at scan rates often used in MFC tests.

Methods
Reactors Construction
Twelve reactors were used, each having a liquid volume of
25 mL. Four MFCs were operated at closed circuit (CC)
using a circuit of titanium wire containing a resistor of
1,000 V (except as noted), four were operated in open
circuit (OC) (no connection), and four were constructed to
be completely anaerobic reactors (ARs) by sealing the
reactors with an end plate. MFCs and ARs had graphite fiber
brush anodes 3.0 cm in outer diameter and 3.0 cm long
(PANEX33 160K, ZOLTEK) (Logan et al., 2007) that were
treated using a high-temperature ammonia gas process
(Cheng and Logan, 2007). MFCs had air-cathodes that were
prepared according to the procedures of Cheng et al.
(2006), with a platinum (Pt) catalyst (0.5 mg/m2 Pt) and
four diffusion layers. All materials were initially sterilized
(using UV light for 2 h or autoclaved at 1218C for 15 min);
although all tests were run using mixed cultures under nonsterile conditions.

Algae Characteristics and Medium Preparation


U. lactuca was purchased as a powder with homogenous
particles of 90200 mm (SigmaAldrich, St. Louis, MO). It
consisted of 317  45 mg/g-substrate of total carbohydrates,
65  1 mg/g-substrate of total protein and had a COD of
833 mg/g-substrate. C. vulgaris was also obtained in a

powder form (Federal Laboratory Corporation, New York,


NY) with particle sizes in the same range as those for
U. lactuca. It consisted of 157  46 mg/g-substrate of total
carbohydrates, 428  15 mg/g-substrate of total protein and
had a COD of 1587 mg/g-substrate. Both materials were
certified to contain the same composition as the natural
sources. Pre-treatment of the U. Lactuca consisted of drying
in a hot-drum (losing 93% of total water) followed by
grinding in stainless steel and tungsten carbide mills.
C. Vulgaris was pre-treated using spray drying. Samples were
stored in a cool (48C) dry environment. No chemical
treatment processes were used. The use of algae in a powder
form ensured MFC efficiencies could be directly compared
based on substrate composition, and not for example algae
cell size.
Algae powders were added to in a medium containing
(g/L): NH4Cl, 0.31; KCl, 0.13; NaH2PO4H2O, 2.45 and
Na2HPO47H2O, 4.57. The conductivity of the medium was
4.9 mS/cm, and it had a pH of 6.8. No additional minerals or
vitamins were added to the medium.

Reactors Inoculation and Operation


Reactors were inoculated using primary clarifier overflow
from the Pennsylvania State University wastewater treatment plant in State College. All reactors were operated in
fed-batch mode at a fixed temperature of 308C. Voltage was
monitored across the resistor every 20 min. Data were
obtained when MFCs produced a repeatable cycle of current
generation. The solution was replaced when the voltage was
reduced to <50 mV.
Maximum power densities were evaluated using three
different methods, all using a high organic matter
concentration (2,500 mg COD/L). The single cycle
method consisted of varying the resistance (between
7,500 and 100 V) from the open circuit voltage (OCV)
every 20 min in a single batch cycle, over a short time period
where the substrate change did not affect power output
(Heilmann and Logan, 2006). For the multiple cycle
method, a single cycle was run for each resistor (Heilmann
and Logan, 2006). The maximum power density value was
taken at steady state conditions, when the current started
to decrease the solution was replaced before using a new
resistor. These two methods were compared to a commonly
used method based on linear sweep voltammetry (LSV)
(Aelterman et al., 2008; Chen et al., 2008; Clauwaert et al.,
2008; Logan et al., 2006). For LSV, the potential was
decreased using a potentiostat at a rate decrease of 1 or
0.1 mV s1 from the OCV to a cell potential of 100 mV.
Substrate degradation was monitored over a batch cycle
for the highest substrate concentration (2,500 mg COD/L)
in terms of chemical oxygen demand (COD), protein,
and carbohydrate concentrations. To assess by-products
formation among the different types of reactors used,
volatile fatty acids (VFAs) and headspace gas composition
were monitored in reactors using U. lactuca. A plastic cap

Velasquez-Orta et al.: Energy From Algae Using MFCs


Biotechnology and Bioengineering

1069

was fitted on the top of selected reactors to create a


headspace for gas accumulation. Gases were sampled using a
gastight syringe (250 mL, Hamilton Samplelock Syringe).
Bulk solution (0.5 mL) was obtained using a sterile syringe,
filtered using glass fiber filters (Whatman, GF/C, 4.7 cm of
diameter, 1.2 mm of pore size), and stored at 208C. Once
experiments were completed, samples were obtained to
characterize the microbial community as described below.

Chemical Analyses
Total and soluble protein, carbohydrate and chemical
oxygen demand (COD) concentrations were monitored for
soluble and total fractions of the bulk liquid according to
Standard Methods (American Public Health Association
(APHA), 1995). For the soluble fraction, samples were prefiltered using glass fiber filters (Whatman, GF/C, 4.7 cm of
diameter, 1.2 mm of pore size). Total and soluble COD
analyses were carried out using method 5220 (Hach COD
system, Hach Company, Loveland, Colorado). Soluble and
total protein concentrations were quantified using the
bicinchoninic acid protein assay kit (SigmaAldrich). A
calibration curve was prepared using standard solutions
ranging from 0.1 to 1.0 g/L of bovine serum albumin.
Soluble and total carbohydrate concentrations were
measured using a colorimetric method based on an
anthrone reagent (Gerhardt et al., 1994). For this measurement, 1 mL of sample was transfer to a COD tube and 2 mL
of a chilled H2SO4 solution (75%, v/v) were added. After
vortexing (30 s), 4 mL of a chilled fresh anthrone solution (2
g/L, 75% (v/v) H2SO4) was added and the sample vortexed
again. COD tubes were placed in a heating-block at 1008C
for 15 min, cooled to room temperature, and analyzed
using disposable cuvettes at 578 nm (UV-1601, Shimadzu
Corporation). Every measurement included a calibration
curve using glucose at concentrations of 503,000 mg/L
(R2 0.99).
Volatile fatty acid (VFA) concentrations (acetate, butyrate
and propionate) were measured in triplicate using a
gas chromatograph (Agilent 6890N) and a 30 m 
0.32 mm  0.5 mm fused-silica capillary column. Before
GC analysis, 50 mL 50% formic acid (v/v in water) were
added to 1 mL samples. Gases (carbon dioxide, methane,
and nitrogen) were analyzed using a gas chromatograph
(helium carrier gas; model 310, SRI Instruments). Since
nitrogen served as a dilution gas, it was removed from the
calculations in order to determine the CO2 and CH4
headspace fractions.

for 30 s and fibers removed. Then, tubes were centrifuged at


5,000g for 7 min. Total DNA was extracted using the
Powersoil DNA isolation kit (Mobio Laboratories, Carlsbad,
CA) and stored at 208C. 16S rRNA gene fragments were
amplified from DNA samples using 16S rRNA gene
fragments amplified using primer 3 50 -CCT ACG GGA
GGC AGC AG-30 with a GC-clamp, and primer 2 50 -ATT
ACC GCG GCT GCT GG-30 (Muyzer et al., 1993), and
analyzed by denaturing gradient gel electrophoresis (DGGE)
with a denaturing gradient ranging from 30% to 55% (100%
denaturant is 7 M urea plus 40% (v/v) formamide in 1
TAE; 40 mM Tris-acetate, 1 mM EDTA, pH 8). DGGE gels
were processed using the Bionumerics software package
(version 3.5, Applied Maths, Austin, TX) to determine the
position and intensity of all bands in all DGGE profiles in
relation to markers run alongside samples in the gel
(Verseveld and Roling, 2008) and to analyze bands present
in DGGE profiles using Dice cluster analysis.

Calculations
MFC potentials were obtained using a data acquisition
system (2700, Keithly, Cleveland, OH). Potentials were
converted to current using Ohms law, E IR, where E is the
voltage, I is the current, and R is the resistance. Power
densities ( Pd) were obtained using the equation: Pd IV/
AAn, where AAn is the cathode surface area (7.07 cm2).
Columbic efficienciesR (Ec) were calculated using (Logan
t
et al., 2006): Ec M 0f I dt=FzvAn DCOD, where: M 32, is
the molecular weight of oxygen, z 4 the number of
electrons transferred per mole of oxygen, F 96485.4
A mol1 Faradays constant, vAn the volume of liquid in
the anode compartment, and DCOD the change in COD
over the batch period of time. Statistical analysis of data was
performed using Minitab1 15.1.0.0 for Windows using
analysis of variance (ANOVA). The regression analysis
relating maximum power densities algae concentrations
was assessed using Sigmaplot1 9.0 for Windows. The
first order exponential rise-to-maximum model gave
the best fit (R2 0.95  0.05) according to the equation:
Pd Pmax 1  ekCS , where Pmax is the maximum power
density, k is the growth rate constant, and Cs is the substrate
concentration.

Results
MFCs Performance Using Different Algae
Concentrations

Microbial Analyses
Biofilms from the anodes were removed from MFCs by
cutting off anode fibers in a sterile laminar flow cabinet.
Fibers were immediately placed in 15 mL centrifuge
tubes containing sterile PBS solution (130 mM NaCl,
10 mM Na2HPO4, pH 7.2). Centrifuge tubes were vortexed

1070

Biotechnology and Bioengineering, Vol. 103, No. 6, August 15, 2009

A repeatable cycle of current generation was obtained by


C. vulgaris and U. lactuca (510 mg COD/L) after five batch
cycles (6 days). The maximum current for each batch cycle
was obtained within 0.76.7 h depending on the substrate
concentration. Batch cycles for MFCs using different
concentrations of C. vulgaris required 25 days, while those

for U. lactuca lasted 27 days (data not shown). When MFCs


were fed algae at a concentration of 2,500 mg COD/L, a
steady state current generation was rapidly achieved (Fig. 1).
MFCs produced a steady state current between 0.52 and
0.55 mA for a period of 3 days in MFCs using U. lactuca, and
1.5 days using C. vulgaris. Following this steady output,
current generation gradually decreased from 0.5 to 0.05 mA
over a period of 4 days.
Maximum power densities increased with substrate
concentration until reaching a plateau (Fig. 2). At high
organic matter concentrations the maximum power
densities ( Pmax) obtained using the different algae were in
the same range, with 440 mW  60 for MFCs using
C. vulgaris and 450 mW  50 for MFCs using U. lactuca.
MFCs fed C. vulgaris had a higher power growth rate
constant (k 2:6  103  0:0012) and hence achieved
Pmax at lower organic matter concentrations than MFCs
using U. lactuca (k 0:9 103  0:0003).
Coulombic efficiencies peaked at relatively low algae
concentrations (from 100 mg COD/L and 500 mg COD/L)
with a maximum of 28% for C. vulgaris and 23% for
U. lactuca (Fig. 3). At substrate concentrations higher than
600 mg/L, coulombic efficiencies decreased and reached a
plateau at 10%. COD removal increased to a maximum of
85  5% at a loading of 1,000 mg COD/L.

Figure 2. Maximum currents with Chlorella vulgaris and Ulva lactuca as a


function of COD concentration.

There were no significant differences in the maximum


power produced between the single cycle and multiple cycle
methods ( P 0.45, ANOVA; Fig. 4). The average maximum
power density produced using these two methods was of
0.98  0.39 W/m2 (277 W/m3) for MFCs using C. vulgaris in

a current density between 0.20 and 0.25 mA/ cm2, and


0.76  0.15 W/m2 (215 W/m3) for MFCs using U. lactuca at
a current density of 0.20 mA/cm2 (Fig. 4). The maximum
power density obtained using LSV was dependant on the
scan rate selected (Fig. 5). A scan rate of 1 mV/s produced
a very high maximum power density of 1.6  0.3 W/m2
(453 W/m3). At a scan rate of 0.1 mV/s, the power density
was much lower (0.9  0.2 W/m2, 255 W/m3; P 0.007,
ANOVA). Based on a comparison of these LSV scan results
with other methods (single and multiple cycles), it appears
that a scan rate of 0.1 mV/s is a better choice than 1 mV/s for
obtaining polarization curves (Fig. 4). The slower scan
rates of 0.1 mV/s required 34 h to complete a polarization
curve. It is likely that a scan rate of 1 mV/s is too fast (30 min
for a cycle) for the bacteria to respond to the changes in
voltage.

Figure 1. MFCs current generation with 2,500 mg COD/L of Chlorella vulgaris or


Ulva lactuca. Reactors were operated using an external resistance of 1,000 V. [Color
figure can be seen in the online version of this article, available at www.interscience.
wiley.com.]

Figure 3. Effect of substrate concentration in CE and COD removal efficiency.


Filled symbols are for CE and empty symbols are for COD removal efficiency

Effect of Polarization Method on the Calculated


Maximum Power

Velasquez-Orta et al.: Energy From Algae Using MFCs


Biotechnology and Bioengineering

1071

Organic Matter Consumption

Figure 4.

MFCs power density outputs from algae using different


polarization methods. Reactors were fed a concentration of 2,500 mg COD/L of
C. vulgaris (a) and of U. lactuca (b). Measurements were done using different
resistances in duplicate reactors. Dotted lines are cell voltages and plain lines are
for power densities. [Color figure can be seen in the online version of this article,
available at www.interscience.wiley.com.]

Figure 5.

Comparison of short single cycle polarizations in MFCs using


2,500 mg COD/L Chlorella vulgaris at different scan rates. [Color figure can be seen
in the online version of this article, available at www.interscience.wiley.com.]

1072

Biotechnology and Bioengineering, Vol. 103, No. 6, August 15, 2009

Removal efficiencies were compared among three types of


reactors: operating MFCs or those with a closed circuit
(CC); MFCs kept in an open circuit (OC); and anaerobic
reactors (ARs) for one operational batch (Fig. 6). COD in
C. vulgaris was better removed in MFCs at CC (1,000 V)
than in other reactors ( P 0.00, ANOVA). Differences in
COD removals among reactors fed U. lactuca were not
significant ( P 0.60, ANOVA). COD removals in reactors
using C. vulgaris were two times higher in the operating
MFCs (60  6%) than in the AR (28%), while COD
removals in reactors using U. lactuca were in the same range
for all reactors (approximately 73  1%). Carbohydrate
removal using both types of algae was highest in operating
MFCs.
Protein removal from C. vulgaris was highest in the
operating MFCs (51  16%) and lowest in ARs (24  3%).
Protein removal from U. lactuca was highest in open circuit
MFCs (62  7%) followed by the closed circuit MFCs
(41  1%) and ARs (15  6%).
Concentrations of the substrate on the basis of soluble and
particulate matter were analyzed for the operating MFCs at
the start and at the end of the experiments (Fig. 7). Initially,
organic matter from both algae contained more particulate
organic matter than soluble organic matter. C. vulgaris
contained two times more protein than carbohydrates while
U. lactuca contained five times more carbohydrates than
protein. Soluble protein was the highest fraction consumed
(83%) in C. vulgaris. In contrast, particulate carbohydrate
was the highest fraction consumed (90%) in U. lactuca.
Total COD consumption was higher for MFCs using
U. lactuca than for MFCs using C. vulgaris. All MFCs
removed more than 50% of the soluble and particulate
COD.

Figure 6. Comparison of organic matter removal in MFCs at closed circuit


(CC), MFCs at open circuit (OCV) and anaerobic reactors (AR) on the basis of protein,
carbohydrates and COD using 2,500 mg/L of algae substrate. Measurements were
done in duplicates, if the difference between replicates was low errors bar do not
appear.

Figure 7. Initial and final organic matter content in MFCs. Organic matter was
measured on the basis of protein, carbohydrates and COD for one operational batch;
i initial and f final.

By-Products in MFCs and Anaerobic Reactors


Using U. lactuca
Differences in headspace composition and volatile fatty
acids production were assessed over one batch cycle for
MFCs and ARs fed 2,500 mg/L U. lactuca. The headspace
composition changed over a period of 7 days (96 h; Fig. 8a).
Methane concentrations were higher for ARs than for MFCs.
ARs reached a maximum methane concentration of 30%
between the 3rd and 5th day while methane concentrations
for MFCs remained below 5%. Carbon dioxide concentrations in MFCs reactors peaked on the first day (14%) and
gradually decreased until the end of the experiment. For ARs
carbon dioxide concentrations remained between 5% and
10%. Acetate concentrations rose rapidly during the first day
for all reactors, following by a gradual decrease in MFCs
(Fig. 8b). ARs had sharps increases and decreases in acetate
concentrations. Propionate was detected in low quantities
(<75 mg/L for all reactors) but no butyrate was detected.

Effect of Type of Reactor Used in


Microbial Communities
Bacterial communities were clustered (Fig. 9) according to
the bioprocess (MFCs or ARs). Communities from AR had
a greater similarity (59.4  11.2%) to MFCs operated
under open circuit conditions than to MFCs operated
for electricity production (45.8  8.3%). There was a low
similarity of microbial community profiles between the
inoculum and MFCs (10.4  0.9% similarity with CC;
11.0  4.9% with OC) or the AR (11.5  5.2% similarity).

Discussion
MFCs using mixed microbial cultures produced significant
power from microalgae and macroalgae when compared
to other substrates. Maximum power densities of

Figure 8. By-products produced in MFC and AR using 2,500 mg/L Ulva lactuca.
a: Headspace gas composition through time; squares: MFC and triangles: anaerobic
reactor. b: Acetate production through time.

0.98 W/m2  0.39 for C. vulgaris and 0.76 W/m2 


0.15 for U. lactuca were higher than those obtained using
paper wastewater (0.7 W/m2) in the same reactor configuration (Huang and Logan, 2008), but lower than those
obtained with acetate (2.4 W/m2) (Logan et al., 2007). The
difference in power generation among substrates is perhaps
due to the complexity of the substrates. Particulate
substrates and macromolecules require more energy for
degradation processes (hydrolysis and fermentation) than
simple soluble compounds that can be directly taken into
the cell.
One important result observed here is that maximum
power densities were over-predicted by polarization
curves obtained using a potentiostat at high scan rates. A
polarization scan rate of 1 mV/s using LSV gave a power
density 44% higher than that obtained at a slower rate of
0.1 mV/s or using fixed resistors for extended periods of
time. This result has important implications in MFCs
analysis since high scan rates have been used to evaluate
maximum power output in several recent MFC studies

Velasquez-Orta et al.: Energy From Algae Using MFCs


Biotechnology and Bioengineering

1073

Figure 9.

Dice cluster analysis of bands present in biofilms. Indicates similarities between the DGGE bands obtained from three different reactor configurations and two
different substrates (Chlorella vulgaris and Ulva lactuca). The cluster analysis was obtained using Bionumerics software.

(Aelterman et al., 2008; Chen et al., 2008; Clauwaert et al.,


2008). This indicates that polarization data developed using
LSV should be confirmed by data obtained using fixed
resistances over longer period of times. For example, in the
single-cycle method an interval of 20 min is used for each
resistor, allowing time for the bacteria to adjust to this new
load. At high scanning rates of 1 mV/s, the LSV approach
applies a potential change that is too rapid for some MFCs to
reach sufficiently steady conditions. Further work should be
done to determine if these changes can also be observed in
MFCs using simple substrates.
Better energy recovery was obtained with microalgae,
C. vulgaris, than with macroalgae, U. lactuca. MFCs fed with
microalgae produced greater power densities at a similar
COD concentration than the macroalgae. In addition,
the COD per gram of C. vulgaris was higher than that
of U. lactuca on a dry weight basis. This difference is
presumably attributable to the nature of the carbon source
in each alga; presumably C. vulgaris contained a higher
quantity of organic compounds than U. lactuca per unit
mass. Apart from the proteins and carbohydrates measured
in C. vulgaris other molecules known to be present such as
lipids and vitamins may have contributed to energy
production. While Coulombic Efficiencies (CEs) obtained
were low (1025%), they were typical for MFCs using
complex substrates, for example, wastewater (16%, Huang
and Logan, 2008) and cellulose (23%, Rezaei et al., 2008).
Presumably as we begin to better understand the factors that
affect CE, these values can be improved. CEs achieved in
MFCs fed C. vulgaris were also slightly higher that those fed

1074

Biotechnology and Bioengineering, Vol. 103, No. 6, August 15, 2009

U. lactuca. CEs decreased at COD concentrations higher


than 600 mg/L. This trend towards a reduction in CE with
organic loading has previously been observed, and is
thought to be due to increased substrate losses over a longer
batch cycle time either through methanogenesis or due to
aerobic degradation of organic matter sustained by oxygen
leakage through the cathode (Feng et al., 2008). Since low
methane compositions and VFA concentrations were
measured for MFCs using U. lactuca it appears that, in
this case, methanogenesis was not the main contributor to
the decrease in CE.
At a high organic matter load, the difference in algae
composition produced different COD removals among
MFCs. The best COD removals were obtained in reactors
using U. lactuca. This was correlated with a higher
degradation of carbohydrate in MFCs using U. lactuca than
of protein in MFCs using C. vulgaris. Moreover, particulate
carbohydrates in MFCs using U. lactuca or C. vulgaris were
more readily consumed than particulate proteins. Analysis
of microbial communities showed that bacterial selection
occurred in all reactors (MFCs and ARs) as a result of the
treatments. Microbial communities clustered according to
the type of bioprocess (MFC at OC, MFC at CC and AR)
used. This implies that the reactor conditions had an effect
in the final microbial selection suggesting that there were
possible differences in the metabolisms occurring in each
reactor.
The utility of power generation from algae using MFCs
can be evaluated in comparison to other methods of energy
production exploiting this source of organic matter. The

Table I.

Comparison of total energy production using different technologies with Chlorella vulgaris and Ulva lactuca.

Technology
Incineration
Anaerobic digestion
Hydrogen production
Oil extraction
Microbial fuel cells

Chlorella vulgaris (kW-h/kg-DW)


a

9.3
9.8b
0.4c
13.5d
2.5e

References

Ulva lactuca (kWh/kg DW)

Minowa and Sawayama (1999)


Golueke and Oswald (1959)
Kim et al. (2006)
Li et al. (2007)
This study

13.5
6.6g
n.a.
n.a.
2.0h

References
Grahame (1973)
Briand and Morand (1997)
n.a.
n.a.
This study

n.a, not available.


a,b,c,e,f,g,h
2.77 kWh/MJ, assuming 30% (for a, b, d, f, g) and 70% (for c) conversion efficiencies to convert energy from heat to power.
b
Using mixed cultures of microalgae containing Chlorella vulgaris, 1 mol CH4 981 kJ.
c
Converted to energy assuming 1 mol of H2 237  103 J (Turner et al., 2008).
d
Converted to grams of algae assuming 0.48 gbiodiesel/g-algae.
e,h
Calculated from the linear regression equation obtained in the plot of joules produced over a range of algae concentrations.

energy obtained per gram of algae with an MFC is currently


low relative to other technologies, based on assuming a 30%
efficiency from heat to power with other technologies
(Table I). For C. vulgaris the highest power output was
derived from oil extraction and methane production from
anaerobic digestion. MFCs and hydrogen production had a
low energy recovery indicating the need for technology
improvement. For U. lactuca the best option seems to be
incineration, but the sustainability of this technology is
debatable (Minowa and Sawayama, 1999). Incineration
could result in the production of hazardous ash and
emission of partially combusted products.

Conclusions
Using microalgae and macroalgae for energy generation
presents several benefits over other biomass sources: they
can grow using different substrates such as CO2 or
wastewater, have high growth rates, and require less space
for cultivation. MFCs using either microalgae or macroalgae
produced relatively high power densities compared to
the use of other substrates in this MFC configuration.
Maximum power densities were best obtained either using
fixed resistances in the circuit for a sufficiently long time for
the system to reach steady conditions, or by LSV using a
slow scan rate (0.1 mV/s). C. vulgaris gave the highest
energy generation per gram of substrate while, the
macroalgae U. lactuca was degraded more efficiently in
MFCs. The substrate composition and bioprocess had an
effect on the final COD removal obtained and the type of
microbial communities present. Carbohydrates contained in
U. lactuca were more completely degraded than proteins in
C. vulgaris. Bioelectricity production using algae in MFCs is
useful as a low temperature method of power generation,
but it needs to be further improved in order to make it
competitive with alternative energy technologies.

The authors greatly appreciate the help of David W. Jones during


laboratory measurements.

References
Aelterman P, Freguia S, Keller J, Verstraete W, Rabaey K. 2008. The anode
potential regulates bacterial activity in microbial fuel cells. Appl
Microbiol Biotechnol 78(3):409418.
American Public Health Association (APHA), American Water Works
Association AWWA, Water Environment Federation WEF. 1995.
Standard methods for the examination of water and wastewater.
Washington.
Becker EW. 2007. Micro-algae as a source of protein. Biotechnol Adv
25(2):207210.
Brecha RJ. 2008. Emission scenarios in the face of fossil-fuel peaking. Energy
Policy 36(9):34923504.
Briand X, Morand P. 1997. Anaerobic digestion of Ulva sp. 1. relationship
between Ulva composition and methanisation. J Appl Phycol 9(6):511
524.
Chen G-W, Choi S-J, Lee T-H, Lee G-Y, Cha J-H, Kim C-W. 2008.
Application of biocathode in microbial fuel cells: Cell performance
and microbial community. Appl Microbiol Biotechnol 79(3):379388.
Cheng S, Logan BE. 2007. Ammonia treatment of carbon cloth anodes to
enhance power generation of microbial fuel cells. Electrochem Commun 9(3):492496.
Cheng S, Liu H, Logan BE. 2006. Increased performance of single-chamber
microbial fuel cells using an improved cathode structure. Electrochem
Commun 8(3):489494.
Clauwaert P, van der Ha D, Verstraete W. 2008. Energy recovery from
energy rich vegetable products with microbial fuel cells. Biotechnol Lett
30(11):19471951.
de-Bashan LE, Hernandez J-P, Morey T, Bashan Y. 2004. Microalgae
growth-promoting bacteria as helpers for microalgae: A novel
approach for removing ammonium and phosphorus from municipal
wastewater. Water Res 38(2):466474.
Deublein D, Steinhauser A. 2008. Biogas from waste and renewable
resources. Weinheim: Wiley-VCH. p. 5779.
Feng Y, Wang X, Logan B, Lee H. 2008. Brewery wastewater treatment using
air-cathode microbial fuel cells. Appl Microbiol Biotechnol 78:873
880.
Gerhardt P, Murray RGE, Wood WA, Krieg NR. 1994. Methods of general
and molecular bacteriology. Washington: American Society for Microiology. p. 518519.
Golueke CG, Oswald WJ. 1959. Biological conversion of light energy to the
chemical energy of methane. Appl Microbiol 7(4):219227.
Grahame J. 1973. Assimilation efficiency of Littorina littorea (L.). J Anim
Ecol 42(2):383389.
Hansson G. 1983. Methane production from marine, green macro-algae.
Resour Conserv 8(3):185194.
Hartman E. 2008. A promising oil alternative: Algae energy. The Washington Post. p. N06.
He Z, Angenent Largus T. 2006. Application of bacterial biocathodes in
microbial fuel cells. Electroanalysis 18(1920):20092015.

Velasquez-Orta et al.: Energy From Algae Using MFCs


Biotechnology and Bioengineering

1075

Heilmann J, Logan BE. 2006. Production of electricity from proteins using a


microbial fuel cell. Water Environ Res 78:531537.
Hirano A, Hon-Nami K, Kunito S, Hada M, Ogushi Y. 1998. Temperature
effect on continuous gasification of microalgal biomass: Theoretical
yield of methanol production and its energy balance. Catal Today 45(1
4):399404.
Huang L, Logan B. 2008. Electricity generation and treatment of
paper recycling wastewater using a microbial fuel cell. Appl Microbiol
Biotechnol 80(2):349355.
Kim M-S, Baek J-S, Yun Y-S, Jun Sim S, Park S, Kim S-C. 2006. Hydrogen
production from Chlamydomonas reinhardtii biomass using a two-step
conversion process: Anaerobic conversion and photosynthetic fermentation. Int J Hydrogen Energy 31(6):812816.
Li X, Xu H, Wu Q. 2007. Large-scale biodiesel production from microalga
Chlorella protothecoides through heterotrophic cultivation in bioreactors. Biotechnol Bioeng 98(4):764771.
Logan BE, Hamelers B, Rozendal R, Schroder U, Keller J, Freguia S,
Aelterman P, Verstraete W, Rabaey K. 2006. Microbial fuel cells:
Methodology and technology. Environ Sci Technol 40(7):51815192.
Logan B, Cheng S, Watson V, Estadt G. 2007. Graphite fiber brush anodes
for increased power production in air-cathode microbial fuel cells.
Environ Sci Technol 41(9):33413346.
Minowa T, Sawayama S. 1999. A novel microalgal system for energy
production with nitrogen cycling. Fuel 78(10):12131215.
Morand P, Briand X. 1999. Anaerobic digestion of Ulva sp. 2. Study of Ulva
degradation and methanisation of liquefaction juices. J Appl Phycol
11(2):164177.
Muyzer G, Waal Ed, Uitterrlinden A. 1993. Profiling of complex microbial
populations by denaturing gradient gel electrophoresis analysis of

1076

Biotechnology and Bioengineering, Vol. 103, No. 6, August 15, 2009

polymerase chain reaction-amplified genes coding for 16rRNA. Appl


Environ Microbiol 59:695700.
Rezaei F, Richard TL, Logan BE. 2008. Enzymatic hydrolysis of cellulose
coupled with electricity generation in a microbial fuel cell. Biotechnol
Bioeng 101(6):11631169.
Riegman R, Kuipers BR, Noordeloos AAM, Witte HJ. 1993. Size-differential
control of phytoplankton and the structure of plankton communities.
Neth J Sea Res 31(3):255265.
Schenk P, Thomas-Hall S, Stephens E, Marx U, Mussgnug J, Posten C,
Kruse O, Hankamer B. 2008. Second generation biofuels: Highefficiency microalgae for biodiesel production. BioEnergy Res
1(1):2043.
Troiano RA, Wise DL, Augenstein DC, Kispert RG, Cooney CL. 1976. Fuel
gas production by anaerobic digestion of kelp. Resour Recov Conserv
2(2):171176.
Turner J, Sverdrup G, Mann MK, Maness P-C, Kroposki B, Ghirardi M,
Evans RJ, Blake D. 2008. Renewable hydrogen production. Int J Energy
Res 32(5):379407.
Ventura MR, Castanon JIR. 1998. The nutritive value of seaweed (Ulva
lactuca) for goats. Small Rumin Res 29(3):325327.
Vergara-Fernandez A, Vargas G, Alarcon N, Velasco A. 2008. Evaluation of
marine algae as a source of biogas in a two-stage anaerobic reactor
system. Biomass Bioenergy 32(4):338344.
Verseveld HWv, Roling WFM. 2008. Cluster analysis and statistical
comparison of molecular community profile data. In: Kowalchuk
GA, Bruijn FJd, Head IM, Akkermans AD, Elsas JDv, editors.
Molecular microbial ecology manual. London: Springer. p. 13731397.
Walker T, Purton S, Becker D, Collet C. 2005. Microalgae as bioreactors.
Plant Cell Rep 24(11):629641.

You might also like