100
The Pennsylvania State University The Graduate School College of Engineering CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON CATALYSTS AND POLYMER MEMBRANE LAYERS FOR MICROBIAL FUEL CELL CATHODES AND AN ANALYSIS OF POWER OVERSHOOT A Dissertation in Environmental Engineering by Valerie Jo Watson 2013 Valerie Jo Watson Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy May 2013

CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

  • Upload
    others

  • View
    5

  • Download
    0

Embed Size (px)

Citation preview

Page 1: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

The Pennsylvania State University

The Graduate School

College of Engineering

CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

CATALYSTS AND POLYMER MEMBRANE LAYERS FOR MICROBIAL FUEL CELL

CATHODES AND AN ANALYSIS OF POWER OVERSHOOT

A Dissertation in

Environmental Engineering

by

Valerie Jo Watson

2013 Valerie Jo Watson

Submitted in Partial Fulfillment of the Requirements

for the Degree of

Doctor of Philosophy

May 2013

Page 2: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

The dissertation of Valerie Watson was reviewed and approved* by the following:

Bruce E. Logan Evan Pugh and Kappe Professor of Environmental Engineering Dissertation Advisor Chair of Committee John M. Regan Associate Professor of Environmental Engineering

Fred Cannon Professor of Environmental Engineering Michael A. Hickner Assistant Professor of Materials Science and Engineering Peggy Johnson Professor of Civil Engineering Head of the Department of Civil and Environmental Engineering

*Signatures are on file in the Graduate School

Page 3: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

iii

ABSTRACT

Microbial fuel cells (MFCs) are a promising technology for treatment of wastewater

streams in combination with electricity production. Commercialization and implementation of

MFCs could eliminate the large energy consumption common in traditional wastewater treatment

and allow for the utilization of this untapped renewable energy source. Polarization curves from

microbial fuel cells (MFCs) often show unexpectedly large drops in voltage with increased

current densities, leading to a phenomenon in the power density curve referred to as “power

overshoot”. Linear sweep voltammetry (LSV, 1 mV s−1) and variable external resistances (at

fixed intervals of 20 min) over a single fed-batch cycle in an MFC both resulted in power

overshoot in power density curves due to anode potentials. Increasing the anode enrichment time

from 30 days to 100 days did not eliminate overshoot, suggesting that insufficient enrichment of

the anode biofilm was not the primary cause. Running the reactor at a fixed resistance for a full

fed-batch cycle (~1 to 2 days), however, completely eliminated the overshoot. These results show

that acclimation at low fixed resistances are needed to stabilize current generation by bacteria in

MFCs, and that even relatively slow LSV scan rates and long times between switching circuit

loads during a fed-batch cycle may produce inaccurate polarization and power density results for

these biological systems.

Membrane separators reduce oxygen flux from the cathode into the anolyte in MFCs, but

water accumulation and pH gradients between the separator and cathode reduces performance. To

avoid these problems, air cathodes were spray-coated (water-facing side) with anion exchange,

cation exchange, and neutral polymer coatings of different thicknesses to incorporate the

separator into the cathode structure. The anion exchange polymer coating resulted in greater

power density (1167 ± 135 mW m−2) than a cation exchange coating (439 ± 2 mW m−2). This

power output was similar to that produced by a Nafion-coated cathode (1114 ± 174 mW m−2), and

Page 4: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

iv

slightly lower than the uncoated cathode (1384 ± 82 mW m−2). Thicker coatings reduced oxygen

diffusion into the electrolyte and increased coulombic efficiency (CE = 56 – 64%) relative to an

uncoated cathode (29 ± 8%), but decreased power production (255–574 mW m−2).

Electrochemical characterization of the cathodes using abiotic anodes in separate reactors showed

that the cathodes with the lowest charge transfer resistance and the highest oxygen reduction

activity produced the most power in MFC tests. The results using hydrophilic cathode separator

layers revealed a tradeoff between power and CE. Cathodes coated with a thin coating of anion

exchange polymer showed the most promise for controlling oxygen transfer while minimally

affecting power production.

Platinum is commonly used as the catalyst in MFC cathodes, but platinum is an

expensive and limited resource. Activated carbon (AC) is a promising material for the

replacement of platinum catalysts because it is inexpensive and can be made from renewable

waste sources, but its catalytic performance in neutral solutions used in MFCs in not well

understood. Commercially available AC powders made from different precursor materials (coal,

peat, coconut shell, hardwood, and phenolic resin) were evaluated as oxygen reduction catalysts,

and tested as cathode catalysts in MFCs. Carbons were characterized in terms of surface

chemistry, specific surface area, and pore volume distribution, and kinetic activities were

compared to carbon black and platinum catalysts using a rotating disk electrode (RDE). Cathodes

using the coal–derived AC had the highest maximum power densities in MFCs (1620 ± 10 mW

m–2) even though this AC had only average catalytic activity, measured by reduction onset

potential (Eonset = 0.09 V), and selectivity, based on number of electrons transferred (n = 2.4).

This coal–based AC also had the lowest specific surface area (550 m2 g–1) among the ACs tested.

Peat–based AC performed similarly in MFC tests (1610 ± 100 mW m–2) but had the best catalyst

performance (Eonset = 0.17 V, n = 3.6) in RDE tests and a lower than average specific surface area

(810 m2 g–1). Hardwood based AC had the highest number of acidic surface functional groups and

Page 5: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

v

a higher specific surface area (1010 m2 g–1), but it had the poorest performance in MFC and

catalyst tests (630 ± 10 mW m–2, Eonset = –0.01V, n = 2.1). There was a strong inverse

relationship between onset potential and the quantity of strong acid (pKa < 8) functional groups,

and a larger fraction of microporosity was negatively correlated with power production in MFCs.

These results showed that surface area alone was a poor predictor of catalyst performance, and

that a high quantity of acidic surface functional groups was detrimental to oxygen reduction and

cathode performance.

Four of the commercially available AC powders (peat, coconut shell, coal, and

hardwood) were treated with ammonia gas at 700 °C in order to improve their performance as

oxygen reduction catalysts. Ammonia treatment resulted in a decrease in oxygen (by 29 – 58%)

and an increase in nitrogen content (total abundance up to 1.8 atomic %) on the carbon surfaces,

which also resulted in an increase in the basicity of the bituminous, peat, and hardwood ACs. The

kinetic activity and selectivity of ammonia–treated carbons were evaluated using a rotating ring

disk electrode (RRDE), and compared to untreated ACs and platinum. All of the ammonia–

treated ACs exhibited better catalytic performance than their untreated precursors, with the

bituminous (treated, Eonset = 0.12 V, n = 3.9; untreated, Eonset = 0.08 V, n = 3.6) and hardwood

(treated, Eonset = 0.03 V, n = 3.3; untreated, Eonset = –0.04 V, n =3.0) based samples showing the

most improvement. These ACs were tested in MFC cathodes made by sandwiching the AC

catalyst and polytetrafluoroethylene (PTFE) binder mixture between two current collectors, one

coated with polydimethylsiloxane (PDMS) diffusion layers on the air–side, and the second one on

the solution–side used to improve power. Cathodes made from the ammonia–treated coal-based

AC had the one of the highest maximum power densities (2450 ± 40 mW m–2). Even though the

ammonia–treated peat–based AC had the greatest ORR activity in RRDE testing, the untreated

sample had higher power production in the MFC tests (2360 ± 230 mW m–2). The treated coconut

and hardwood derived ACs outperformed the untreated precursor ACs in both electrochemical

Page 6: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

vi

and MFC testing. These results show that reduction in oxygen abundance and increase in nitrogen

functionalities on the surface of ACs can increase the catalytic performance for oxygen reduction

in neutral media.

Page 7: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

vii

TABLE OF CONTENTS

List of Figures .......................................................................................................................... x

List of Tables ........................................................................................................................... xii

Acknowledgements .................................................................................................................. xiii

Chapter 1 Introduction ............................................................................................................. 1

1.1 Energy Demand and the Environment ....................................................................... 1 1.2 Microbial Fuel Cells ................................................................................................... 2

1.2.1 Oxygen Reduction at the MFC Cathode ......................................................... 2 1.2.2 Activated Carbon Catalysts ............................................................................. 3 1.2.3 Membranes and Separators ............................................................................. 5

1.3 Analysis Methods ....................................................................................................... 7 1.3.1 MFC Power Density and Overshoot ............................................................... 7 1.3.2 Catalyst Performance ...................................................................................... 8

1.4 Conclusions ................................................................................................................ 9 1.5 Objectives and Clarification of Contributions ........................................................... 10 1.6 References .................................................................................................................. 11

Chapter 2 Analysis of polarization methods for elimination of power overshoot in microbial fuel cells ........................................................................................................... 15

Abstract ............................................................................................................................ 15 2.1 Introduction ................................................................................................................ 16 2.2 Experimental procedures ............................................................................................ 17

2.2.1 MFC reactor construction and operation ......................................................... 17 2.2.2 Analysis ........................................................................................................... 18

2.3 Results ........................................................................................................................ 19 2.3.1. Polarization by varied resistance, single-cycle method .................................. 19 2.3.2. Polarization by varied resistance, multiple-cycle method .............................. 21 2.3.3 Polarization by LSV ........................................................................................ 21

2.4 Discussion .................................................................................................................. 22 2.5 Acknowledgements .................................................................................................... 24 2.6 References .................................................................................................................. 24

Chapter 3 Polymer coatings as separator layers for microbial fuel cell cathodes .................... 25

Abstract ............................................................................................................................ 25 3.1 Introduction ................................................................................................................ 26 3.2 Materials and methods ............................................................................................... 28

3.2.1 Polymers .......................................................................................................... 28 3.2.2 Cathode construction ....................................................................................... 29 3.2.3 MFC reactor construction and operation ......................................................... 30 3.2.4 Analysis ........................................................................................................... 31

Page 8: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

viii

3.3 Results ........................................................................................................................ 32 3.3.1 MFC performance ........................................................................................... 32 3.3.2 Electrochemical performance .......................................................................... 35 3.3.3 Oxygen diffusion and biofilm growth ............................................................. 37

3.4 Discussion .................................................................................................................. 38 3.5 Acknowledgements .................................................................................................... 41 3.6 References .................................................................................................................. 41

Chapter 4 Influence of Chemical and Physical Properties of Activated Carbon Powders on Oxygen Reduction Catalysis and Performance in Microbial Fuel Cells ..................... 43

Abstract ............................................................................................................................ 43 4.1 Introduction ................................................................................................................ 44 4.2 Materials and Methods ............................................................................................... 46

4.2.1 Catalyst Materials ............................................................................................ 46 4.2.2 Physical and Chemical Analyses ..................................................................... 47 4.2.3 RDE Analysis .................................................................................................. 48 4.2.4 MFC Experiments ........................................................................................... 49

4.3 Results and Discussion ............................................................................................... 50 4.3.1 MFC Performance ........................................................................................... 50 4.3.2 Catalyst Activity and Selectivity ..................................................................... 52 4.3.3 Effect of Oxygen Functional Groups on ORR Catalysis................................. 54 4.3.4 Effect of Microporosity on Power Production ................................................ 56 4.3.5 Functional Group Analysis Using XPS ........................................................... 57 4.3.6 Implications of AC properties for MFC performance ..................................... 59

4.4 Acknowledgements .................................................................................................... 59 4.5 References .................................................................................................................. 60

Chapter 5 Improvement of Oxygen Reduction Catalysis in Neutral Solutions using Ammonia Treated Activated Carbons and Performance in Microbial Fuel Cells ........... 62

Abstract ............................................................................................................................ 62 5.1 Introduction ................................................................................................................ 63 5.2 Materials and Methods ............................................................................................... 65

5.2.1 Activated Carbons and Ammonia Treatment .................................................. 65 5.2.2 Chemical Surface Analysis ............................................................................. 65 5.2.3 Rotating Ring-Disk Electrochemical Analysis ................................................ 66 5.2.4 MFC Experiments ........................................................................................... 67

5.3 Results and Discussion ............................................................................................... 69 5.3.1 MFC performance ........................................................................................... 69 5.3.2 Catalyst Activity and Selectivity ..................................................................... 71 5.3.3 Effect of Surface Chemistry ............................................................................ 73

5.4 Conclusion ................................................................................................................. 76 5.5 Acknowledgements .................................................................................................... 77 5.6 References .................................................................................................................. 77

Chapter 6 Conclusions and Future Work ................................................................................. 79

Appendix A Supplemental Information to Chapter 4 ..................................................... 81

Page 9: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

ix

Appendix B Supplemental Information to Chapter 5 ...................................................... 86

Page 10: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

x

LIST OF FIGURES

Figure 2-1. (A) Power density and (B) polarization curves for single-cycle (20 min) resistance changes at 30 days (◊) and 100 days (Δ) and multiple-cycle resistance changes (). ...................................................................................................................... 20

Figure 2-2. Electrode (A=anode and C=cathode) potential measurements (vs. Ag/AgCl) during cell polarization for single-cycle (20 min) resistance changes at 30 days (◊) and 100 days (Δ) and multiple-cycle resistance changes (). .......................................... 20

Figure 2-3. (A) Power density and (B) polarization curves from LSV at 1 mV s−1 for three consecutive scans. ............................................................................................................ 22

Figure 3-1. (A) Power density and (B) polarization curves for polymer-coated cathodes. ..... 34

Figure 3-2. Electrode potential measurements (vs. Ag/AgCl) during cell polarization. .......... 34

Figure 3-3. Coulombic efficiencies for cycles run at 1000 Ω. ................................................. 35

Figure 3-4. EIS of coated and uncoated cathodes at 0.1V (vs. Ag/AgCl) (200mM PBS). ...... 36

Figure 3-5. LSV of coated and uncoated cathodes (100mM PBS). ......................................... 37

Figure 3-6. Optical images of biofilm growth on cathodes (100 days). .................................. 38

Figure 3-7. Inverse relationship between Rct and power density. ............................................ 39

Figure 3-8. Inverse relationship between CE and power density. ............................................ 40

Figure 4-1. A) Power density production and B) electrode potentials from polarization of MFCs using AC cathodes compared to Pt/C (100 mM Phosphate Buffer; Open symbols represent cathode potentials, closed symbols are anode potentials). ................. 51

Figure 4-2. A) LSV current response (per disk area) of the AC catalyst at the disk electrode compared to Pt/C and carbon black (100 mM Phosphate Buffer, 2100rpm) B) Average number of electrons (n) transferred (estimated by Koutecky-Levich RDE analysis) during oxygen reduction. ......................................................................... 53

Figure 4-3. Acidic/Oxygen functional groups determined by potentiometric titration. A) Bituminous and peat based activated carbon samples have similar functional groups. B) Other activated carbons have a larger variety of acidic groups. ................................. 55

Figure 4-4. The onset potential of the oxygen reduction reaction is inversely related to the amount of strong acid functional groups present on the activated carbons tested. .......... 55

Figure 4-5. Maximum power density (per m2 projected cathode surface) of the MFCs using the activated carbon cathodes is inversely related to the A) surface area (without W1 pvalue=0.0006) and B) micropore volume (without W1 pvalue=0.0011) of the powdered carbons with the exception of sample W1. ................. 57

Page 11: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

xi

Figure 4-6. A) Maximum power density (normalized to cathode surface area) of MFCs with activated carbon cathodes are not directly related to oxygen content of activated carbon powders determined by XPS. B) Chemical state of oxygen detected by XPS varies for activated carbon powders tested (e.g. increased signal in the adsorbed O2/H2O region for sample B1.) ........................................................................................ 58

Figure 5-1. A) Power density production and B) electrode potential during cell polarization of MFCs using AC cathodes compared to Pt/C. (100 mM Phosphate Buffer; open symbols indicate cathode potentials, closed symbols anode potentials.) .... 70

Figure 5-2. A) H2O2 detection based on oxidation current at the Pt ring during oxygen reduction at the catalyst on the disk electrode. B) Oxygen reduction current response during LSV of AC catalysts at the disk electrode compared to Pt/C (100 mM Phosphate Buffer, 2100 rpm). C) Average number of electrons transferred (measured by RRDE analysis) during oxygen reduction. ................................................ 72

Figure 5-3. Proton binding isotherms for treated and untreated A) bituminous, B) peat, C) coconut shell, and D) hardwood based activated carbons. ............................................... 74

Figure 5-4. N1s peaks on ammonia treated ACs from XPS show presence of nitrogen groups on the surface of the treated AC. .......................................................................... 75

Figure 5-5. Atomic % of oxygen and nitrogen on the surface of treated and untreated AC catalysts measured using XPS and the relationship to onset potential of the ORR measured with RRDE. ...................................................................................................... 76

Figure A-1: Example of RDE LSV data for bituminous coal based sample (B1) collected at rotation rates from 100 – 2100 RPM. ........................................................................... 81

Figure A-2: Example of K-L analysis for bituminous coal based sample (B1) where the slope of the line is used to calculate n and the y-intercept is the inverse ik. ..................... 82

Figure A-3: Potentiometric titration curves showing protons bound (positive Q) or released (negative Q). The isotherm data was then further analyzed using SAIEUS software to quantify the type (pKa) and quantity of acidic functional groups. ................ 83

Figure A-4: Inverse correlation between quantity of strong acid functional groups on the AC catalyst powder and the power production in an MFC using the AC cathode (a) with and (b) without the inclusion of the bituminous coal based sample (B1) ................ 84

Figure A-5: Cumulative pore volume distribution of AC catalyst powders measured by argon adsorption and DFT analysis .................................................................................. 85

Figure B-1. (A) Power production increases with the additional stainless steel current collector. (B) Cathodes (open symbols) with the additional current collector operate at a higher potential. Anodes (filled symbols) perform the same with both cathode configurations. (100 mM Phosphate buffer, Data of MFC cathodes without the extra current collector is from Chapter 4.) ................................................................................ 86

Page 12: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

xii

LIST OF TABLES

Table 3-1. Properties of polymers and cathode coatings. ........................................................ 29

Table 3-2. Oxygen flux, combined solution and membrane resistance, charge transfer resistance, and maximum power density of cathodes. ..................................................... 33

Page 13: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

xiii

ACKNOWLEDGEMENTS

I would like to whole heartedly thank everyone who made the last several years possible,

especially Dr. Bruce Logan for supporting me through this research journey. I thank him for

introducing me to the incredibly interesting concept of microbial based electricity production

from wastewater treatment. I am truly honored to have been able to work with him. I would also

like to thank the other members of my committee, Dr. Jay Regan, Dr. Fred Cannon, and Dr. Mike

Hickner for their time and support in this research effort.

I would like to thank all of the members of the Logan Lab throughout my time here,

especially Dr. Rachel Wagner, who was always there to encourage me to keep moving forward. I

also appreciate the assistance of Tim Byrne and Dr. Cesar Neito Delgado with their knowledge of

activated carbon testing.

A special thank you to the National Science Foundation (NSF) Graduate Research

Fellowship program and the King Abdullah University of Science and Technology (KAUST) for

funding this work.

Most importantly I would like to thank my family. My parents, John and Carol Prothero,

and sister, Jacqueline Bealla, for never doubting that I could do this. My in-laws, Donald and

Diana Watson, for helping feed and take care of my children when I couldn’t. My husband,

Steven Watson, for putting up with me and supporting me and taking care of the kids and dealing

with a very messy house and on and on and on. And finally, to my children, Alina and Kira

Watson, for believing I could do anything and loving me even if I couldn’t. I started this journey

wanting to be an inspiration to you, but instead, you became my inspiration. Thank you all, (and

all those that I didn’t have time and space to mention, but I am thinking about gratefully right

now) for your incredible support. Without all of you (and God), none of this would have been

possible.

Page 14: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

1

Chapter 1

Introduction

1.1 Energy Demand and the Environment

According to the 2011 United States Energy Information Administration worldwide

report, world energy consumption is projected to grow by 53% from 2008 to 2035. In 2009, the

US was the largest total energy consumer (94.5 quadrillion Btu) and accounted for about 20% of

the world’s energy consumption, although it only contained 5% of the world’s population. Most

worldwide energy consumption is currently derived from fossil fuels, but there are many issues

associated with their use. Fossil fuels are non-renewable resources and they are mainly sourced

from countries with volatile political and economic environments. In 2011, the US was the

world’s largest importer and consumer of petroleum. Burning fossil fuels also produces the main

source of carbon dioxide emissions. Increased anthropogenic CO2 emissions have been linked to

climate change which endangers global ecosystems. The US is the second largest producer of the

world’s CO2 emissions, with China being the number one producer. However most of the

projected increase in global CO2 emissions is attributed to developing countries, with global CO2

emissions projected to increase by 43% from 2008 to 2035[1, 2].

Wastewater treatment consumes a substantial amount of energy. It is estimated that

treatment of wastewater containing mostly organic compounds consumes about 15 GW of power,

which is equivalent to about 3% of US electricity production. It is also estimated that domestic,

industrial, and agricultural (animal) wastewater combined contains about 17 GW of power, which

is mostly lost in current treatment processes [3]. Developing low cost, low energy consumption

Page 15: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

2

methods to treat these waste sources, while utilizing or capturing the energy contained in them,

would be beneficial to both environmental and energy demand issues.

1.2 Microbial Fuel Cells

Microbial fuel cells (MFCs) are a promising technology for treatment of wastewater

streams in combination with electricity production [3, 4]. In a typical design, bacteria oxidize

organic waste in anaerobic conditions and transfer electrons to an anode. Electrons are conducted

through a circuit to the passively air-fed cathode consisting of a porous carbon structure with an

oxygen reduction catalyst, where oxygen is ideally reduced to water through a 4e– transfer

pathway, producing electrical current [4]. Commercialization and implementation of MFCs could

eliminate the large energy consumption common in traditional wastewater treatment (mostly

through the elimination of aeration of the waste stream) and allow for the utilization of this

untapped renewable energy source. Although improvements in MFC design and performance has

progressed rapidly over the past several years, more technological advances are needed for MFCs

to become commercially viable. Specifically, improvement in cathode design and low-cost

materials are needed in order to increase power production, decrease material costs, and improve

long–term stability [5].

1.2.1 Oxygen Reduction at the MFC Cathode

Oxygen reduction air cathodes are commonly used in MFCs since oxygen is readily

available, has a positive redox potential, and is environmentally innocuous. Unfortunately, the

oxygen reduction reaction (ORR) is kinetically slow in common MFC conditions (neutral pH and

ambient temperatures) [6]. Platinum particles loaded onto high surface area carbon black (e.g.

Page 16: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

3

Vulcan XC-72) are commonly used to catalyze the ORR in MFCs and other fuel cells because of

its well-known high activity and performance, but platinum is expensive, finite in abundance, and

prone to inhibition by contaminants found in the waste streams to be treated [7, 8].

Depending on the catalyst, the ORR proceeds through either a 4e– pathway producing

water or hydroxide [9] or 2e– pathway producing hydrogen peroxides as intermediates [10, 11].

Currently cathode material costs account for 47-75% of MFC capital costs [12]. Several other

catalyst materials have been considered for use in MFCs, including other metal compounds such

as cobalt and iron tetramehoxyphenylporphyrin (TMPP) or phthalocyanine (Pc) [7] and

manganese oxides [13, 14]. Recently there have been promising studies using activated carbon

(AC) powder based air–cathodes [15-18].

1.2.2 Activated Carbon Catalysts

Activated carbon (AC) is relatively inexpensive and can be made from a variety of

materials that are readily available, including several waste sources [19-21]. Coal (Bituminous

and Lignite), wood, phenol resin, coconut shells, and silk fibroin have been used as AC

precursors. Utilization of different precursors as well as different activation processes (e.g. steam

vs. chemical activation) can lead to differences in the surface functional groups, surface area, and

pore structure (micro-, meso-, and macropore development) of the AC [19-21]. For instance, silk

fibroin was used as an AC precursor because the material contained no metal elements and

eighteen types of amino acids that contain nitrogen atoms [20].

The use of AC as an ORR catalyst has been evaluated by several researchers, but most of

the work has been done under conditions that are not applicable to MFCs, such as using acidic

and alkaline solutions commonly used in other types of fuel cells. In these other solutions,

researchers have found impacts of specific surface area, pore structure, and precursor material on

Page 17: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

4

the AC catalyst activity, but the results are not always in agreement on which characteristics have

the most influence on the ORR [20, 22]. A silk based AC was compared to a phenolic resin–

derived AC, and the materials showed different ORR activity that did not seem to correlate with

specific surface area [20]. X-ray photoelectron spectroscopy (XPS) was used to analyze the types

of nitrogen atoms present in AC materials (e.g. pyridine-like, pyrrole-like, and quaternary) and it

was found that the presence of quaternary nitrogen atoms improved the performance of the

catalyst for oxygen reduction. Unfortunately no analysis of pore size distribution was mentioned

in this study.

Iron and nitrogen loaded ACs with different pore size distributions were tested as oxygen

reduction catalysts [22]. The pore sizes were varied for the same carbons by heating the materials

at 500°C in air for 20 to 180 minutes. The researchers found that surface area and pore size varied

with the “burn-off” time and that the pore size had an important effect on catalytic activity, with

pore sizes around 13Å showing the best performance. The relationship between specific surface

area and catalyst activity was not linear. Again, nitrogen content of the ACs was also found to

have an effect on the catalyst activity for oxygen reduction. In a similar study, researchers found

that only pore sizes greater than 15Å can be utilized for oxygen reduction in alkaline media [10].

ACs have also been tested as catalysts in MFC cathodes, but most often they are used as a

high surface area catalyst support for platinum or other metal catalysts [23]. In recent studies, AC

powder based cathodes have been successful in attaining power densities similar to those

achieved with commonly used platinum cathodes. An MFC with an AC cathode made using a

proprietary process, which contained a polytetrafluoroethylene (PTFE) binder and a nickel

current collector, produced 1220 mW m–2, compared to 1060 mW m–2 using a cathode with a Pt

catalyst and a Nafion binder [15]. An MFC incorporating a similar AC cathode structure that had

a polydimethylsiloxane (PDMS) coated cloth diffusion layer reached 1255-1310 mW m–2

compared to 1295 mW m–2 with a standard Pt/C cathode [16]. However, very little testing has

Page 18: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

5

been done to compare the activity of ACs obtained from different precursor materials and with

different pore structures in MFCs. A cathode that was formed by rolling (rather than pressing) the

AC and PTFE catalyst layer onto a stainless steel mesh current collector produced 1086 mW m–2

in an MFC with a high surface area (1701 m2 g–1) AC, and 1355 mW m–2 with a lower surface

area (576 m2 g–1) AC powder. Power production using a standard Pt/C cathode as a baseline

reference was not reported [18]. The higher power production by the lower surface area AC

cathode was attributed to a more uniform distribution of microporosity. However, only two ACs

were compared and there was no analysis of AC surface chemistry, which could have affected the

ORR.

Several studies have shown that the number of nitrogen functional groups on carbon

surfaces can be increased by treatment with ammonia gas at elevated temperatures [11, 24-26].

During the process of incorporating the nitrogen into the carbon structure, there is a

corresponding reduction in acidic oxygen groups as the oxygen atoms are desorbed from the

carbon surface as CO/CO2. This rearrangement of surface functional groups results in an increase

in the basic properties of the carbon surface at the expense of acidic properties [24, 27-31].

Nitrogen incorporation on carbon surfaces has been shown to increase the catalytic activity and

selectivity for oxygen reduction through a four electron pathway in both acidic and alkaline

environments, but it has not been examined under neutral pH conditions [11, 32, 33].

1.2.3 Membranes and Separators

Some MFCs include an ion exchange membrane in the electrolyte compartment between

the anode and cathode. However, membranes have been shown to negatively impact the power

production of the MFC by increasing the internal resistance of the cell and inducing pH gradients

during cell operation [34, 35]. Cations such as Na+, K+, and NH4+ are preferentially transferred

Page 19: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

6

through cation exchange membranes (CEMs) due to their higher concentrations in water, rather

than protons present at lower concentrations, in order to maintain charge balance. This results in a

decrease in performance due to pH changes [6, 34]. Anion exchange membranes (AEMs)

outperform CEMs and other types of membranes in MFCs and microbial electrolysis cells

(MECs) mostly due to lower internal resistances that result from lower charge transport resistance

[35-38]. Charge balance can be facilitated by transfer of buffer anions (such as phosphate) when

using an AEM [35]. However, both AEMs and CEMs negatively impact MFC performance due

to the formation of pH gradients at the electrodes [39].

MFC power production can be improved by removing the membrane from the system

[40] and reducing the electrode spacing to decrease ohmic losses. When the electrodes become

closely spaced, however, a separator is needed to prevent short circuiting and also to reduce

oxygen diffusion into the anode chamber which can adversely affect power production [41].

Oxygen diffusion into the anode chamber negatively affects MFC performance by serving as an

alternative electron acceptor for the facultative bacteria at the anode. If the bacteria use the

oxygen as the terminal electron acceptor instead of the anode current collector, the coulombic

efficiency (CE) will decrease, the anode potential will become more positive, and the current

density will decrease [42]. Cloth separators have been used to decrease oxygen diffusion into the

anolyte, but over time the cloth became completely degraded by the bacteria in the reactor [43,

44]. Positioning a glass fiber separator next to the cathode in an MFC with 2 cm electrode spacing

has been shown to increase CE to 80% compared to 30% without a separator [43]. However,

power production with the separator decreased from 896 mW m−2 to 791 mW m−2 as a result of

decreased cathode potential and increased ohmic resistance. To improve power production, the

electrode spacing was decreased using the separator which prevented short-circuiting. With the

decreased electrode spacing, the power density increased to 1195 mW m−2 while maintaining CE

at 80%. In the same study, growth of biofilm on the cathode was also found to improve CE over

Page 20: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

7

time due to a decrease in oxygen diffusion into the electrolyte from the air cathode, but the

biofilm also hindered proton migration to the cathode and limited power production [43].

Currently more work needs to be done in order to develop a material that can be used to separate

the anode and cathode electrodes/chambers without decreasing power production.

1.3 Analysis Methods

1.3.1 MFC Power Density and Overshoot

Power production is one of the main measures of MFC performance, but estimates of the

amount of power that can be produced in an MFC are a function of the technique used to obtain

polarization data. Linear sweep voltammetry (LSV) is commonly used in MFC studies to obtain

polarization data, but high scan rates can overestimate power production [45]. An alternate

approach is to vary the circuit resistance at fixed time intervals, ranging from 10 s to 24 h, or even

running each resistance for a full cycle [46, 47]. A common problem often encountered when

evaluating polarization curves is “power overshoot” [45, 48-50]. Power overshoot refers to the

response of the system at high current densities (past maximum power) in a power density curve

where the cell voltage and current drop very quickly resulting in a doubling back of the power

density curve, producing lower power than previously measured at lower current densities [50].

One hypothesis on the cause of this power overshoot is that as the current resistance is decreased

the bacteria on the anode are unable to produce sufficient current at lower voltages [50]. Accurate

methods are needed to ensure that power densities reported by different researchers reflect the

performance of the MFC in stable conditions.

Page 21: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

8

1.3.2 Catalyst Performance

In order to evaluate catalyst candidates for ORR activity, many studies have employed

the use of a rotating disk electrode (RDE) in order to isolate mass transfer effects and focus on the

kinetics of the reaction [21, 51, 52]. Catalysts are applied in a uniform thin Nafion bound film to

a glassy carbon disk electrode and evaluated by comparing current production during a scan of

the RDE potential in the range of the ORR. Reduction current free from mass transfer effects (ik)

can be determined using the Koutecky-Levich equation

1𝑖

= 1𝑖𝑘

+ 1𝑖𝑑

= 1𝑛𝐹𝐴𝑘𝐶𝑂2

− 1

0.62𝑛𝐹𝐴𝐷𝑂22 3⁄ 𝜐−1 6⁄ 𝐶𝑂2𝜔

1 2⁄ (1-1)

where i is the measured current, ik the kinetic current, id the diffusion-limiting current, F

Faraday’s constant, A the projected surface area of the disk electrode, k the rate constant, CO2 the

concentration of oxygen in solution, DO2 the diffusion coefficient of oxygen, υ the kinematic

viscosity, and ω the rotation rate of the electrode [13]. By plotting -1/i vs. ω-1/2, the number of

electrons transferred during oxygen reduction can be determined by using the slope of the linear

regression and ik can be determined by the y-intercept [13, 21, 53]. RDE analyses have been used

to study the ORR catalysis of many materials, including AC powders [18] and other materials

such as carbon–supported magnesium oxide nanoparticles [13, 14], and FeTMPP and FePc [7],

where the RDE results were well correlated with MFC performance.

Rotating Ring Disk Electrode (RRDE) experiments can be used to directly determine the

fraction of peroxide produced at the catalyst during oxygen reduction while simultaneously

evaluating the overall ORR activity of the catalyst [51]. While the potential of the glassy carbon

disk electrode is scanned, the platinum ring electrode is held at a potential where the oxidation of

peroxide is diffusion limited (1.2 V vs. RHE). The amount of current produced at the ring

electrode indicates the amount of peroxide formed at the catalyst on the disk electrode. The 4e–

Page 22: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

9

pathway is ideal for fuel cell cathodes because it can induce higher currents for increased power

production. The number of electrons transferred measured using a rotating ring-disk electrode

(RRDE) can be influenced by the proportion of the reactions taking place including other

reactions that may occur [51], such as:

Reduction of H2O2: H2O2 + 2H+ + 2e– → 2H2O

Decomposition of H2O2: 2H2O2 → 2H2O + O2

This RRDE method was used in silk fibroin AC testing to simultaneously measure activity and

selectivity [20]. The researchers determined that 25% of the oxygen reacted followed the 2e–

pathway to H2O2 production verifying the overall 3.5e– transfer per mole of O2. The average

number of electrons transferred (n) in the ORR at the disk electrode can be calculated based on

the amount of H2O2 detected using [54]

𝑛 = 4𝑖𝑑𝑖𝑠𝑘𝑖𝑑𝑖𝑠𝑘+𝑖𝑟𝑖𝑛𝑔 𝑁⁄

(1–4)

where idisk is reduction current at the disk, iring the oxidation current at the ring, and N is the

collection efficiency of the RRDE.

1.4 Conclusions

MFC technology is a promising innovation for treating wastewater streams while

simultaneously salvaging energy that is typically wasted. In order to ensure the

commercialization of these systems, the capital costs of these systems must be reduced while

increasing performance. Many advances have been made to reduce the cost of the materials in a

short period of time. Still, more work needs to be done to find low-cost cathode materials in order

to decrease the cost and increase the power production of MFCs.

Page 23: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

10

1.5 Objectives and Clarification of Contributions

The objective of this research was to improve the understanding and performance of

MFCs, with a special focus on cathode materials and performance. The first Chapter provided an

overview into the need for alternative technologies for wastewater treatment and energy

production, as well as an introduction to MFC technology.

A study of the use of different methods to obtain power density analysis in MFCs and

their effect on the observation of power overshoot was described in Chapter 2. I was the primary

author and contributor on this research, as I performed all the experiments and analysis. The

coauthor (Dr. Bruce Logan) contributed to the editing on the manuscript.

A method of isolating the cathode from bacterial growth was presented in Chapter 3. I

was the primary author of this paper. My co-authors selected the polymer materials to test based

on our objectives and the MFC environment and Dr. Tomonori Saito prepared the polymer

solutions and provided the material characteristics listed in Table 3-1. I constructed the cathodes,

coated them with the polymer materials, and performed the electrochemical and MFC analysis.

All of the coauthors contributed to the editing of the paper.

The comparison of several different types of powdered AC and their catalytic and MFC

performance was presented in Chapter 4. I realized there was a need to further characterize the

AC used in MFC electrodes to better understand their performance since most research about

ACs and AC electrodes are done in acidic and alkaline environments and there was little research

comparing different types of activated carbon as the ORR catalyst in MFC cathodes. I therefore

obtained commercially produced ACs made from different precursor materials and that had

different surface areas in order to study the relationship between the chemical and physical

characteristics of the ACs and their performance as ORR catalysts in neutral solutions. I

performed the experimentation and analysis with the exception of running the potentiometric

Page 24: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

11

titrations which were done by Dr. Cesar Nieto Delgado, and the XPS operation which was done

by Vince Bojan. I was the primary author and the co-author (Dr. Bruce Logan) contributed to the

editing of the draft manuscript.

Since the results reported in Chapter 4 indicated that that strong acid oxygen groups were

detrimental to ORR catalysis in ACs, I decided to study the effect of increasing the amount of

nitrogen groups. In Chapter 5, the comparison in performance of ammonia treated activated

carbons to the non-treated samples was presented. I was the primary author and prepared the

treated activated carbons. I performed all of the electrochemical and MFC performance

experimentation and analysis. The potentiometric titrations were run by Dr. Cesar Nieto

Delgado, and the XPS was performed by Vince Bojan. The co-author (Dr. Bruce Logan)

contributed to the editing of the draft manuscript.

1.6 References

1. Administration, U.S.E.I. International Energy Statistics. 2013 [cited 2013 January 30, 2013]; Available from: http://www.eia.gov/countries/data.cfm.

2. Administration, U.S.E.I., International Energy Outlook 2011, D.o. Energy, Editor. 2011. 3. Logan, B. and K. Rabaey, Conversion of Wastes into Bioelectricity and Chemicals by

Using Microbial Electrochemical Technologies. Science, 2012. 337: p. 686-690. 4. Logan, B.E., Microbial Fuel Cells. 2008, New York: John Wiley & Sons. 5. Rismani-Yazdi, H., et al., Cathodic limitations in microbial fuel cells: An overview.

Journal of Power Sources, 2008. 180: p. 683-694. 6. Zhao, F., et al., Challenges and constraints of using oxygen cathodes in microbial fuel

cells. Environmental Science and Technology, 2006. 40(17): p. 5193-5199. 7. Birry, L., et al., Application of iron-based cathode catalysts in a microbial fuel cell.

Electrochimica Acta, 2011. 56: p. 1505-1511. 8. Harnisch, F., S. Wirth, and U. Schröder, Effects of substrate and metabolite crossover on

the cathodic oxygen reduction reaction in microbial fuel cells: Platinum vs. iron(II) phthalocyanine based electrodes. Electrochemistry Communications, 2009. 11(11): p. 2253-2256.

9. Popat, S., et al., Importance of OH- Transport from Cathodes in Microbial Fuel Cells. ChemSusChem, 2012. 5: p. 1071-1079.

10. Qu, D., Investigation of oxygen reduction on activated carbon electrodes in alkaline solution. Carbon, 2007. 45: p. 1296-1301.

Page 25: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

12

11. Zhong, R.-S., et al., Effect of carbon nanofiber surface functional groups on oxygen reduction in alkaline solution. Journal of Power Sources, 2013. 225: p. 192-199.

12. Rozendal, R.A., et al., Towards practical implementation of bioelectrochemical wastewater treatment. Trends in Biotechnology, 2008. 26(8): p. 450-459.

13. Chen, Y., et al., Stainless steel mesh coated with MnO2/carbon nanotube and polymethlyphenyl siloxane as low-cost and high-performance microbial fuel cell cathode materials. Journal of Power Sources, 2012. 201: p. 136-141.

14. Roche, I. and K. Scott, Carbon-supported manganese oxide nanoparticles as electrocatalysts for oxygen reduction reaction (orr) in neutral solution. Journal of Applied Electrochemistry, 2009. 39: p. 197-204.

15. Zhang, F., et al., Power generation using an activated carbon and metal mesh cathode in a microbial fuel cell. Electrochemistry Communications, 2009. 11(11): p. 2177-2179.

16. Wei, B., et al., Development and evaluation of carbon and binder loading in low-cost activated carbon cathodes for air-cathode microbial fuel cells. RSC Advances, 2012. 2: p. 12751-12758.

17. Dong, H., et al., A novel structure of scalable air-cathode without Nafion and Pt by rolling activated carbon and PTFE as catalyst layer in microbial fuel cells. Water Res, 2012. 46: p. 5777-5787.

18. Dong, H., H. Yu, and X. Wang, Catalysis kinetics and porous analysis of rolling activated carbon-PTFE air-cathode in microbial fuel cells. Environ Sci Technol, 2012. 46: p. 13009-13015.

19. Matsis, V.M. and H.P. Grigoropoulou, Kinetics and equilibrium of dissolved oxygen adsorption on activated carbon. Chemical Engineering Science, 2008. 63: p. 609-621.

20. Iwazaki, T., et al., High oxgen-reduction activity of silk-derived activated carbon. Electrochemistry Communications, 2009. 11: p. 376-378.

21. Maruyama, J. and I. Abe, Carbonized hemoglobin functioning as a cathode catalyst for polymer electrolyte fuel cells. Chemistry of Materials, 2006. 18: p. 1303-1311.

22. Yang, R., T.R. Dahn, and J.R. Dahn, Fe-N-C oxygen reduction catalysts supported on "burned-off" activated carbon. Journal of The Electrochemical Society, 2009. 156(4): p. B493-B498.

23. Aelterman, P., et al., Microbial fuel cells operated with iron chelated air chathodes. Electrochimica Acta, 2009. 54: p. 5754-5760.

24. Chen, W., F.S. Cannon, and J.R. Rangel-Mendez, Ammonia-tailoring of GAC to enhance perchlorate removal. I: Characterizationof NH3 thermally tailored GACs. Carbon, 2005. 43: p. 573-580.

25. Arrigo, R., et al., Tuning the acid/base properties of nanocarbons by functionalization via ammination Journal of the American Chemical Society, 2010. 132: p. 9616-9630.

26. Shafeeyan, M.S., et al., A review on surface modification of activated carbon for carbon dioxide adsorption. Journal of Analytical and Applied Pyrolysis, 2010. 89: p. 143-151.

27. Mangun, C.L., et al., Surface chemistry, pore sizes and adsorption properties of activated carbon fibers and precursors treated with ammonia. Carbon, 2001. 39: p. 1809-1820.

28. Biniak, S., et al., The characterization of activated carbons with oxygen and nitrogen surface groups. Carbon, 1997. 35(12): p. 1799-1810.

29. Szymanski, G.S., et al., The effect of the gradual thermal decomposition of surface oxygen species on the chemical and catalytic properties of oxidized activated carbon. Carbon, 2002. 40: p. 2627-2639.

30. Shen, W., Z. Li, and Y. Liu, Surface chemical functional groups modification of porous carbons. Recent Patents on Chemical Engineering, 2008. 1: p. 27-40.

Page 26: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

13

31. Hulicova-Jurcakova, et al., Combined effect of nitrogen- and oxygen-containing functional groups of microporous activated carbon on its electrochemical performance in supercapacitors. Advanced Functional Materials, 2009. 19: p. 438-447.

32. Kruusenberg, I., et al., Non-platinum cathode catalysts fo alkaline membrane fuel cells. International Journal of Hydrogen Energy, 2012. 37: p. 4406-4412.

33. Nallathambi, V., et al., Development of High Performance Carbon Composite Catalyst for Oxygen Reduction Reaction in PEM Proton Exchange Membrane Fuel Cells. Journal of Power Sources, 2008. 183: p. 34-42.

34. Rozendal, R.A., H.V.V. Hamelers, and C.J.N. Guisman, Effects of membrane cation transport on pH and microbial fuel cell performance. Environmental Science & Technology, 2006. 40(17): p. 5206-5211.

35. Kim, J.R., et al., Power generation using different cation, anion and ultrafiltration membranes in microbial fuel cells. Environmental Science & Technology, 2007. 41(3): p. 1004-1009.

36. Rozendal, R., et al., Performance of single chamber biocatalyzed electrolysis with different types of ion exchange membranes. Water Research, 2007. 41: p. 1984-1994.

37. Zuo, Y., S. Cheng, and B.E. Logan, Ion Exchange Membrane Cathodes for Scalable Microbial Fuel Cells. Environmental Science & Technology, 2008. 42(18): p. 6967-6972.

38. Sleutels, T.H.J.A., et al., Ion transport resistance in Microbial Electrolysis Cells with anion and cation exchange membranes. International Journal of Hydrogen Energy, 2009. 34(9): p. 3612-3620.

39. Harnisch, F., U. Schröder, and F. Scholz, The Suitability of Monopolar and Bipolar Ion Exchange Membranes as Separators for Biological Fuel Cells. Environmental Science & Technology, 2008. 42(5): p. 1740-1746.

40. Liu, H. and B.E. Logan, Electricity generation using an air-cathode single chamber microbial fuel cell in the presence and absence of a proton exchange membrane. Environ Sci Technol, 2004. 38(14): p. 4040-4046.

41. Logan, B., Scaling up microbial fuel cells and other bioelectrochemical systems. Applied Microbiology and Biotechnology, 2010. 85(6): p. 1665-1671.

42. Harnisch, F. and U. Schröder, Selectivity versus Mobility: Separation of Anode and Cathode in Microbial Bioelectrochemical Systems. ChemSusChem, 2009. 2(10): p. 921-926.

43. Zhang, X., et al., Separator Characteristics for Increasing Performance of Microbial Fuel Cells. Environmental Science & Technology, 2009. 43(21): p. 8456-8461.

44. Fan, Y., H. Hu, and H. Liu, Enhanced Coulombic efficiency and power density of air-cathode microbial fuel cells with an improved cell configuration. Journal of Power Sources, 2007. 171(2): p. 348-354.

45. Velasquez-Orta, S.B., T.P. Curtis, and B.E. Logan, Energy from algae using microbial fuel cells. Biotechnology and Bioengineering, 2009. 103(6): p. 1068-1076.

46. Zuo, Y., et al., Tubular membrane cathodes for scalable power generation in microbial fuel cells. Environmental Science & Technology, 2007. 41(9): p. 3347-3353.

47. Menicucci, J.H., et al., Procedure for determining maximum sustainable power generated by microbial fuel cells. Environmental Science & Technology, 2006. 40(3): p. 1062-1068.

48. Aelterman, P., et al., Continuous electricity generation at high voltages and currents using stacked microbial fuel cells. Environmental Science & Technology, 2006. 40: p. 3388-3394.

Page 27: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

14

49. Kim, J.R., et al., Modular tubular microbial fuel cells for energy recovery during sucrose wastewater treatment at low organic loading rate. Bioresource Technology, 2010. 101(4): p. 1190-1198.

50. Ieropoulos, I., J. Winfield, and J. Greenman, Effects of flow-rate, inoculum and time on the internal resistance of microbial fuel cells. Bioresource Technology, 2010. 101(10): p. 3520-3525.

51. Paulus, U.A., et al., Oxygen reduction on a high-surface area Pt/Vulcan carbon catalyst: a thin-film rotating ring-disk electrode study. Journal of Electroanalytical Chemistry, 2001. 495: p. 134-145.

52. Schmidt, T.J., et al., Characterization of high-surface-area electrocatalysts using a rotating disk electrode configuration. Journal of The Electrochemical Society, 1998. 145(7): p. 2354-2358.

53. Bard, A.J. and L.R. Faulkner, Electrochemical Methods: Fundamentals and Applications. 2nd ed. 2001, New York: Wiley.

54. Kim, J.R., et al., Application of Co-naphthalocyanine (CoNPc) as alternative cathode catalyst and support structure for microbial fuel cells. Bioresource Technology, 2011. 102: p. 342-347.

55. Yu, E.H., et al., Microbial Fuel Cell Performance with non-Pt Cathode Catalysts. Journal of Power Sources, 2007. 171: p. 275-281.

56. Gojkovic, S.L., S. Gupta, and R.F. Savinell, Heat-treated iron(III) tetramethoxyphenyl porphyrin supported on high-area carbon as an electrocatalyst for oxygen reduction - I. Characterization of the electrocatalyst. Journal of the Electrochemical Society, 1998. 145: p. 3493-3499.

57. Bandosz, T.J., J. Jagiello, and C. Contescu, Characterization of the surfaces of activated carbons in terms of their acidity constant distributions. Carbon, 1993. 31(7): p. 1193-1202.

58. Seredych, M. and T.J. Bandosz, Investigation of the enhancing effects of sulfur and/or oxygen functional groups of nanoporous carbons on adsorption of dibenzothiophenes. Carbon, 2011. 49: p. 1216-1224.

Page 28: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

15

Chapter 2

Analysis of polarization methods for elimination of power overshoot in microbial fuel cells

Abstract

Polarization curves from microbial fuel cells (MFCs) often show an unexpectedly large drop in

voltage with increased current densities, leading to a phenomenon in the power density curve

referred to as “power overshoot”. Linear sweep voltammetry (LSV, 1 mV s−1) and variable

external resistances (at fixed intervals of 20 min) over a single fed-batch cycle in an MFC both

resulted in power overshoot in power density curves due to anode potentials. Increasing the anode

enrichment time from 30 days to 100 days did not eliminate overshoot, suggesting that

insufficient enrichment of the anode biofilm was not the primary cause. Running the reactor at a

fixed resistance for a full fed-batch cycle (~1 to 2 days), however, completely eliminated the

overshoot in the power density curve. These results show that long times at a fixed resistance are

needed to stabilize current generation by bacteria in MFCs, and that even relatively slow LSV

scan rates and long times between switching circuit loads during a fed-batch cycle may produce

inaccurate polarization and power density results for these biological systems.

This chapter was published as: Watson, V. J. and Logan, B. E., Analysis of polarization methods for elimination of power overshoot in microbial fuel cells. Electrochemistry Communications 2011, 13, (1), 54-56.

Page 29: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

16

2.1 Introduction

Much of the research involving the use of microbial fuel cells (MFC) for combined

electricity production and wastewater treatment is focused on producing the most power through

improved reactor designs [1]. However, estimates of the amount of power that can be produced in

an MFC are a function of the technique used to obtain polarization data. Linear sweep

voltammetry (LSV) is commonly used in MFC studies to obtain polarization data, but high scan

rates can overestimate power production [2]. An alternate approach is to vary the circuit

resistance at fixed time intervals, ranging from 10 s to 24 h [3,4]. There have been few studies

comparing the different techniques, but in one study it was found that power production with scan

rates higher than 0.1 mV s−1 produced higher power densities than those where the circuit

resistance was varied [2]. A common problem often encountered when evaluating polarization

curves is “power overshoot” [2,5–7]. Power overshoot refers to the response of the system at high

current densities (past the maximum power) in a power density curve where the cell voltage and

current drop very quickly resulting in a doubling back of the power density curve, producing

lower power than previously measured for the lower current densities [7]. One hypothesis on the

cause of this power overshoot is that as the current resistance is decreased the bacteria on the

anode are unable to produce sufficient current at lower voltages [7]. However, there does not

seem to be a correlation in the literature between the magnitude of current density and power

curve shape.

Accurate methods are needed to ensure that power densities reported by different

researchers reflect the true performance of the MFC. So far there has been no study on how

different polarization techniques might affect the development of power overshoot or methods to

eliminate it. We therefore examined MFCs that exhibited power overshoot when analyzed using

common LSV and fixed resistances methods, and showed that power overshoot could be

Page 30: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

17

eliminated by allowing sufficient time for the biofilm to adjust to a change in resistance by using

a single fixed resistance for each separate fed-batch cycle.

2.2 Experimental procedures

2.2.1 MFC reactor construction and operation

Cube-shaped MFCs with a cylindrical chamber (28 mL, 7 cm2 cross section) were

constructed without a membrane as previously described [8]. The brush anode was constructed

from carbon fibers (PANEX®33 160 K, ZOLTEK) wound into a titanium wire core (2.5 cm

diameter, 2.5 cm length, 0.22 m2 surface area) which was heat treated at 450 °C [9] and placed

horizontally in the center of the chamber. Air cathodes (projected surface area of 7 cm2) were

made from carbon cloth (30 wt.% wet proofing polymer, #B1B30WP, BASF Corp.) with four

PTFE diffusion layers and 0.5 mg-Pt cm−2 [10]. The electrode spacing was 2.5 cm (center of the

anode to the face of the cathode).

Seven MFCs were inoculated using effluent from another MFC operated under similar

conditions (50% v/v inoculum and medium) [3] at 30 °C in a controlled-climate room, and were

covered to exclude light. The medium was a 100 mM phosphate buffer solution (PBS) containing

(g L−1): 9.125 Na2HPO4, 4.904 NaH2PO4·H2O, 0.31 NH4Cl, and 0.13 KCl; pH 7, vitamins and

minerals [11]; and 1 g L−1 sodium acetate. The electrodes were connected through a 1000 Ω

resistor, except as noted. Once an MFC produced ≥100 mV at 1000 Ω, no additional inoculum

was added to the medium over subsequent fed-batch cycles. MFCs were considered enriched and

ready for testing once they achieved the same maximum voltage for three consecutive batch

cycles [3].

Page 31: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

18

2.2.2 Analysis

The voltage across the resistor was recorded every 30 min using a multimeter data

acquisition system (model 2700 Keithley Instruments, Cleveland, OH). Polarization was

performed once the voltage stabilized after the MFC was fed. Polarization curves were obtained

by three different methods. In the first method (single-cycle), conducted on days 30 and 100,

various external resistances (OCV, 1000, 500, 250, 100, 75, 50, and 25 Ω, except where noted)

were connected across the MFC, with each resistance being connected for 20 min and the voltage

recorded using a digital multimeter (Model 83 III, Fluke) over a single batch cycle [12]. For the

second method (multiple-cycle), conducted after 100 days, the maximum sustainable voltage over

the cycle (typically sustained for 7 to 30 h depending on the total length of the cycle) was

recorded using a single resistor (decreasing for each batch) over a complete fed-batch cycle [12].

Each resistance was tested for three consecutive cycles to ensure that the voltage response was

unchanged with successive cycles. The third method, linear sweep voltammetry (LSV), was run

after the multiple-cycle method. LSVs were run three times at the recommended scan rate of 1

mV s−1 over a range of 0.5 V starting from the measured open circuit voltage [13]. For the applied

resistance methods, current density was calculated from I=E/R, where I is the current, E the

measured voltage, and R the external resistance, and normalized to the projected cathode surface

area. Power densities were calculated using P=IE, and normalized by the projected cathode

surface area [13].

Page 32: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

19

2.3 Results

2.3.1. Polarization by varied resistance, single-cycle method

Polarization curves obtained using the single-cycle method (20 min intervals) exhibited a

steep drop in voltage at higher current densities resulting in power overshoot occurrences in

power density curves. In an example power density curve (Figure 2-1), for the 30 day test the

MFC produced a maximum power of 856mWm−2 (0.22 mA cm−2) before the power rapidly

decreased. The measured electrode potentials indicate that the anode potential was responsible for

power overshoot. At the point of overshoot, the anode potential rapidly became more positive

(from −0.354 V to +0.022 V) and the current decreased while the cathode potential returned to a

value consistent with that previously measured at that current (Figure 2-2). Not all reactors tested

exhibited power overshoot, and therefore additional tests were conducted to further investigate

this phenomenon.

To rule out substrate depletion at the end of the polarization cycle as the cause of the

power overshoot, each MFC was refilled with fresh medium, left to stabilize at 1000 Ω (30 min to

1 h) and polarization testing was started at a lower resistance (500 Ω instead of OCV). The same

rapid increase in anode potential was still observed under these new starting conditions (−0.341 V

to 0.034 V) at the same resistances (from 250 Ω to 100 Ω).

In order to see if the overshoot was caused by incomplete enrichment of the anode with

biofilm, the MFCs were maintained over repeated fed-batch cycles for another 70 days. After 100

days of operation, polarization curves still exhibited power overshoot using the 20 min single-

cycle method (Figure 2-1) (1027mWm−2, 0.24 mA cm−2). As before, the anode potential dropped

off when changing resistance from 250 Ω to 100 Ω (Figure 2-2).

Page 33: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

20

Figure 2-1. (A) Power density and (B) polarization curves for single-cycle (20 min) resistance changes at 30 days (◊) and 100 days (Δ) and multiple-cycle resistance changes ().

Figure 2-2. Electrode (A=anode and C=cathode) potential measurements (vs. Ag/AgCl) during cell polarization for single-cycle (20 min) resistance changes at 30 days (◊) and 100 days (Δ) and multiple-cycle resistance changes ().

Page 34: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

21

2.3.2. Polarization by varied resistance, multiple-cycle method

Polarization data using the multiple-cycle method (a separate resistor for each fed-batch

cycle) conducted after 100 days, produced power density curves without overshoot. At low

current densities, the power curve followed that obtained using the single-cycle (20 min) method,

but at the point where the other curves dropped off, the multiple-cycle curve extended to a higher

power density (1296 mW m−2) with increased current production (0.61 mA cm−2) (Figure 2-1).

The anode potential did not undergo a rapid increase during the multiple-cycle method as it did

when measuring current after 20 min in the single-cycle method (Figure 2-2).

The fed-batch cycle curves resembled those reported previously [8], where the cell

voltage increases rapidly (over a few hours) after the reactor is fed, and then stabilizes for most of

the cycle. Polarization data were obtained during this stable voltage period. Only one full cycle is

needed at each different resistance for the multiple-cycle method. Power production over three

consecutive cycles at the same applied resistance did not show any noticeable change in

maximum voltage from the first cycle to the third cycle (data not shown).

2.3.3 Polarization by LSV

Power density curves obtained using LSV (following the above multiple-cycle results)

also exhibited power overshoot as shown by a doubling back of the power density curve (Figure

2-3). The maximum power recorded for cycle 1 (2530 mW m−2, 0.56 mA cm−2) was higher than

for cycle 2 (1840 mWm−2, 0.44 mA cm−2) and cycle 3 (1860 mW m−2, 0.44 mA cm−2). In

addition, there was another type of power overshoot in that the maximum power densities in three

of the LSV curves were much higher than power densities measured by the applied resistance

Page 35: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

22

methods. Also, none of the LSV cycles measured power at current densities as high as was those

measured using the multiple-cycle method.

Figure 2-3. (A) Power density and (B) polarization curves from LSV at 1 mV s−1 for three consecutive scans.

2.4 Discussion

Power overshoot was observed in power density curves obtained using the single-cycle

(20 min intervals) and LSV (1 mV s−1) methods even with MFCs enriched for 100 days or more.

However, power density curves obtained using the multiple-cycle method did not exhibit

overshoot even at current densities up to 0.82 mA cm−2. The overshoot resulted from a rapid

increase in the anode potential as resistance to current flow was decreased, indicating electron

transfer limitation at the anode. The limitation was most likely related to a slow response from the

Page 36: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

23

microbes to adjust to the new resistance [7]. When the biofilm was given sufficient time to adjust

to a set resistance by fixing the resistance over an entire cycle (multiple-cycle method), the

biofilm produced increased currents at lower voltages. Since the maximum cell potential did not

change during 3 cycles at the same resistance, the improved performance using the multiple-cycle

method was not a consequence of the long-term enrichment of the anode community. Instead,

these results showed that the biofilm needed much more time to adapt to the applied resistance

than could be obtained in brief intervals at fixed resistances. Additional data in our laboratory has

shown that even 1-hr intervals or slower LSV scan rates do not eliminate overshoot. For fed-batch

MFCs, longer time periods are problematic as the series of resistances needed to produce a

polarization curve cannot be obtained over the whole cycle due to depletion of the substrate.

Thus, it is recommended that the full cycle (or in other cases at least a day or more) be used at a

fixed resistance when obtaining polarization data in MFCs that exhibit overshoot.

It was also observed that the maximum power produced in a single-cycle polarization

curve was less than that produced in a multiple-cycle curve. Thus, power would be

underestimated as a result of reporting data where power overshoot occurs. A full polarization

curve should be obtained in order to see if overshoot is present. Maximum power densities

obtained by LSV were all higher than that obtained using either of the resistance methods,

consistent with previous studies [2].

Our results show that it is important to use a reliable and consistent method for measuring

maximum current densities in MFCs in order to obtain valid results concerning maximum power

densities. If overshoot occurs, it may not be possible to properly compare maximum power from

different MFC studies or different conditions within the same study. For fed-batch MFCs, the use

of multiple cycles of data, each at different fixed resistances, offers the best method to obtain

polarization data for producing power curves representative of performance during steady

operation conditions.

Page 37: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

24

2.5 Acknowledgements

This research was supported under a National Science Foundation (NSF) Graduate

Research Fellowship, NSF grant CBET-0730359, and the King Abdullah University of Science

and Technology (KAUST) (Award KUS-I1-003-13).

2.6 References

1. B. Logan, Appl. Microbiol. Biotechnol. 85 (2010) 1665–1671. 2. S.B. Velasquez-Orta, T.P. Curtis, B.E. Logan, Biotechnol. Bioeng. 103 (2009)

1068–1076. 3. Y. Zuo, S. Cheng, D. Call, B.E. Logan, Environ. Sci. Technol. 41 (2007) 3347–

3353. 4. J.H. Menicucci, H. Beyenal, E. Marsili, R.A. Veluchamy, G. Demir, Z.

Lewandowski, Environ. Sci. Technol. 40 (2006) 1062–1068. 5. P. Aelterman, K. Rabaey, T.H. Pham, N. Boon, W. Verstraete, Environ. Sci.

Technol. 40 (2006) 3388–3394. 6. J.R. Kim, G.C. Premier, F.R. Hawkes, J. Rodríguez, R.M. Dinsdale, A.J. Guwy,

Bioresour. Technol. 101 (2010) 1190–1198. 7. I. Ieropoulos, J. Winfield, J. Greenman, Bioresour. Technol. 101 (2010) 3520–

3525. 8. H. Liu, B.E. Logan, Environ. Sci. Technol. 38 (2004) 4040–4046. 9. Y. Feng, Q. Yang, X. Wang, B.E. Logan, J. Power Sources 195 (2010) 1841–

1844. 10. S. Cheng, H. Liu, B.E. Logan, Electrochem. Commun. 8 (2006) 489–494. 11. O. Bretschger, A. Obraztsova, C.A. Sturm, I.S. Chang, Y.A. Gorby, S.B. Reed,

D.E. Culley, C.L. Reardon, S. Barua, M.F. Romine, J. Zhou, A.S. Beliaev, R. Bouhenni, D. Saffarini, F. Mansfeld, B.-H. Kim, J.K. Fredrickson, K.H. Nealson, Appl. Environ. Microbiol. 73 (2007) 7003–7012.

12. J. Heilmann, B.E. Logan, Water Environ. Res. 78 (2006) 531–537. 13. B.E. Logan, P. Aelterman, B. Hamelers, R. Rozendal, U. Schröder, J. Keller, S.

Freguiac, W. Verstraete, K. Rabaey, Environ. Sci. Technol. 40 (2006) 5181–5192.

Page 38: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

25

Chapter 3

Polymer coatings as separator layers for microbial fuel cell cathodes

Abstract

Membrane separators reduce oxygen flux from the cathode into the anolyte in microbial fuel cells

(MFCs), but water accumulation and pH gradients between the separator and cathode reduces

performance. Air cathodes were spray-coated (water-facing side) with anion exchange, cation

exchange, and neutral polymer coatings of different thicknesses to incorporate the separator into

the cathode. The anion exchange polymer coating resulted in greater power density (1167 ± 135

mW m−2) than a cation exchange coating (439 ± 2 mW m−2). This power output was similar to

that produced by a Nafion-coated cathode (1114 ± 174mW m−2), and slightly lower than the

uncoated cathode (1384 ± 82 mW m−2). Thicker coatings reduced oxygen diffusion into the

electrolyte and increased coulombic efficiency (CE = 56–64%) relative to an uncoated cathode

(29 ± 8%), but decreased power production (255–574 mW m−2). Electrochemical characterization

of the cathodes ex situ to the MFC showed that the cathodes with the lowest charge transfer

resistance and the highest oxygen reduction activity produced the most power in MFC tests. The

results on hydrophilic cathode separator layers revealed a tradeoff between power and CE.

Cathodes coated with a thin coating of anion exchange polymer show promise for controlling

oxygen transfer while minimally affecting power production.

This chapter was published as: Watson, V. J.; Saito, T.; Hickner, M. A.; Logan, B. E., Polymer coatings as separator layers for microbial fuel cell cathodes. Journal of Power Sources 2011, 196, (6), 3009-3014.

Page 39: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

26

3.1 Introduction

Microbial fuel cells (MFCs) represent one of the latest innovations for the treatment of

wastewater streams. MFC technology combines waste treatment with electricity production [1].

Typical MFCs consist of a microbe-enriched anode where organic matter is oxidized. Electrons

are conducted through a circuit to the air-fed cathode consisting of a porous carbon structure with

platinum catalyst, where oxygen is reduced to water [2]. Some MFCs include an ion exchange

membrane in the electrolyte compartment between the anode and cathode. However, membranes

have been shown to negatively impact the power production of the MFC by increasing the

internal resistance of the cell and inducing pH gradients during cell operation [3,4]. MFC power

production can be improved by removing the membrane from the system [5] and reducing the

electrode spacing to decrease ohmic losses. When the electrodes become closely spaced,

however, a separator is needed to prevent short circuiting and also to reduce oxygen diffusion into

the anode chamber which can adversely affect power production [1,6].

The performance characteristics of membrane separators have been investigated in

bioelectrochemical systems, including cation exchange (CEM), anion exchange (AEM), bipolar,

and ultrafiltration membranes [4,7–10]. It has been shown that the cations (Na+, K+, and NH4+)

are preferentially transferred through the CEM due to their high concentrations rather than

protons to maintain charge balance, and as a result there is a decrease in performance due to pH

changes [3,11]. AEMs outperform CEMs and other types of membranes in MFCs and microbial

electrolysis cells (MECs) mostly due to lower internal resistances that result from lower charge

transport resistance [4,8,9,12]. Charge balance can be facilitated by transfer of buffer anions

(such as phosphate) when using an AEM [4]. However, both AEMs and CEMs negatively impact

microbial fuel cell performance due to the formation of a pH gradient at the electrodes [7].

Page 40: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

27

Oxygen diffusion into the anode chamber negatively affects MFC performance by

serving as an alternative electron acceptor for the facultative bacteria at the anode. If the bacteria

use the oxygen as the terminal electron acceptor instead of the anode current collector, the

coulombic efficiency (CE) will decrease, the anode potential will become more positive, and the

current density will decrease [13]. Cloth (J-cloth) separators have been used to decrease oxygen

diffusion into the anolyte, but over time the cloth became completely degraded by the bacteria in

the reactor [14,15]. Positioning a glass fiber separator next to the cathode in an MFC with 2 cm

electrode spacing has been shown to increase CE to 80% compared to 30% without a separator

[14]. Power production with the separator decreased from 896 mW m−2 to 791 mW m−2 as a result

of decreased cathode potential and increased ohmic resistance. To improve power production, the

electrode spacing was decreased using the separator which prevented short-circuiting. With the

decreased electrode spacing, the power density increased to 1195 mW m−2 while maintaining CE

at 80%. In the same study, growth of biofilm on the cathode was also found to improve CE over

time due to a decrease in oxygen diffusion into the electrolyte from the air cathode, but the

biofilm also hindered proton migration to the cathode and limited power production [14].

Zhang et al. [16] placed AEMs and CEMs in the electrolyte compartment directly

adjacent to the cathode and obtained around 90% CE. However, the membranes deformed after

several cycles due to membrane swelling during ion and water transport, and the deformation

created a void space between the membrane and electrode filled with water and gas. The water

trapped between the membrane and the cathode had a higher pH than the anode chamber and

decreased the cathode potential. The researchers used stainless steel mesh to keep the membrane

pressed against the cathode and prevent water accumulation behind the membrane. In this

configuration, the ohmic resistance of the reactor decreased from 120Ω to 15Ω with the AEM and

from 49Ω to 16Ω with the CEM, while the power density increased from 16 W m−3 to 46 W m−3

Page 41: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

28

with the AEM and from 21 W m−3 to 32 W m−3 with the CEM [16]. In previous studies when

Nafion was hot-pressed on to carbon cloth preventing deformation, the CE increased from 9–12%

to 40–50%, but the power density decreased from 12.5 W m−3 to 6.6 W m−3 [5]. Hot pressing the

membrane to the cathode decreased the power by increasing the ohmic resistance of the

membrane, likely due to adverse effects of the bonding process on membrane permeability [4].

Therefore, it is important to incorporate the membrane into the cathode to prevent deformation

while also striving to minimize the ohmic resistance of the membrane.

In next-generation MFC systems, a separator between the anode and cathodes will be

important to facilitate minimum electrode spacing while preventing short circuiting of the

electrodes [1]. The separators must limit oxygen diffusion to the anolyte while not impeding

proton transfer to the cathode catalyst. This study explored the use of spray coating for applying

thin layers of hydrophilic cation exchange, anion exchange, and neutral polymers to the

electrolyte side of the cathode structure and measured the layers’ effect on power production and

CE with respect to polymer type, oxygen diffusivity, and biofilm growth at the cathode.

3.2 Materials and methods

3.2.1 Polymers

Bisphenol A-based poly(sulfone) (Udel P-3500 LCD, Mw 79,000 g mol−1, 1.24 g cm−3)

and poly(phenylsulfone) (Radel R-5500, Mw 63,000 g mol−1, 1.29 g cm−3) were kindly donated by

Solvay Advanced Polymers, LLC. Radel was aminated (A-Radel, ion exchange capacity of IEC =

2.64 meq g−1) or sulfonated (S-Radel, IEC = 2.54meq g−1) as previously described [17–19].

Nafion solution (Nafion® 117 solution), ∼5 wt% in a mixture of lower aliphatic alcohols and

water was purchased from Aldrich and used as received. Poly(styrene)-b-poly(ethylene oxide)

Page 42: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

29

diblock copolymer PS156-b-PEO110 (PEO-110, Mn 21,100 g mol−1, Mw/Mn = 1.01) was

synthesized as previously described [20], where subscripted numbers denote the corresponding

number of repeat units of each block. Polymer solutions (5 wt%) were prepared by dissolving

Udel and PEO-110 in tetrahydrofuran and A-Radel and S-Radel in methanol. Properties of each

of the polymers are summarized in Table 3-1.

Table 3-1. Properties of polymers and cathode coatings.

Cathode Type IEC

(meq g−1) IECν (wet) (meq cm−3)

Water uptake

Solvent Density (g cm−3)

Weight (mg)

Thickness (dry) (µm)

Thickness (wet) (µm)

Nafion-62 CEM 0.91 1.59 20% Aliph-Alc 2.10 122.3 ± 0.8 52 ± 0.3 62 ± 0.4 A-Radel-146 AEM 2.64 1.22 180% MeOH 1.29 75.8 ± 0.1 52 ± 0.0 146 ± 0.0 A-Radel-67 AEM 2.64 1.22 180% MeOH 1.29 34.7 ± 4.5 24 ± 1.9 67 ± 2.3 S-Radel-60 CEM 2.54 1.93 70% MeOH 1.29 51.7 ± 0.7 35 ± 0.3 60 ± 0.4 S-Radel-47 CEM 2.54 1.93 70% MeOH 1.29 40.7 ± 1.0 28 ± 0.4 47 ± 0.5 PEO110-101 N-Phila 0 0 50% THF 1.10 83.8 ± 6.6 67 ± 2.8 101 ± 3.4 Udel-32 N-Phobb 0 0 0% THF 1.24 44.5 ± 2.8 32 ± 1.2 32 ± 1.4 a Neutral – hydrophilic polymer. b Neutral – hydrophobic polymer.

3.2.2 Cathode construction

Platinum-catalyzed air cathodes (projected surface area of 7cm2) were constructed from

carbon cloth containing 30 wt% wet proofing polymer (#B1B30WP, BASF Corp.) with PTFE

diffusion layers, and 0.5 mg-Pt cm−2 catalyst loading [21]. The polymer layers were applied to the

cathodes in layers using an air brush (Paache, BearAir, S. Easton, MA). The sprayed polymer

coating was allowed to dry between layers and then checked for resistivity using a handheld

digital multimeter (Model 83 III, Fluke) and weighed to determine the amount of coating applied

(Table 3-1). Once coated, all cathode surfaces produced a resistance greater than the

Page 43: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

30

measurement range of the multimeter, effectively electrically insulating the electrolyte-facing

surface of the cathode. Two cathodes were coated for each polymer tested. The average wet and

dry thicknesses of the coatings were calculated from the measured mass of the applied polymer,

the density of the dry polymer, and the polymer’s water uptake (Table 3-1). The thicknesses of

the polymer layers applied to the solution side of the cathode structure are included in the names

of the samples as indicated by the number to the right of the dash, for instance A-Radel-146

indicates that the A-Radel coating on those cathodes averaged 146 µm in thickness.

3.2.3 MFC reactor construction and operation

Cube-shaped MFCs were constructed as previously described [5]. The anode chamber

was a 28 mL cylindrical chamber (7 cm2 cross section) bored into a Lexan block. The brush

anode was constructed from carbon fibers (PANEX®33 160K, ZOLTEK) wound into a titanium

wire core (2.5 cm diameter, 2.5 cm length, and 0.22 m2 surface area) which was heat treated at

450 °C [22] then placed horizontally in the center of the cylinder. The electrode spacing was

2.5cm (center of the anode to the face of the cathode).

Effluent from the anode chamber of an enriched MFC operated under similar conditions

to those in this study was used for the mixed culture inoculum. The medium used in MFC

performance tests was a 100 mM phosphate buffer solution (PBS) (9.125 g L−1 Na2HPO4, 4.904 g

L−1 NaH2PO4·H2O, 0.31 g L−1 NH4Cl, and 0.13 g L−1 KCl; pH 7) with vitamins and minerals [23]

and 1g L−1 sodium acetate. The PBS concentration was doubled to 200 mM in tests where

indicated. MFCs were inoculated with a 50% (v v−1) inoculum of effluent and medium and were

covered to exclude light. The electrodes were connected through a 1000 Ω resistor, except as

noted. Once an MFC produced ≥100 mV, no additional inoculum was added to the medium over

Page 44: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

31

subsequent fed batch cycles. All MFCs were operated at 30 C in a controlled climate room. The

MFCs were considered enriched and ready for testing once they achieved the same maximum

voltage for three consecutive batch cycles. Once the MFC anodes were enriched, the uncoated

cathodes used for startup were removed and the coated cathodes and new uncoated cathodes were

placed in the reactors. All MFC tests were conducted with duplicate cathodes and averages over

two cycles were reported with standard deviations for n = 4 (two cathodes over two cycles),

except for CE which was averaged from duplicate reactors over three cycles (n = 6).

3.2.4 Analysis

The voltage across the resistor was recorded every 30 min using a multimeter (model

2700 Keithley Instruments, Cleveland, OH) with a computerized data acquisition system.

Polarization curves were obtained by applying a different external resistance to the circuit for a

complete batch cycle and the maximum sustainable voltage (typically sustained for 7–30 h

depending on the total length of the cycle) was recorded for each resistance. Current density was

calculated from I = E/R, where I is the current, E the measured voltage, and R the external

resistance, and normalized to the projected cathode surface area. Power densities were calculated

using P = IE, and normalized by the projected cathode surface area [24].

CE was calculated from the ratio of the total electrical charge produced during the

experiment (at 1000 Ω) to the theoretical amount of electrons available from the oxidation of

acetate to carbon dioxide. Therefore, CE [%] = (CEx/CTh) × 100, where 𝐶𝐸𝑥 = ∑ (𝐸𝑖𝑡𝑖)/𝑅𝑇𝑡=1 , CTh

= FbMv, F is Faraday’s constant (96,485 C mol e−1), b is the number of moles of electrons

available per mole of substrate (8 mol e (mol acetate)−1), M is the acetate concentration (mol L−1),

and v is the volume of liquid in the anode chamber (L) [24].

Page 45: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

32

The oxygen flux into the electrolyte chamber through each cathode was calculated by

measuring the change in dissolved oxygen concentration (NeoFox, Ocean Optics Inc., FL) over

time in a stirred abiotic MFC reactor (30 mL) without an active anode as previously described

[4].

The impedance of each cathode half-cell was measured by electrochemical impedance

spectroscopy (EIS) at 0.1 V (vs. Ag/AgCl) over a frequency of 100,000–0.1 Hz with sinusoidal

perturbation of 10 mV using a potentiostat (PC 4/750, Gamry Instrument Inc.) at 30 °C. The half-

cell consisted of a 7 cm2 platinum disk counter electrode set parallel to the test cathode and

equipped with an Ag/AgCl reference electrode (+0.2 V vs. NHE) (RE-5B, BASI, IN). The test

cell was filled with 200 mM PBS (13 mL, pH 7) without substrate or other nutrients. The

combined solution and membrane resistances (Rs + Rm) were obtained from Nyquist impedance

plots at the point where Zimag was equal to zero at high frequency. The charge transfer resistance

(Rct) for each cathode was estimated from a semi-circular fit of the charge transfer impedance in

the Nyquist plot [25,26].

The oxygen reduction response of each cathode was measured by linear sweep

voltammetry (LSV) using the same experimental setup as with EIS at a scan rate of 1 mV s−1 over

the range of 0.6 to −0.3 V (vs. Ag/AgCl) with current interrupt correction. The oxygen reduction

activity of the cathodes was measured in both 200 mM and 100 mM PBS solution (pH 7).

3.3 Results

3.3.1 MFC performance

MFCs with cathodes coated with a thin layer of anion exchange polymer (A-Radel-67)

produced approximately the same maximum power (Table 3-2) as the cells with cathodes coated

Page 46: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

33

with Nafion of similar thickness (Nafion-62) (Figure 3-1). MFC tests with uncoated cathodes

resulted in slightly higher power density than Nafion-62 or A-Radel-67 coated cathodes. MFCs

with cation exchange Radel polymer coatings on the cathode (S-Radel-47) produced much less

power than the cells with the uncoated control cathodes, as did reactors with thin layers of Udel

hydrophobic polymer, Udel-32. The A-Radel-146 and S-Radel-60 coated cathodes had thicker

coatings and produced less power than the cathodes coated with a thinner layer of the same

polymer (A-Radel-67 and S-Radel-47), most likely due to increased impedance of proton transfer.

The observed differences in power production during polarization were due to differences in

cathode potentials (Figure 3-2) since anode potentials did not vary over the current density range

tested.

Table 3-2. Oxygen flux, combined solution and membrane resistance, charge transfer resistance, and maximum power density of cathodes.

Cathode Oxygen flux

(mg cm-2 h-1) Rs + Rm (Ω) Rct (Ω) Maximum power density (mWm−2)

Uncoated 0.055 5 19 1384 ± 82

Nafion-62 0.022 7 57 1114 ± 174 A-Radel-146 0.010 7 85 574 ± 32 A-Radel-67 0.023 7 22 1167 ± 135 S-Radel-60 0.004 9 >100 255 ± 28 S-Radel-47 0.012 7 >100 439 ± 2 PEO110-101 0.002 7 >100 307 ± 9 Udel-32 0.008 18 >100 266 ± 16 a Neutral – hydrophilic polymer. b Neutral – hydrophobic polymer.

Page 47: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

34

Figure 3-1. (A) Power density and (B) polarization curves for polymer-coated cathodes.

Figure 3-2. Electrode potential measurements (vs. Ag/AgCl) during cell polarization.

Page 48: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

35

The CEs of the MFCs ranged between 29% and 64% (Figure 3-3; fixed external

resistance of 1000 Ω). Cathodes with thicker coatings of the same polymer type had higher CEs

(A-Radel-146, 56 ± 2%; S-Radel-60, 64 ± 5%) than the cathodes with thinner coatings (ARadel-

67, 33 ± 8%; S-Radel-47, 40 ± 10%).

Figure 3-3. Coulombic efficiencies for cycles run at 1000 Ω.

3.3.2 Electrochemical performance

The A-Radel-67 cathode had the lowest impedance (Rs + Rm =7 Ω and Rct =22 Ω) of all

the coated cathodes (Figure 3-4 and Table 3-2) and only slightly higher resistances than the

uncoated cathode (Rs + Rm =5 Ω and Rct =19 Ω). Since Rm is zero for the uncoated cathode, Rs is 5

Ω for the half-cell control geometry used in these experiments. Thus, for all coated cathodes

except Udel-32, the coating added an Rm of between 2 and 4 Ω. However, larger effects of the

coatings can be observed in the Rct, most likely due to the decrease of reactant concentration at

the catalyst or a decrease in available catalyst sites. A-Radel-67 had the smallest increase in Rct

(+3 Ω) compared to the uncoated cathode while Nafion-62 showed an Rct of 38 Ω greater than the

Page 49: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

36

uncoated control cathode. S-Radel-60, S-Radel-47, PEO110-101, and Udel-32 coated cathodes

had Rct values of greater than 100 Ω.

Figure 3-4. EIS of coated and uncoated cathodes at 0.1V (vs. Ag/AgCl) (200mM PBS).

The effect of the coatings on the oxygen reduction performance of the cathodes can be

observed by the decrease in current density during LSV testing compared to the uncoated control

(Figure 3-5). The current densities obtained from LSV for each coated cathode showed the same

trends as power production in MFC tests. For example, A-Radel-67 produced higher current

densities in LSV and the maximum power in the MFC tests, and S-Radel-60 and Udel-32

produced the lowest current densities in LSV and the lowest power densities. LSV showed

similar trends between cathodes using either the 100 mM PBS or 200 mM PBS solution.

Page 50: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

37

Figure 3-5. LSV of coated and uncoated cathodes (100mM PBS).

3.3.3 Oxygen diffusion and biofilm growth

The coatings applied to the cathode decreased the rate of oxygen diffusion into the anode

solution as demonstrated by the measured oxygen flux into the electrolyte compartment (Table 3-

2). The decrease in oxygen diffusion was not exclusively a function of the amount of polymer

applied (i.e., the thickness or the weight of the coating), but was a combined result of the type of

polymer, the processing of the layer (e.g. solvent used for coating deposition), and the coating

thickness on the cathode. The oxygen flux was inversely related to the CE during MFC testing.

The cathodes with the highest rate of oxygen diffusion and the lowest CE developed a significant

layer of biofilm after 100 days of operation (Figure 3-6). S-Radel-60, PEO110-101, A-Radel-60,

and Udel-32 cathodes did not develop a visible biofilm layer.

Page 51: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

38

Figure 3-6. Optical images of biofilm growth on cathodes (100 days).

3.4 Discussion

Of the coated cathodes, the MFCs using the A-Radel-67 cathodes produced the highest

power density. In general, MFCs that had cathodes with the lowest Rct achieved the highest power

density (Figure 3-7). The A-Radel (AEM) had a lower Rct and higher power production than both

of the CEMs (S-Radel and Nafion). The better performance of the A-Radel is consistent with

results in previous studies comparing AEM and CEM separators in MFCs [4,8] which indicate

phosphate anions buffer pH changes and maintain charge balance. We therefore conclude that

positively charged quaternary ammonium groups on the AEM layer aid anion transport and result

in less accumulation of cations compared to CEM layers. The preferential anion transport in

AEMs may also decrease the pH gradient toward the catalyst moiety compared to that of CEM

layers [9]. The AEM had higher water uptake than the CEM when comparing similar IECs and

polymer backbones (i.e. A-Radel and S-Radel). The higher water uptake of the AEM likely

Page 52: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

39

decreased its ion transport resistance which could have also contributed to the higher power

density output of AEM coating than that of the CEM coating.

Figure 3-7. Inverse relationship between Rct and power density.

Although S-Radel and Nafion are both CEMs, the S-Radel hindered power production

more than the Nafion coating of the same thickness, which is reflected in the increase in Rct. The

greater Rct can be explained by considering the IECν, the volumetric concentration of ions in the

swollen polymer. The S-Radel had a higher IECν than Nafion and as seen in previous studies, the

higher IECν can impede proton diffusion at neutral pH [19]. Sulfonate groups in the CEM layers

were most likely saturated with Na+ and K+ rather than H+ due to the high Na+ and K+

concentration in the electrolyte. The accumulated cations hindered proton diffusion through the

CEM layer on the cathode and within the electrode where the layers had penetrated the porous

structure. It is also possible that polymer seepage into the cathode pores inhibited oxygen

transport to the catalyst surface, which increased Rct. Rct of the A-Radel cathodes increased and

the corresponding maximum power density decreased as the applied layer thickness increased and

the same effect was observed for the S-Radel cathodes.

The uncharged, hydrophilic polymer coatings of PEO110 had agreater Rct than the A-

Radel coating of similar thickness (A-Radel-146 compared to PEO110-101, and A-Radel-67

Page 53: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

40

compared to Udel-32), most likely due to the A-Radel coating having a greater water uptake

(Table 3-1) and therefore less impedance to proton transfer. The significant increase in Rct for

PEO110-101, with a reasonably high water uptake, implies that anion transport to decrease the

pH gradient in an AEM may be an important factor in the resulting Rct.

In general, there was an inverse relationship between the maximum power density and

CE (Figure 3-8), except for the S-Radel-47, which had a lower power density (440 ± 4 mW m−2)

and lower CE (40 ± 10%) than the A-Radel-146 (574 ± 32 mW m−2 and CE = 56 ± 2%). The

cathode coatings with lower oxygen permeability did not show an improvement in anode

potential resulting from a decrease in oxygen intrusion, most likely due to biofilm formation on

coatings with higher oxygen permeability, which limited oxygen diffusion to the anode. The

biofilm formation on the cathodes with higher oxygen permeability was most likely the cause of

the decrease in CE compared to the less permeable cathodes. Despite similar anode performance,

the cathodes with less oxygen permeability and higher CE produced less power due to an increase

in Rct caused by the increased resistance of the coatings to either proton or oxygen diffusion to the

catalyst surface.

Figure 3-8. Inverse relationship between CE and power density.

Page 54: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

41

Polymer coated cathodes can be useful in MFC designs as further efforts are made to

develop polymer coatings that facilitates proton transfer to the cathode but limit oxygen diffusion

into the electrolyte and provide an electrically insulating surface. Anion exchange polymers such

as A-Radel, integrated as a thin membrane coating into MFC cathodes, have potential for

controlling oxygen diffusion into the MFC while minimally affecting power production.

3.5 Acknowledgements

This research was supported under a National Science Foundation Graduate Research

Fellowship, National Science Foundation Grant CBET-0730359, and the King Abdullah

University of Science and Technology (KAUST) (Award KUS-I1-003-13). Thanks to Solvay

Advanced Polymers for the donation of Radel® and Udel® polymer and to Justin Tokash for

insights into EIS theory and application.

3.6 References

1. B. Logan, Appl. Microbiol. Biotechnol. 85 (2010) 1665–1671. 2. B.E. Logan, Microbial Fuel Cells, John Wiley & Sons, Inc., Hoboken, NJ, 2007. 3. R.A. Rozendal, H.V.V. Hamelers, C.J.N. Guisman, Environ. Sci. Technol. 40

(2006) 5206–5211. 4. J.R. Kim, S. Cheng, S.-E. Oh, B.E. Logan, Environ. Sci. Technol. 41 (2007)

1004–1009. 5. H. Liu, B.E. Logan, Environ. Sci. Technol. 38 (2004) 4040–4046. 6. W.-W. Li, G.-P. Sheng, X.-W. Liu, H.-Q. Yu, Bioresour. Technol.,

doi:10.1016/j.biortech.2010.03.090. 7. F. Harnisch, U. Schroder, F. Scholz, Environ. Sci. Technol. 42 (2008) 1740–

1746. 8. Y. Zuo, S. Cheng, B.E. Logan, Environ. Sci. Technol. 42 (2008) 6967–6972. 9. R. Rozendal, H.V.M. Hamelers, R.J. Molenkamp, C.J.N. Buisman, Water Res.

41 (2007) 1984–1994.

Page 55: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

42

10. L. Zhuang, C. Feng, S. Zhou, Y. Li, Y. Wang, Process Biochem. 45 (2010) 929–934.

11. F. Zhao, F. Harnisch, U. Schroder, F. Scholz, P. Bogdanoff, I. Herrmann, Environ. Sci. Technol. 40 (2006) 5193–5199.

12. T.H.J.A. Sleutels, H.V.M. Hamelers, R.A. Rozendal, C.J.N. Buisman, Int. J. Hydrogen Energy 34 (2009) 3612–3620.

13. F. Harnisch, U. Schröder, ChemSusChem 2 (2009) 921–926. 14. X. Zhang, S. Cheng, X. Wang, X. Huang, B.E. Logan, Environ. Sci. Technol. 43

(2009) 8456–8461. 15. Y. Fan, H. Hu, H. Liu, J. Power Sources 171 (2007) 348–354. 16. X. Zhang, S. Cheng, X. Huang, B.E. Logan, Biosens. Bioelectron. 25 (2010)

1825–1828. 17. M.R. Hibbs, M.A. Hickner, T.M. Alam, S.K. McIntyre, C.H. Fujimoto, C.J.

Cornelius, Chem. Mater. 20 (2008) 2566–2573. 18. M. Mehanna, T. Saito, J. Yan, M. Hickner, X. Cao, X. Huang, B.E. Logan,

Energy Environ. Sci. 3 (2010) 1114–1120. 19. T. Saito, M.D. Merrill, V.J. Watson, B.E. Logan, M.A. Hickner, Electrochim.

Acta 55 (2010) 3398–3403. 20. T. Saito, T.H. Roberts, T.E. Long, M. Hickner, B.E. Logan, Energy Environ.

Sci., 2010, doi:10.1039/c0ee00229a. 21. S. Cheng, H. Liu, B.E. Logan, Electrochem. Commun. 8 (2006) 489–494. 22. Y. Feng, Q. Yang, X. Wang, B.E. Logan, J. Power Sources 195 (2010) 1841–

1844. 23. O. Bretschger, A. Obraztsova, C.A. Sturm, I.S. Chang, Y.A. Gorby, S.B. Reed,

D.E. Culley, C.L. Reardon, S. Barua, M.F. Romine, J. Zhou, A.S. Beliaev, R. Bouhenni, D. Saffarini, F. Mansfeld, B.-H. Kim, J.K. Fredrickson, K.H. Nealson, Appl. Environ. Microbiol. 73 (2007) 7003–7012.

24. B.E. Logan, P. Aelterman, B. Hamelers, R. Rozendal, U. Schröder, J. Keller, S. Freguiac, W. Verstraete, K. Rabaey, Environ. Sci. Technol. 40 (2006) 5181–5192.

25. S. Cheng, H. Liu, B.E. Logan, Environ. Sci. Technol. 40 (2006) 2426–2432. 26. Z. He, N. Wagner, S.D. Minteer, L.T. Angenent, Environ. Sci. Technol. 40

(2006) 5212–5217.

Page 56: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

43

Chapter 4

Influence of Chemical and Physical Properties of Activated Carbon Powders on Oxygen Reduction Catalysis and Performance in Microbial Fuel Cells

Abstract

Commercially available activated carbon (AC) powders made from different precursor

materials (coal, peat, coconut shell, hardwood, and phenolic resin) were evaluated as oxygen

reduction catalysts, and tested as cathode catalysts in microbial fuel cells (MFCs). Carbons were

characterized in terms of surface chemistry, specific surface area and pore volume distribution,

and kinetic activities were compared to carbon black and platinum catalysts using a rotating disk

electrode (RDE). Cathodes using the coal–derived AC had the highest maximum power densities

in MFCs (1620 ± 10 mW m–2) even though this AC had only average catalytic activity and

selectivity (Eonset = 0.09 V, n = 2.4), and the lowest specific surface area (550 m2 g–1) among these

materials. Peat-based AC performed similarly in MFC tests (1610 ± 100 mW m–2) but had the

best catalyst performance (Eonset = 0.17 V, n = 3.6) in RDE tests and a lower than average specific

surface area (810 m2 g–1). Hardwood based AC had the highest number of acidic surface

functional groups and a higher specific surface area (1010 m2 g–1), but it had the poorest

performance in MFCs and catalysis tests (630 ± 10 mW m–2, Eonset = –0.01V, n = 2.1). There was

a strong inverse relationship between onset potential and the quantity of strong acid (pKa < 8)

functional groups, and a larger fraction of microporosity was negatively correlated with power

production in MFCs. These results show that surface area alone is a poor predictor of catalyst

performance, and that a high quantity of acidic surface functional groups can be detrimental to

oxygen reduction and cathode performance.

Page 57: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

44

4.1 Introduction

Microbial fuel cells (MFCs) are a promising technology for treatment of wastewater

streams in combination with electricity production [1]. MFCs can reduce energy consumption for

wastewater treatment through the elimination of the need for wastewater aeration, and allow for

the utilization of an untapped renewable energy source in the wastewater organic matter. MFCs

consist of a microbe–enriched anode where organic matter is oxidized, and a circuit through

which electrons are conducted to (typically) an air-fed cathode, consisting of a porous carbon

structure and an oxygen reduction catalyst where oxygen is reduced [2]. Power production from

MFCs is often limited by the overpotential of the oxygen reduction reaction (ORR) at the

cathode, and the ORR is negatively impacted by the conditions of neutral pH and ambient

temperature inherent to MFCs. Depending on the catalyst selected, the ORR proceeds through

either a 4e– pathway producing water or hydroxide [3], or a 2e– pathway producing hydrogen

peroxides as an intermediate [4].

To limit the large cathode overpotentials, platinum is often used as a catalyst for oxygen

reduction, but it is an expensive material and a limited resource. Cathode materials account for

47-75% of MFC capital costs [5], and therefore it is important to choose less expensive materials

as the cathode catalyst. Several catalysts have been considered for use in MFCs, including other

metal compounds such as cobalt and iron tetramehoxyphenylporphyrin (TMPP) or

phthalocyanine (Pc) [6, 7] and manganese oxides [8, 9]. Recently, promising results have been

obtained using activated carbon (AC) powder based air-cathodes [10-13]. ACs are especially

interesting as they can be made from many different renewable waste materials such as coconut

shells, wood chips, and sawdust, making them an inexpensive and renewable resource.

AC powder based cathodes have produced power densities by MFCs similar to or slightly

higher than those made with a platinum catalyst. An MFC with an AC cathode made using a

Page 58: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

45

proprietary process, which contained a polytetrafluoroethylene (PTFE) binder and a nickel

current collector, produced 1220 mW m–2, compared to 1060 mW m–2 using a cathode with a Pt

catalyst and a Nafion binder [10]. An MFC incorporating a similar AC cathode structure that had

a polydimethylsiloxane (PDMS) coated cloth diffusion layer reached 1255-1310 mW m–2

compared to 1295 mW m–2 with a standard Pt/C cathode [11]. Using a rolling process to produce

a cathode that consisted of an AC/PTFE layer supported by a stainless steel mesh current

collector produced 1086 mW m–2 in an MFC with a high surface area (1701 m2 g–1) AC, and 1355

mW m–2 with a lower surface area (576 m2 g–1) AC powder. Power production using a standard

Pt/C cathode was not reported [13]. The higher power production by the lower surface area AC

cathode was attributed to a more uniform distribution of microporosity. However, only two ACs

were compared and there was no analysis of AC surface chemistry, which could have affected the

ORR.

The catalytic activity of ACs and other materials can be evaluated independently of mass

transfer limitations using a rotating disk electrode (RDE) and Koutecky-Levich modeling.

Because the method requires only a small sample of material, and it is a relatively quick

comparative analysis that allows the researcher to select promising samples for further testing,

RDE is a common tool used to evaluate catalysts before testing them as fuel cell cathode catalysts

[14]. The evaluation of the material is based on kinetic rates and reaction pathways under non-

mass transfer limited conditions. RDE analyses have been used to study the ORR catalysis of

many materials, including AC powders [13] and other materials such as carbon supported

magnesium oxide nanoparticles [8, 9], and FeTMPP and FePc [7], where the RDE results were

well correlated with MFC performance.

The ORR mechanism at the AC catalyst is not well understood, especially in the neutral

pH and phosphate or carbonate buffered solutions used in many MFC studies. Both physical and

chemical characteristics of the AC catalyst are important to its performance. The AC surface can

Page 59: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

46

contain many different chemical functional groups, and specific surface areas and pore size

distributions vary between different materials. The most common heteroatom found in AC

functional groups is oxygen, which is present in chemical groups that can take on acidic and basic

characteristics. Acid oxygen groups are often analyzed by potentiometric titration (PT) [4, 15-17].

Basic functional groups cannot be determined by PT, and so XPS is often used to detect groups

with a pKa >10 [16].Using these methods can provide insight into the nature and quantity of the

functional groups found on the surface of the ACs and possibly their role in catalysis.

In order to better understand the factors that affect the performance of AC cathodes in

MFCs, nine different ACs made from four different precursor materials were examined as

catalysts for oxygen reduction in terms of kinetics and selectivity (based on number of electrons

transferred) using RDEs, in neutrally buffered solutions. The catalytic rates obtained under these

conditions relevant to MFC operation were then compared in terms of chemical and physical

properties that included relative abundance of acidic oxygen functional groups, specific surface

area, and pore volume distribution. The results of these kinetic and material property analyses

were compared to the power production obtained using different ACs in the cathodes of MFCs.

4.2 Materials and Methods

4.2.1 Catalyst Materials

Nine different samples were chosen to represent a range of physical and chemical

characteristics found in commercially available AC powders. The ACs used were: peat based

carbons, Norit SX1 (P1), SXPlus (P2), and SXUltra (P3) (Norit, USA); coconut shell based

carbons, YP50 (C1) (Kuraray Chemical, Japan), USP8325C (C2) (Carbon Resources, USA),

ACP1250C (C3) (Charcoal House, USA); a hardwood carbon, Nuchar SA-1500 (W1)

Page 60: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

47

(MeadWestvaco, USA); a phenolic resin based carbon, RP20 (R1) (Kurraray Chemical, Japan);

and a bituminous coal carbon, CR325B (B1) (Carbon Resources, USA). The performance of

these carbons in terms of ORR were compared with those of carbon black XC-72 (CB), and

Pt(10%) in carbon black XC-72 (PtC) (Fuel Cell Store, USA).

4.2.2 Physical and Chemical Analyses

Detailed incremental surface area and pore volume distributions were determined for

each AC from argon adsorption isotherms (at 87.3K) determined from progressively increasing

relative pressures of 10–6 to 0.993 atm atm–1 (ASAP2003, Micrometrics Instrument Corp., GA) as

described previously [18]. Pore size distributions were calculated from the isotherms using

Density Functional Theory (DFT) modeling software (Micrometrics Instrument Corp., GA) [18].

Three pore size classifications were used based on the International Union of Pure and Applied

Chemistry (IUPAC) definitions, of micropores <2 nm, mesopores between 2–50 nm, and

macropores >50 nm [19].

The elements present on the surface of the AC powder samples were identified by X-ray

photoelectron spectroscopy (XPS) (Axis Ultra XPS, Kratos Analytical, UK, monochrome AlKα

source, 1486.6eV). CASA XPS software was used for the elemental and peak fitting analysis of

O1s (531–536 eV) and C1s (285–289 eV) signals [16].

Potentiometric titrations were performed using a DL53 automatic titrator (Mettler

Toledo, USA) in the pH range of 3–11, with NaOH (0.1 M) used as the titrant and NaCl (0.01 M)

as the electrolyte. Before titration, samples were adjusted to a ~pH 3 using HCl (0.1 M). Proton

binding isotherms were measured and deconvoluted using the SAIEUS numerical procedure to

obtain the distribution of acidity constants [15, 20, 21]. This analysis produces separate peaks that

denote a difference in type of functional group, with the area under the peak corresponding to the

Page 61: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

48

quantity of functional groups detected in mmol g–1 based on binding/release of protons during

titration.

4.2.3 RDE Analysis

Catalyst ink was prepared by adding 30 mg of the powdered sample (except Pt/C which

was 6 mg to represent the same loading comparison used in the cathodes) to 3 ml of DMF and

homogenized with a sonifier (S-450A, Branson, country) fitted with a 1/8 inch micro tip, pulsed

at 50% for 15 minutes, in an ice bath. Nafion (5 wt% solution, 270 µl) was added and the solution

was mixed for an additional 15 minutes. The ink solution (10 µl) was drop coated onto a 5 mm

diameter glassy carbon disk (Pine Instruments, USA) and allowed to dry overnight. The disk was

prepared before coating by polishing with 5.0 and 0.05 µm alumina paste and cleaned in an

ultrasonic bath for 30 minutes.

All RDE experiments were run first in nitrogen sparged solution, before switching to an

air sparged 100 mM phosphate buffer solution. Solutions were sparged for 30 min before LSVs

were run and then the gas was streamed into the headspace for the duration of the experiment). In

order to clean the electrode surface of possible contaminants or excess oxygen trapped in the

pores of the carbon, the disk potential was cycled between 0.4 and –1.0 V at 100 mV s–1 until the

current response was the same from cycle to cycle. Then, the potential of the disk electrode was

scanned from 0.4 to 1.0 V at 10 mV s–1 and rotation rates of 100 to 2100 rpm. The current

obtained under nitrogen sparging was subtracted from that obtained under air sparging to obtain

the faradaic current attributed to oxygen reduction [14]. Catalyst activity was evaluated by the

onset potential (Eonset) and limiting current (ilim). Kinetic current (ik) and average number of

Page 62: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

49

electrons transferred (n) in the ORR were obtained from the Koutecky-Levich (K-L) analysis

using the following equation.

1𝑖

= 1𝑖𝑘

+ 1𝑖𝑑

= 1𝑛𝐹𝐴𝑘𝐶𝑂2

− 1

0.62𝑛𝐹𝐴𝐷𝑂22 3⁄ 𝜐−1 6⁄ 𝐶𝑂2𝜔

1 2⁄ (4-1)

where i is the measured current, ik the kinetic current, id the diffusion-limiting current, F

Faraday’s constant, A the projected surface area of the disk electrode, k the rate constant, CO2 the

concentration of oxygen in solution, DO2 the diffusion coefficient of oxygen, υ the kinematic

viscosity, and ω the rotation rate of the electrode [9].

4.2.4 MFC Experiments

AC cathodes (31 mg cm–2 loading, projected surface area of 7 cm2) were constructed as

previously described [11], except that two PDMS diffusion layers were applied to the air side of

the stainless steel mesh current collector (50×50 mesh, type 304, McMaster-Carr, OH) prior to

application of the carbons [22]. AC powder was mixed with 10 wt% PTFE binder (in a 60%

emulsion) and spread evenly onto the solution side and pressed at 4.54 metric ton-force for 20

min (Carver press, Model 4386, Carver Inc., IN) [11]. Cathodes were made with carbon black

(XC-72) following the same technique. Platinum-catalyzed air cathodes (projected surface area of

7 cm2) were constructed from carbon cloth (30 wt% wet proofing, Fuel Cell Earth LLC) with four

PTFE diffusion layers, and a catalyst loading of 0.5 mg-Pt cm–2 (on carbon black XC-72) [23].

Cube-shaped MFCs were constructed as previously described [24]. The anode chamber

was a 28 mL cylindrical chamber (7 cm2 cross section) bored into a Lexan block. The anodes

were carbon fiber brushes with a titanium wire core (2.5 cm diameter, 2.5 cm length, and 0.22 m2

surface area) which was heat treated at 450 °C [25] and then placed horizontally in the center of

the cylinder. The electrode spacing was 2.5 cm (center of the anode to the face of the cathode).

The MFCs were inoculated using effluent from an MFC operated under conditions similar to

those used here. The medium in MFC tests was a 100 mM phosphate buffer solution (PBS) (9.13

Page 63: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

50

g L–1 Na2HPO4, 4.90 g L–1 NaH2PO4·H2O, 0.31 g L–1 NH4Cl, and 0.13 g L–1 KCl; pH 7) amended

with vitamins and minerals [26] and 1 g L–1 sodium acetate. Anodes were inoculated and

acclimated under the same conditions in MFCs containing standard Pt/C carbon cloth cathodes,

and then tested with the different cathodes. A 1000 Ω resistor was used during acclimation, and

then the resistance was changed to lower resistance (100 Ω) for several cycles before running

polarization tests to avoid power overshoot [27]. All MFCs were operated at 30 °C in a constant

temperature controlled room. Once the MFC produced a steady voltage for 3 cycles, the Pt/C

cathodes were removed and replaced with the AC cathodes, CB cathodes, or new Pt/C cathodes.

All MFC tests were conducted in duplicate.

The voltage across the resistor was recorded every 30 min using a multimeter (model

2700 Keithley Instruments, Cleveland, OH) with a computerized data acquisition system.

Polarization curves were obtained by applying a different external resistance to the circuit for a

complete batch cycle (multiple cycle method), and the average sustainable voltage was recorded

for each resistance. Current density was calculated from I = E/R, where I is the current, E the

measured voltage, and R the external resistance, and normalized to the projected cathode surface

area. Power densities were calculated using P = IE, and normalized by the projected cathode

surface area [28].

4.3 Results and Discussion

4.3.1 MFC Performance

Based on polarization results, MFCs using cathodes made from different AC powders

had quite different maximum power densities (Figure 4-1A). These ranged from 1620 ± 10 mW

m–2 (0.48 ± 0.00 mA cm–2) using a bituminous coal (B1) to 630 ± 10 mWm–2 (0.36 ± 0.00 mA

cm–2) for the hardwood (W1). One of the peat-based cathodes (P2) had a high power density of

Page 64: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

51

1610 ± 100 mW m–2 that was similar to that obtained with cathodes using AC derived from

bituminous coal (B1). The MFCs using a standard Pt/C cathode all produce a maximum power

density of 2110 ± 0 mW m–2 (0.55 ± 0.00 mA cm–2). The lower maximum power densities

produced by the AC cathodes was due to decreased cathode potentials (increased cathode

overpotential), as the anodes in all of the MFCs maintained similar working potentials (Figure 4-

1B). The best performing MFCs had cathodes with the lowest overpotentials, as seen with the

bituminous sample (B1) operating potential at –0.07 ± 0.00 V (vs. Ag/AgCl) at peak power, while

the hardwood AC (W1) which had the lowest power operated at –0.23 ± 0.01 V (a 230% decrease

in potential). The standard Pt/C cathode potential was –0.01 ± 0.00 V.

Figure 4-1. A) Power density production and B) electrode potentials from polarization of MFCs using AC cathodes compared to Pt/C (100 mM Phosphate Buffer; Open symbols represent cathode potentials, closed symbols are anode potentials).

0

400

800

1200

1600

2000

Pow

er D

ensit

y [m

W/m

2 ]

A

-0.5

-0.4

-0.3

-0.2

-0.1

0.0

0.1

0.2

0.3

0.4

0.0 0.2 0.4 0.6 0.8 1.0 1.2

Elec

trod

e Po

tent

ial [

v]

Current Density [mA/cm2]

Pt/C B1P2 P3P1 C2CB C3C1 R1W1

B

Page 65: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

52

4.3.2 Catalyst Activity and Selectivity

Catalyst performance evaluated using LSV and RDE produced trends in performance

with the different AC precursor materials that generally were similar to those obtained with the

MFCs (Figure 4-2A, Figure A-1). However, the peat-based AC (P2) had the greatest oxygen

reduction activity (Eonset = 0.17 V and ilim = 0.87 mA cm–2 at 2100 rpm), and the bituminous coal

(B1) sample had only average performance (Eonset = 0.09 V and ilim = 0.78 mA cm–2), compared to

MFC tests where the bituminous coal sample produced a higher peak power density than the peat

(P2) sample. It is possible that the differences in MFC and RDE performance were partially due

to the differences in the catalyst layer formation (pressing with PTFE binder versus drop–coating

with Nafion binder); however the point of using the RDE was to test the catalyst performance

without diffusion limitations found in MFC cathode testing. Also, the amount of binder used in

both cases is small, but the large amount of AC used in the MFC cathodes, compared to the thin

layer used in RDE tests, can increase diffusion and electrical conductivity limitations. The

hardwood (W1) AC activity in terms of oxygen reduction again had the worst performance (Eonset

= –0.01 V and ilim = 0.73 mA cm–2), but all AC materials had superior performance to carbon

black (Eonset = –0.06 V and ilim = 0.66 mA cm–2) or the plain glassy carbon disk (Eonset = –0.40 V

and ilim = 0.60 mA cm–2). The AC materials also had less catalytic activity than the Pt/C catalyst

(Eonset = 0.36 V and ilim = 1.11 mA cm–2).

Page 66: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

53

Figure 4-2. A) LSV current response (per disk area) of the AC catalyst at the disk electrode compared to Pt/C and carbon black (100 mM Phosphate Buffer, 2100rpm) B) Average number of electrons (n) transferred (estimated by Koutecky-Levich RDE analysis) during oxygen reduction.

The performance of the catalysts was examined using the K-L analysis to focus on kinetic

current (ik) and the average number of electrons transferred (n) in the ORR without the effects of

diffusion limitation (Figure A-2). ACs that had larger ilim and more positive Eonset also had higher

kinetic current production, where the peat-based (P2) current (ik = 1.1 mA cm–2 at –0.2 V) was

greater than the bituminous coal (B1) (ik = 0.5 mA cm–2) and the hardwood (W1) (ik = 0.4 mA

cm–2) AC catalysts. The selectivity of the catalyst, estimated by the number of electrons

transferred in the ORR, of the peat-based (P2) activated carbon catalyst at –0.2 V (Figure 2B) was

near four electrons with n = 3.6, indicating a mixed reaction that tended toward H2O/OH–

-1.2

-1.0

-0.8

-0.6

-0.4

-0.2

0.0

-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4

Curr

ent D

ensit

y [m

A/cm

2 ]

Disk Potential [V vs. Ag/AgCl]

PtCP2P3P1C1C2C3B1R1W1CBGC

A

1.0

1.5

2.0

2.5

3.0

3.5

4.0

-0.7 -0.6 -0.5 -0.4 -0.3 -0.2 -0.1 0.0 0.1

n

Disk Potential [V vs. Ag/AgCl]

PtC P2P1 P3C3 C1C2 B1R1 W1CB GC

B

Page 67: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

54

formation either through a direct 4e– reduction or through a series of reductions where the

peroxide product was further reduced at the catalyst. The bituminous (B1) (n = 2.4) and

hardwood (W1) (n = 2.1) based samples were closer to a 2e– reduction, where the reduction

product was mostly through peroxide formation without further catalytic reduction.[4].

4.3.3 Effect of Oxygen Functional Groups on ORR Catalysis

The prevalence and variety of acidic functional groups on the surface of the AC samples,

determined by potentiometric titration analysis, influenced their catalytic activity for reduction of

oxygen. ACs made from the same or similar precursor materials had acidic functional groups

with similar pKa values (Figure 4-3, Figure A-3). The two best performing AC samples peat (P2)

and bituminous coal based (B1) had very similar acidic functional groups present, with pKa

values around 4.5, 6.5, 8.5 and 10 (Figure 4-3A). The quantity (mmol g–1) of acidic functional

groups present on the surface of the AC samples, estimated by the area under the curve at each

pKa , varied for the different carbons There was a greater quantity of strong acid groups (pKa < 8,

typically attributed to carboxyl goups) for the bituminous coal sample (B1, 0.17 mmol g–1) than

the three peat based samples (P1, 0.06 mmol g–1; P2, 0.05 mmol g–1; P3, 0.06 mmol g–1). The ACs

made from the other precursor materials showed a larger variety of acidic functional groups

(Figure 4-3B), with the hardwood sample (W1, 0.36 mmol g–1) having the largest quantity of

strong acid functional groups. This suggests that the AC surface chemistry, based on the quantity

of strong acid groups, had a strong influence on the activity of the AC catalyst for ORR (Figure

4-4). An increase in strong acid functional groups led to a decrease in the onset potential of

oxygen reduction (p = 0.00003). However, the presence of strong acid functional groups did not

correlate as well with the power densities produced when these carbons were used in the cathodes

of the MFCs (Figure A-4).

Page 68: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

55

Figure 4-3. Acidic/Oxygen functional groups determined by potentiometric titration. A) Bituminous and peat based activated carbon samples have similar functional groups. B) Other activated carbons have a larger variety of acidic groups.

Figure 4-4. The onset potential of the oxygen reduction reaction is inversely related to the amount of strong acid functional groups present on the activated carbons tested.

0

0.1

0.2

0.3

0.4

0.5

F(pK

) [m

mol

/g]

P1P2P3B1

A

0

0.1

0.2

0.3

0.4

0.5

3 4 5 6 7 8 9 10 11

F(pK

) [m

mol

/g]

pKa

C1C2C3R1W1

B

P2P3P1

B1

W1

C2C1

C3R1

y = -1.7x + 0.3

0

0.1

0.2

0.3

0.4

-0.05 0 0.05 0.1 0.15 0.2

Stro

ng A

cid

Grou

ps (p

K a<8

) [m

mol

/g]

Eonset [V vs. Ag/AgCl]

P-value=3x10-5

Page 69: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

56

4.3.4 Effect of Microporosity on Power Production

The ACs had cumulative surface areas ranging from 550 m2 g–1 (B1) to 1440 m2 g–1 (R1),

and cumulative pore volumes between 0.3 mL g–1 (B1) and 1.1 mL g–1 (W1) (Figure 4-5). A

majority of the surface area was attributed to micropores (Figure 4-5A). With the exception of the

hardwood AC (W1), there was a strong inverse relationship between the surface area of the AC

powder and the maximum power density achieved in the MFCs. The bituminous sample (B1),

which had the least surface area, produced the highest power density in MFC tests, and the

phenolic resin sample (R1), which had the most surface area, produced one of the lowest power

densities. A similar trend was observed with micropore volume (Figure 5B), where the power

density increased inversely with the micropore volume of the carbon. Most likely the micropores

hinder diffusion of the reactants to the catalytic functional sites on the activated carbon, as well as

the diffusion of the reduction product from the pores, thereby negatively impacting the

favorability of the reaction. The negative effect of increased microporosity of the AC powders on

power production explains the higher power production of the MFC using the bituminous based

cathode (B1) despite the average catalytic performance of the bituminous based AC in the RDE

analysis. In the case of the hardwood (W1) based cathode, the lower microporosity did not

compensate for the poor intrinsic catalytic performance of the hardwood based AC powder based

on RDE tests.

Page 70: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

57

Figure 4-5. Maximum power density (per m2 projected cathode surface) of the MFCs using the activated carbon cathodes is inversely related to the A) surface area (without W1 pvalue=0.0006) and B) micropore volume (without W1 pvalue=0.0011) of the powdered carbons with the exception of sample W1.

4.3.5 Functional Group Analysis Using XPS

The quantity of acidic functional groups measured by potentiometric titration is

commonly attributed to oxygen containing groups [15]. Although the onset potential was

inversely correlated to the quantity of strong acid/oxygen containing functional groups detected,

neither the onset potential nor the power density obtained were correlated with the total atomic

percent of oxygen present on the AC surface as determined by XPS (Figure 4-6A). For instance,

the bituminous based AC (B1) had the most oxygen atoms present (9.4%), but only a moderate

0

300

600

900

1200

1500

1800

0

300

600

900

1200

1500

1800

R1 C1 C2 C3 W1 P3 P1 P2 B1

Pow

er D

ensit

y [m

W/m

2 ]

Cum

ulat

ive

Surf

ace

Area

[m2 /

g]

meso micro PowerA

0

300

600

900

1200

1500

1800

0.0

0.2

0.4

0.6

0.8

1.0

1.2

R1 C1 C3 C2 W1 P3 P1 P2 B1

Pow

er D

ensit

y [m

W/m

2 ]

Cum

ulat

ive

Pore

Vol

ume

[ml/g

]

B

Page 71: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

58

abundance of strong acid functional groups, and average ORR catalytic activity. The chemical

state of the oxygen present varied between the samples (Figure 4-6B), where the bituminous

based (B1) sample had a larger amount of oxygen present in the adsorbed O2/H2O region (BE =

536 – 536.6 eV) [16]. This suggests that not all oxygen functional groups (e.g. quinones) are

detrimental to oxygen reduction catalysis, in agreement with AC ORR catalysis studies using

acidic or alkaline media [29]. Trace amounts of nitrogen were found in the Peat and Bituminous

samples, with no detectable nitrogen in the remaining samples.

Figure 4-6. A) Maximum power density (normalized to cathode surface area) of MFCs with activated carbon cathodes are not directly related to oxygen content of activated carbon powders determined by XPS. B) Chemical state of oxygen detected by XPS varies for activated carbon powders tested (e.g. increased signal in the adsorbed O2/H2O region for sample B1.)

0

500

1000

1500

2000

2500

3000

3500

4000

525530535540545

CPS

Binding Energy [eV]

C3R1C1B1C2W1P1P2P3

B

0

300

600

900

1200

1500

1800

P1 P2 P3 C2 C1 R1 W1 C3 B10

2

4

6

8

10

Pow

er D

ensit

y [m

W/m

2 ]

Atom

ic %

O2

A

Page 72: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

59

4.3.6 Implications of AC properties for MFC performance

Both the surface chemistry and the pore structure of the AC catalyst affected performance

of the catalyst in the cathode of an MFC. AC catalysts selected for neutral buffered environments,

like those in MFCs, should have less acidic surface functional groups, which can hinder the ORR

activity. The effect of basic (rather than acidic) functional groups on ORR catalysis should be

further investigated. ACs used in MFC cathodes should therefore not be chosen solely because

they have the largest surface area, but instead ACs should be selected that have a moderate

amount of micropore volume and surface area to avoid the negative impact of diffusion

limitations to the active catalyst sites. Binders, AC loading, and manufacturing methods can also

affect the diffusion characteristics of the cathodes and their performance relative to standard Pt/C

cathodes [11, 13], and therefore this aspect of cathode construction should also be taken into

consideration when constructing MFC cathodes. Longevity of AC catalysts in MFC cathodes is

also an issue for MFC applications [30], and therefore changes in surface chemistry and rates of

mass transfer to catalytic sites should also be considered in future studies.

4.4 Acknowledgements

The authors thank Dr. Cesar Nieto Delgado for assistance with the potentiometric

titration and Vince Bojan for assistance with XPS analysis. The authors acknowledge support

from the King Abdullah University of Science and Technology (KAUST) by Award KUS-I1-

003-13 and the National Science Foundation Graduate Research Fellowship Program (NSF-

GRFP).

Page 73: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

60

4.5 References

1. Logan, B. and K. Rabaey, Conversion of Wastes into Bioelectricity and Chemicals by Using Microbial Electrochemical Technologies. Science, 2012. 337: p. 686-690.

2. Logan, B.E., Microbial Fuel Cells. 2008, New York: John Wiley & Sons. 3. Popat, S., et al., Importance of OH- Transport from Cathodes in Microbial Fuel Cells.

ChemSusChem, 2012. 5: p. 1071-1079. 4. Zhong, R.-S., et al., Effect of carbon nanofiber surface functional groups on oxygen

reduction in alkaline solution. Journal of Power Sources, 2013. 225: p. 192-199. 5. Rozendal, R.A., et al., Towards practical implementation of bioelectrochemical

wastewater treatment. Trends in Biotechnology, 2008. 26(8): p. 450-459. 6. Yu, E.H., et al., Microbial Fuel Cell Performance with non-Pt Cathode Catalysts. Journal

of Power Sources, 2007. 171: p. 275-281. 7. Birry, L., et al., Application of iron-based cathode catalysts in a microbial fuel cell.

Electrochimica Acta, 2011. 56: p. 1505-1511. 8. Roche, I. and K. Scott, Carbon-supported manganese oxide nanoparticles as

electrocatalysts for oxygen reduction reaction (orr) in neutral solution. Journal of Applied Electrochemistry, 2009. 39: p. 197-204.

9. Chen, Y., et al., Stainless steel mesh coated with MnO2/carbon nanotube and polymethlyphenyl siloxane as low-cost and high-performance microbial fuel cell cathode materials. Journal of Power Sources, 2012. 201: p. 136-141.

10. Zhang, F., et al., Power generation using an activated carbon and metal mesh cathode in a microbial fuel cell. Electrochemistry Communications, 2009. 11(11): p. 2177-2179.

11. Wei, B., et al., Development and evaluation of carbon and binder loading in low-cost activated carbon cathodes for air-cathode microbial fuel cells. RSC Advances, 2012. 2: p. 12751-12758.

12. Dong, H., et al., A novel structure of scalable air-cathode without Nafion and Pt by rolling activated carbon and PTFE as catalyst layer in microbial fuel cells. Water Res, 2012. 46: p. 5777-5787.

13. Dong, H., H. Yu, and X. Wang, Catalysis kinetics and porous analysis of rolling activated carbon-PTFE air-cathode in microbial fuel cells. Environ Sci Technol, 2012. 46: p. 13009-13015.

14. Gojkovic, S.L., S. Gupta, and R.F. Savinell, Heat-treated iron(III) tetramethoxyphenyl porphyrin supported on high-area carbon as an electrocatalyst for oxygen reduction - I. Characterization of the electrocatalyst. Journal of the Electrochemical Society, 1998. 145: p. 3493-3499.

15. Bandosz, T.J., J. Jagiello, and C. Contescu, Characterization of the surfaces of activated carbons in terms of their acidity constant distributions. Carbon, 1993. 31(7): p. 1193-1202.

16. Seredych, M. and T.J. Bandosz, Investigation of the enhancing effects of sulfur and/or oxygen functional groups of nanoporous carbons on adsorption of dibenzothiophenes. Carbon, 2011. 49: p. 1216-1224.

17. Szymanski, G.S., et al., The effect of the gradual thermal decomposition of surface oxygen species on the chemical and catalytic properties of oxidized activated carbon. Carbon, 2002. 40: p. 2627-2639.

Page 74: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

61

18. Moore, B.C., et al., Changes in GAC pore structure during full-scale water treatment at Cincinnati: a comparison between virgin and thermally reactivated GAC. Carbon, 2001. 39: p. 789-807.

19. Rouquerol, J., et al., Recommendations for the characterization of porous solids. Pure and Applied Chemistry, 1994. 66(8): p. 1739-1758.

20. Jagiello, J., Stable numerical solution of the adsorption integral equation using splines. Langmuir, 1994. 10(8): p. 2778-2785.

21. Bandosz, T.J., et al., Efffect of surface chemisty on sorption of water and methanol on activated carbons. Langmuir, 1996. 12: p. 6480-6486.

22. Zhang, X., et al., Improved performance of single-chamber microbial fuel cells through control of membrane deformation. Biosensors and Bioelectronics, 2010. 25: p. 1825-1828.

23. Cheng, S., H. Liu, and B.E. Logan, Increased performance of single-chamber microbial fuel cells using an improved cathode structure. Electrochem Commun, 2006. 8: p. 489-494.

24. Liu, H., R. Ramnarayanan, and B.E. Logan, Production of electricity during wastewater treatment using a single chamber microbial fuel cell. Environmental Science & Technology, 2004. 38(7): p. 2281-2285.

25. Feng, Y., et al., Treatment of carbon fiber brush anodes for improving power generation in air–cathode microbial fuel cells. Journal of Power Sources, 2010. 195: p. 1841 - 1844.

26. Bretschger, O., et al., Current production and metal oxide reduction by Shewanella oneidensis MR-1 wild type and mutants. Appl Environ Microbiol, 2007. 73(21): p. 7003-7012.

27. Hong, Y., et al., Adaptation to high current using low external resistances eliminates power overshoot in microbial fuel cells. Biosens Bioelectron, 2011. 28(1): p. 71-76.

28. Logan, B.E., et al., Microbial fuel cells: methodology and technology. Environ Sci Technol, 2006. 40(17): p. 5181-5192.

29. Zhang, J., ed. PEM Fuel Cell Electrocatalysts and Catalyst Layers: Fundamentals and Applications. Vol. XXII. 2008. 489.

30. Zhang, F., D. Pant, and B.E. Logan, Long-term performance of activated carbon air cathodes with different diffusion layer porosities in microbial fuel cells. Biosens Bioelectron, 2011. 30(1): p. 49-55.

Page 75: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

62

Chapter 5

Improvement of Oxygen Reduction Catalysis in Neutral Solutions using Ammonia Treated Activated Carbons and Performance in Microbial Fuel

Cells

Abstract

Commercially available activated carbon (AC) powders made from different precursor

materials (peat, coconut shell, coal, and hardwood) were treated with ammonia gas at 700 °C and

evaluated as oxygen reduction catalysts in neutral pH phosphate buffer for application in

microbial fuel cell (MFC) cathodes. Ammonia treatment resulted in a decrease in oxygen (by 29

– 58%) and an increase in nitrogen content (total abundance up to 1.8 atomic %) on the carbon

surfaces, which also resulted in an increase in the basicity of the bituminous, peat, and hardwood

ACs. The kinetic activity and selectivity of ammonia treated carbons were evaluated using a

rotating ring-disk electrode (RRDE) and compared to untreated ACs and platinum. All of the

ammonia treated ACs exhibited better catalytic performance than their untreated precursors, with

the bituminous (treated, Eonset = 0.12 V, n = 3.9; untreated, Eonset = 0.08 V, n = 3.6) and hardwood

(treated, Eonset = 0.03 V, n = 3.3; untreated, Eonset = –0.04 V, n = 3.0) based samples showing the

most improvement. Cathodes using the ammonia treated coal based AC had the highest maximum

power densities in MFCs (2450 ± 40 mW m–2). Even though the ammonia treated peat based AC

had the greatest ORR activity in RRDE testing, the untreated sample had higher power

production in the MFC tests (2360 ± 230 mW m–2). The treated coconut and hardwood derived

ACs outperformed the untreated precursor ACs in both electrochemical and MFC testing. These

results show that reduction in oxygen abundance and increase in nitrogen functionalities on the

surface of ACs can increase the catalytic performance for oxygen reduction in neutral media.

Page 76: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

63

5.1 Introduction

Microbial fuel cells (MFCs) are a promising option for reduction of energy costs

associated with the treatment of wastewater sources [1]. Power production from MFCs is limited

by the overpotential of the oxygen reduction reaction (ORR) at the cathode, which is negatively

impacted by the conditions of neutral pH and ambient temperature common in MFCs. Depending

on the catalyst properties, the ORR can proceed through either a 4e– pathway producing water or

hydroxide [2], or 2e– pathway producing hydrogen peroxide as an intermediate [3]. Peroxides can

be further reduced through an additional 2e– reduction step, resulting in a mixed reduction

pathway that can approach an apparent four electron transfer to the cathode [3].

In order to achieve commercial viability, low cost materials are essential to the success of

MFC technology. Activated carbons (ACs) are inexpensive ORR catalysts that can be made from

several biomass waste streams such as coconut shells, wood chips and sawdust. They have a

complex surface chemistry that can be tailored to improve their performance for the desired

application. AC powder based cathodes have produced power densities in MFCs similar to or

slightly higher than those made with a typical platinum catalyst. An MFC with an AC cathode,

that consisted of a polytetrafluoroethylene (PTFE) binder and a nickel current collector, produced

1220 mW m–2, compared to 1060 mW m–2 using a cathode with a Pt catalyst and a Nafion binder

[4]. An MFC incorporating a similar AC cathode structure with a polydimethylsiloxane (PDMS)

coated cloth diffusion layer reached 1255-1310 mW m–2 compared to 1295 mW m–2 with a

standard Pt/C cathode [5]. Another type of AC air cathode made by rolling out an AC/PTFE

layer on a stainless steel mesh current collector produced between 1086 – 1355 mW m–2 using

two different types of AC powders. Power production using a standard Pt/C cathode was not

reported [6]. In the research reported in Chapter 4, a greater abundance of strong acid functional

Page 77: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

64

groups of the surface of AC powders was found to be detrimental to the ORR catalytic activity in

MFC environments.

Several studies have shown that the number of nitrogen functional groups on carbon

surfaces can be increased by treatment with ammonia gas at elevated temperatures [3, 7-9].

During the process of incorporating the nitrogen into the carbon structure, there is a

corresponding reduction in acidic oxygen groups as the oxygen atoms are desorbed from the

carbon surface as CO/CO2. This rearrangement of surface functional groups results in an increase

in the basic properties of the carbon surface at the expense of acidic properties [9-14]. Nitrogen

incorporation on carbon surfaces has been shown to increase the catalytic activity and selectivity

for oxygen reduction through a four electron pathway in both acidic and alkaline environments,

but it has not been examined under neutral pH conditions [3, 15, 16].

In order to improve the performance of AC cathodes in MFCs, ACs made from four

different precursor materials were treated with ammonia gas (5%, balance He) at 700 °C and

examined as catalysts for oxygen reduction in neutrally buffered solution. Treated ACs were

evaluated in terms of activity and selectivity using a rotating ring-disk electrode (RRDE) and

linear sweep voltammetry (LSV). The catalytic activity observed, in conditions relevant to MFC

operation (pH 7, 30 °C), was then compared to the change in surface chemistry that included the

relative abundance of surface oxygen and nitrogen functional groups. The results of the kinetic

and chemical property analyses were compared to the power production produced using the

ammonia treated ACs in the cathodes of MFCs.

Page 78: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

65

5.2 Materials and Methods

5.2.1 Activated Carbons and Ammonia Treatment

Four AC samples that were previously studied for ORR catalysis (Chapter 4) were

chosen to represent a range of physical and chemical characteristics found in commercially

available AC powders. The ACs used were: a peat based carbon, Norit SXPlus (P, Norit, USA); a

coconut shell based carbon, YP50 (C, Kuraray Chemical, Japan); a hardwood carbon, Nuchar

SA-1500 (W1, MeadWestvaco, USA); and a bituminous coal carbon, CR325B (B1, Carbon

Resources, USA). The base ACs were treated with ammonia gas (5% in helium) at 700 °C using a

vertical cylindrical glass tube reactor in a programmable furnace (model 3210, Applied Test

Systems, Inc., Butler, PA). Before heating, the furnace (including sample) was purged with ultra

pure nitrogen gas for 30 min. Gas flow was changed to dilute ammonia while temperature was

ramped at 5 °C min–1, then held at 700 °C for 1 hr. The furnace and sample were purged with

ultra pure nitrogen gas while cooling to ambient temperature. The performance of these carbons

(denoted as –N) in terms of ORR catalysis were compared to the base AC samples, as well as

carbon black XC–72 (CB), and Pt (10%) in carbon black XC–72 (PtC) (Fuel Cell Store, USA).

5.2.2 Chemical Surface Analysis

The elements present on the surface of the AC powder samples were identified by X-ray

photoelectron spectroscopy (XPS, Axis Ultra XPS, Kratos Analytical, UK, monochrome AlKα

source, 1486.6eV). A base survey scan was performed first, followed by a detailed scan of C1s

(285–289 eV), O1s (531–536 eV), and N1s (398–406 eV) signals [17]. CASA XPS software was

used for the elemental and peak fitting analysis.

Page 79: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

66

Potentiometric titrations were performed using a DL53 automatic titrator (Mettler

Toledo, USA) in the pH range of 3–11, with NaOH (0.1 M) used as the titrant and NaCl (0.01 M)

as the electrolyte. Before titration, samples were adjusted to a ~pH 3 using HCl (0.1 M). Proton

binding isotherms were measured and deconvoluted using the SAIEUS numerical procedure to

obtain the distribution of acidity constants [18-20]. This analysis produces separate peaks that

denote a difference in type of functional group, with the area under the peak corresponding to the

quantity of functional groups detected in mmol g–1 based on binding/release of protons during

titration.

5.2.3 Rotating Ring-Disk Electrochemical Analysis

Catalyst ink was prepared by adding 30 mg of the powdered sample to 3 ml of DMF and

homogenized with a sonifier (S-450A, Branson, country) fitted with a 1/8 inch micro tip, pulsed

at 50% for 15 minutes, in an ice bath. Nafion (270 µL; 5 wt% solution) was added and the

solution was mixed for an additional 15 minutes. The ink solution (10 µL) was drop coated onto a

5 mm diameter glassy carbon disk (Pine Instruments, USA) and allowed to dry overnight. The

disk was prepared before coating by polishing with 5.0 and 0.05 µm alumina paste and cleaned in

an ultrasonic bath for 30 minutes.

All RRDE experiments were run first in nitrogen sparged solution in order to obtain the

baseline current, before switching to an air sparged 100 mM phosphate buffer solution. Solutions

were sparged for 30 min before LSVs were run, and then the gas was streamed into the headspace

for the duration of the experiment. In order to clean the electrode surface of possible

contaminants or excess oxygen trapped in the pores of the carbon, the disk potential was cycled

between 0.4 and 1.0 V at 100 mV s–1 until a consistent current response was observed from one

cycle to the next. All potentials are reported vs. 3M Ag/AgCl reference electrodes (0.197 V vs.

Page 80: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

67

SHE). The potential of the disk electrode was then scanned from 0.4 to –1.0 V at 10 mV s–1 and

rotation rates of 100 to 2100 rpm, while the potential of the platinum ring was held constant at

0.62 V for H2O2 oxidation. The current obtained under nitrogen sparging was subtracted from that

obtained under air sparging to obtain the faradaic current attributed to oxygen reduction [21].

Catalyst activity was evaluated by the onset potential (Eonset) and limiting current (ilim) [21]. The

average number of electrons transferred (n) in the ORR at the disk electrode was calculated based

on the amount of H2O2 detected using [22]

𝑛 = 4𝑖𝑑𝑖𝑠𝑘𝑖𝑑𝑖𝑠𝑘+𝑖𝑟𝑖𝑛𝑔 𝑁⁄

(5–1)

where idisk is reduction current at the disk, iring the oxidation current at the ring, and N is the

collection efficiency of the RRDE.

5.2.4 MFC Experiments

AC cathodes (31 mg cm–2 loading, projected surface area of 7 cm2) were constructed as

previously described [5], except that the diffusion layer consisted of two PDMS layers that were

applied to the air side of the stainless steel mesh current collector (50×50 mesh, type 304,

McMaster-Carr, OH) prior to application of the carbons [23]. In addition, a second stainless steel

mesh current collector was pressed onto the solution side of the activated carbon to improve

electrical conductivity of the cathode. AC powder was mixed with 10 wt% PTFE binder (in a

60% emulsion) and spread evenly onto the solution side of the PDMS coated mesh, and pressed

at 4.54 metric ton-force for 20 min (Carver press, Model 4386, Carver Inc., IN) [5]. Pt-catalyzed

air cathodes (projected surface area of 7 cm2) were constructed from carbon cloth (30 wt% wet

proofing, Fuel Cell Earth LLC) with four PTFE diffusion layers, and a catalyst loading of 0.5 mg-

Page 81: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

68

Pt cm–2 (on carbon black XC-72) [24] to benchmark AC cathodes against this commonly used

cathode.

Cube-shaped MFCs were constructed as previously described [25]. The anode chamber

was a 28 mL cylindrical chamber (7 cm2 cross section) bored into a Lexan block. The anodes

were carbon fiber brushes with a titanium wire core (2.5 cm diameter, 2.5 cm length, and 0.22 m2

surface area) which was heat treated at 450 °C [26] and then placed horizontally in the center of

the cylinder. The electrode spacing was 2.5 cm (from center of the anode to the cathode). The

MFCs were inoculated using effluent from an MFC operated under conditions similar to those

used here. The medium in MFC tests was a 100 mM phosphate buffer solution (PBS) (9.13 g L–1

Na2HPO4, 4.90 g L–1 NaH2PO4·H2O, 0.31 g L–1 NH4Cl, and 0.13 g L–1 KCl; pH 7) amended with

vitamins and minerals [27] and 1 g L–1 sodium acetate. Anodes were inoculated and acclimated

under the same conditions in MFCs containing standard Pt/C carbon cloth cathodes, and then

tested with the different cathodes. A 1000 Ω resistor was used during acclimation, and then the

resistance was changed to lower resistance (100 Ω) for several cycles before running polarization

tests to avoid power overshoot [28]. All MFCs were operated at 30 °C in a constant temperature

controlled room. Once the MFC produced a steady voltage for 3 cycles, the Pt/C cathodes were

removed and replaced with the AC cathodes, CB cathodes, or new Pt/C cathodes. All MFC tests

were conducted in duplicate.

The voltage across the resistor was recorded every 30 min using a multimeter (model

2700 Keithley Instruments, Cleveland, OH) with a computerized data acquisition system.

Polarization curves were obtained by applying a different external resistance to the circuit for a

complete batch cycle (multiple cycle method), and the average sustainable voltage was recorded

for each resistance. Current density was calculated from I = E/R, where I is the current, E the

measured voltage, and R the external resistance, and normalized to the projected cathode surface

Page 82: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

69

area. Power densities were calculated using P = IE, and normalized by the projected cathode

surface area [29].

5.3 Results and Discussion

5.3.1 MFC performance

In general, cathodes made from ammonia treated AC powders had higher maximum

power densities in MFCs than the cathodes made with the corresponding untreated AC (Figure 5-

1A). The highest power was obtained using the treated bituminous coal AC (B–N) cathode (2450

± 43 mW m–2; 0.83 ± 0.01 mA cm–2), which was a 28% increase in power compared to the MFC

with the untreated bituminous AC (B) cathode (1910 ± 188 mW m–2; 0.73 ± 0.04 mA cm–2), and a

16% improvement in maximum power compared to a standard Pt/C cathode (2100 ± 1 mW m–2;

0.55 ± 0.01 mA cm–2). Treating the hardwood AC powder (W–N) produced the largest

improvement, with 53% higher power production than the untreated hardwood (W). Ammonia

treatment of the coconut AC (C–N) resulted in a 29% increase in maximum power. The peat-

based cathodes were the exception to improved power with treatment, as the untreated peat AC

(P) cathode had a higher power density of 2360 ± 230 mW m–2 (0.81 ± 0.04 mA cm–2) than the

treated AC (P–N) cathode (1860 ± 84 mW m–2; 0.72 ± 0.01 mA cm–2).

Page 83: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

70

Figure 5-1. A) Power density production and B) electrode potential during cell polarization of MFCs using AC cathodes compared to Pt/C. (100 mM Phosphate Buffer; open symbols indicate cathode potentials, closed symbols anode potentials.)

Power and current densities of the untreated base ACs obtained here were higher than

those in the previous study (Chapter 4) due to the addition of a second stainless steel mesh current

collector on the solution side of the cathode (Figure B-1A). These increases in maximum power

densities resulted from increased operating potentials of the cathodes (decreased cathode

overpotential), since the anodes in all of the MFCs operated at similar potentials (Figure 5-1B,

0

500

1000

1500

2000

2500

Pow

er D

ensit

y [m

W m

-2]

A

-0.5

-0.4

-0.3

-0.2

-0.1

0.0

0.1

0.2

0.0 0.2 0.4 0.6 0.8 1.0 1.2

Elec

trod

e Po

tent

ial [

V]

Current Density [mA cm-2]

P-N PB-N BW-N WC-N CPtC

B

Page 84: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

71

Figure B-1B). MFCs that produced the most power at a given current density had cathodes with

the lowest overpotentials, such as the treated bituminous sample (B–N) (–0.05 ± 0.00 V at peak

power), which had a 43% increase in working potential compared to the untreated AC.

5.3.2 Catalyst Activity and Selectivity

Ammonia treatment of the ACs improved catalytic activity for oxygen reduction in all

samples based on evaluation using LSV and RRDE (Figure 5- 2). Despite the reduced power

production observed in the MFC cathode tests, the ammonia treated peat-based AC (P–N) had the

greatest oxygen reduction activity (Eonset = 0.16 V and ilim = 0.96 mA cm–2 at 2100 rpm), and was

improved from the untreated peat AC performance (Eonset = 0.14 V and ilim = 0.83 mA cm–2). The

other treated ACs had catalytic improvements that aligned with the increased power production

observed in MFC tests. The treated bituminous coal (B–N) sample had reduced overpotential with

a similar limiting current (Eonset = 0.12 V and ilim = 0.78 mA cm–2) compared to the untreated

sample (B) (Eonset = 0.08 V and ilim = 0.76 mA cm–2). The activity of the hardwood AC sample

improved after ammonia treatment with both a reduction in overpotential and an increase in

oxygen reduction current activity (W–N, Eonset = 0.03 V and ilim = 0.82 mA cm–2; W, Eonset = –0.04

V and ilim = 0.67 mA cm–2). Even though the peat based (P) and ammonia treated bituminous (B–

N) AC achieved higher power densities than standard Pt/C cathodes in MFC tests, all of the AC

materials had less catalytic activity than the Pt/C catalyst in RDDE tests (Eonset = 0.36 V and ilim =

1.28 mA cm–2), suggesting that mass transfer to the catalyst material is important for MFC

performance.

Page 85: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

72

Figure 5-2. A) H2O2 detection based on oxidation current at the Pt ring during oxygen reduction at the catalyst on the disk electrode. B) Oxygen reduction current response during LSV of AC catalysts at the disk electrode compared to Pt/C (100 mM Phosphate Buffer, 2100 rpm). C) Average number of electrons transferred (measured by RRDE analysis) during oxygen reduction.

0.000

0.002

0.004

0.006

0.008

0.010

I ring

(mA)

WW-NCC-NPP-NBB-NCBGCPtC

A

2.0

2.5

3.0

3.5

4.0

-1 -0.8 -0.6 -0.4 -0.2 0 0.2

n

Disk Potential [V vs Ag/AgCl]

PtCP-NPB-NBC-NCW-NWCBGC

C

-1.4

-1.2

-1.0

-0.8

-0.6

-0.4

-0.2

0.0

-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4

J disk

(mA

cm-2

)

Disk Potential (V vs Ag/AgCl)

B

Page 86: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

73

The selectivity of the catalysts for a complete four electron reduction (and limited H2O2

production) was evaluated using RRDE, and the average number of electrons transferred (n) was

calculated using an empirical collection efficiency of 0.2 (Figure 5-2C). The selectivity of the

catalyst improved after ammonia treatment for all of the samples tested. The treated bituminous

(B–N) AC had the most consistent, near complete reduction of oxygen over the range of

potentials with an average n = 3.85. The treated peat-based (P–N) AC catalyst was also near four

electrons with n = 3.9 at –0.2 V, but this tapered off to n = 3.5 as the disk potential was decreased

to –1 V. Peroxide formation was detected by the platinum ring electrode for all of the ACs tested.

This indicated that there was a mixed reaction pathway with some catalytic sites possibly

reducing oxygen to H2O/OH– through a direct four electron reduction, with others sites reducing

oxygen to H2O2 through a two electron transfer route, and then a second reduction pathway where

the peroxide product was further reduced to H2O/OH– by an additional two electron reduction [3].

5.3.3 Effect of Surface Chemistry

Ammonia treatment of the AC samples increased the basicity of the carbon surface for

the bituminous, peat, and hardwood based samples (Figure 5-3) as expected based on previous

studies [9-14]. PT analysis of the coconut based sample did not show a noticeable change in basic

properties. Acidic oxygen functional groups could not be reliably quantified using this

deconvolution of the proton binding isotherm because some nitrogen functional groups have pKas

~ 4.5 and 9 [14]. However, the measured increase in the basicity is evidence that the heat

treatment increased the relative amount of basic to acidic groups on the carbon surface.

Page 87: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

74

Figure 5-3. Proton binding isotherms for treated and untreated A) bituminous, B) peat, C) coconut shell, and D) hardwood based activated carbons.

Based on the XPS analysis, the ammonia treatment process successfully increased the

amount of nitrogen groups present on the surface of the ACs to a measureable level (Figure 5-4).

The hardwood based AC incorporated the most nitrogen, and the type of groups present were

similar among all of the treated sample types. From the quantitative analysis (Figure 5-5),

ammonia treatment reduced the relative abundance oxygen atoms by between 29 to 58% on the

surface of all ACs, with the peat and hardwood losing the largest percent and the coconut shell

AC losing the least. The corresponding gain in surface nitrogen groups resulted in nitrogen levels

between 0.9 and 1.8 atomic %. The decrease in relative abundance along with the increase in

-0.1

0.1

0.3

0.5

0.7

0.9

H+Bo

und

(Q) [

mm

ol/g

]

B-NB

A P-NP

B

3 5 7 9 11pH

W-NW

D

-0.6

-0.4

-0.2

0.0

0.2

0.4

0.6

3 5 7 9 11

H+Bo

und

(Q) [

mm

ol/g

]

pH

C-NC

C

Page 88: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

75

surface nitrogen groups resulted in the increase in catalytic activity for oxygen reduction

observed through the increase in onset potential, current response, and increased electron transfer

numbers nearing a four electron reduction pathway. These results are similar to those reported

for treated carbon nanofibers tested in an alkaline solution [3], where oxygen reduction at the

electropositive carbon sites adjacent to nitrogen groups follow the Yeager Model which leads to a

direct four electron reduction. This is in contrast to the Pauling model where end-on oxygen

adsorption to the electropositive site at the carbonyl group leads to a two electron transfer and

peroxide production, which can be further reduced through a sequential two electron pathway on

any oxygen functional group [3].

Figure 5-4. N1s peaks on ammonia treated ACs from XPS show presence of nitrogen groups on the surface of the treated AC.

0

200

400

600

800

1000

1200

1400

395400405410

CPS

Binding Energy [eV]

W-NC-NP-NB-N

Page 89: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

76

Figure 5-5. Atomic % of oxygen and nitrogen on the surface of treated and untreated AC catalysts measured using XPS and the relationship to onset potential of the ORR measured with RRDE.

5.4 Conclusion

Ammonia treatment of AC powders resulted in an increase in ORR catalytic activity and

selectivity in a neutral phosphate buffer solution due to an increase in nitrogen and decrease in

acidic oxygen surface functional groups. Selection of an ideal AC catalyst for neutral buffered

environments, like those in MFCs, should therefore focus on the presence of surface nitrogen

groups. Stability of the treated AC catalysts should be explored in MFC applications, since

increasing the basicity of AC surfaces has been shown to increase the adsorption of dissolved

organic matter and other contaminants commonly found in wastewater. Adsorption of these

molecules may block active surface sites, which can interfere with ORR catalysis [10, 13].

-0.05

0.00

0.05

0.10

0.15

0.20

0

1

2

3

4

5

6

7

8

9

10

B B-N P P-N C C-N W W-N

E ons

et[V

vs A

g/Ag

Cl]

Atom

ic %

O 1s % N 1s % Eonset

Page 90: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

77

5.5 Acknowledgements

The authors thank Dr. Cesar Nieto Delgado for assistance with the potentiometric

titration and Vince Bojan for assistance with XPS analysis. The authors acknowledge support

from the King Abdullah University of Science and Technology (KAUST) by Award KUS-I1-

003-13 and the National Science Foundation Graduate Research Fellowship Program (NSF-

GRFP).

5.6 References

1. Logan, B. and K. Rabaey, Conversion of Wastes into Bioelectricity and Chemicals by Using Microbial Electrochemical Technologies. Science, 2012. 337: p. 686-690.

2. Popat, S., et al., Importance of OH- Transport from Cathodes in Microbial Fuel Cells. ChemSusChem, 2012. 5: p. 1071-1079.

3. Zhong, R.-S., et al., Effect of carbon nanofiber surface functional groups on oxygen reduction in alkaline solution. Journal of Power Sources, 2013. 225: p. 192-199.

4. Zhang, F., et al., Power generation using an activated carbon and metal mesh cathode in a microbial fuel cell. Electrochemistry Communications, 2009. 11(11): p. 2177-2179.

5. Wei, B., et al., Development and evaluation of carbon and binder loading in low-cost activated carbon cathodes for air-cathode microbial fuel cells. RSC Advances, 2012. 2: p. 12751-12758.

6. Dong, H., H. Yu, and X. Wang, Catalysis kinetics and porous analysis of rolling activated carbon-PTFE air-cathode in microbial fuel cells. Environ Sci Technol, 2012. 46: p. 13009-13015.

7. Arrigo, R., et al., Tuning the acid/base properties of nanocarbons by functionalization via ammination Journal of the American Chemical Society, 2010. 132: p. 9616-9630.

8. Shafeeyan, M.S., et al., A review on surface modification of activated carbon for carbon dioxide adsorption. Journal of Analytical and Applied Pyrolysis, 2010. 89: p. 143-151.

9. Chen, W., F.S. Cannon, and J.R. Rangel-Mendez, Ammonia-tailoring of GAC to enhance perchlorate removal. I: Characterization of NH3 thermally tailored GACs. Carbon, 2005. 43: p. 573-580.

10. Mangun, C.L., et al., Surface chemistry, pore sizes and adsorption properties of activated carbon fibers and precursors treated with ammonia. Carbon, 2001. 39: p. 1809-1820.

11. Biniak, S., et al., The characterization of activated carbons with oxygen and nitrogen surface groups. Carbon, 1997. 35(12): p. 1799-1810.

12. Szymanski, G.S., et al., The effect of the gradual thermal decomposition of surface oxygen species on the chemical and catalytic properties of oxidized activated carbon. Carbon, 2002. 40: p. 2627-2639.

Page 91: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

78

13. Shen, W., Z. Li, and Y. Liu, Surface chemical functional groups modification of porous carbons. Recent Patents on Chemical Engineering, 2008. 1: p. 27-40.

14. Hulicova-Jurcakova, et al., Combined effect of nitrogen- and oxygen-containing functional groups of microporous activated carbon on its electrochemical performance in supercapacitors. Advanced Functional Materials, 2009. 19: p. 438-447.

15. Kruusenberg, I., et al., Non-platinum cathode catalysts fo alkaline membrane fuel cells. International Journal of Hydrogen Energy, 2012. 37: p. 4406-4412.

16. Nallathambi, V., et al., Development of High Performance Carbon Composite Catalyst for Oxygen Reduction Reaction in PEM Proton Exchange Membrane Fuel Cells. Journal of Power Sources, 2008. 183: p. 34-42.

17. Seredych, M. and T.J. Bandosz, Investigation of the enhancing effects of sulfur and/or oxygen functional groups of nanoporous carbons on adsorption of dibenzothiophenes. Carbon, 2011. 49: p. 1216-1224.

18. Bandosz, T.J., J. Jagiello, and C. Contescu, Characterization of the surfaces of activated carbons in terms of their acidity constant distributions. Carbon, 1993. 31(7): p. 1193-1202.

19. Jagiello, J., Stable numerical solution of the adsorption integral equation using splines. Langmuir, 1994. 10(8): p. 2778-2785.

20. Bandosz, T.J., et al., Efffect of surface chemisty on sorption of water and methanol on activated carbons. Langmuir, 1996. 12: p. 6480-6486.

21. Gojkovic, S.L., S. Gupta, and R.F. Savinell, Heat-treated iron(III) tetramethoxyphenyl porphyrin supported on high-area carbon as an electrocatalyst for oxygen reduction - I. Characterization of the electrocatalyst. Journal of the Electrochemical Society, 1998. 145: p. 3493-3499.

22. Kim, J.R., et al., Application of Co-naphthalocyanine (CoNPc) as alternative cathode catalyst and support structure for microbial fuel cells. Bioresource Technology, 2011. 102: p. 342-347.

23. Zhang, X., et al., Improved performance of single-chamber microbial fuel cells through control of membrane deformation. Biosensors and Bioelectronics, 2010. 25: p. 1825-1828.

24. Cheng, S., H. Liu, and B.E. Logan, Increased performance of single-chamber microbial fuel cells using an improved cathode structure. Electrochem Commun, 2006. 8: p. 489-494.

25. Liu, H., R. Ramnarayanan, and B.E. Logan, Production of electricity during wastewater treatment using a single chamber microbial fuel cell. Environmental Science & Technology, 2004. 38(7): p. 2281-2285.

26. Feng, Y., et al., Treatment of carbon fiber brush anodes for improving power generation in air–cathode microbial fuel cells. Journal of Power Sources, 2010. 195: p. 1841 - 1844.

27. Bretschger, O., et al., Current production and metal oxide reduction by Shewanella oneidensis MR-1 wild type and mutants. Appl Environ Microbiol, 2007. 73(21): p. 7003-7012.

28. Hong, Y., et al., Adaptation to high current using low external resistances eliminates power overshoot in microbial fuel cells. Biosens Bioelectron, 2011. 28(1): p. 71-76.

29. Logan, B.E., et al., Microbial fuel cells: methodology and technology. Environ Sci Technol, 2006. 40(17): p. 5181-5192.

Page 92: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

79

Chapter 6

Conclusions and Future Work

Through the research included in this dissertation we found that MFC polarization curves

measured by running for a full cycle at each resistance allows the anode performance to

stabilize and eliminate power overshoot. We also showed that polymer layers on the liquid

side of the cathode surface can hinder oxygen diffusion into the anode chamber, decrease

biofilm growth on the cathode and increase coulombic efficiency. Also, coating the cathode

with polymer layers increased the charge transfer resistance of the cathode, which decreased

power production, but the anion exchange polymers did not increase the resistance as much

as the other polymers studied. In the study of AC catalysts, we saw that both surface

chemistry and pore structure influenced the ORR performance and that increased surface area

or microporosity did not lead to improved performance. We also found that strong acid

oxygen functional groups (such as carboxyl groups) hindered the oxygen reduction activity of

the ACs, and that an increase in nitrogen groups and decrease in oxygen groups on the AC

surface (resulting from ammonia treatment) resulted in increased ORR activity.

Despite recent findings and improvements, further research is essential to the successful

development of MFC technology, such as:

• Improvement to the cathode structure to address diffusion and conductivity concerns

• Development of cathodes using activated carbon fiber as a way to increase

conductivity and stability of the cathode structure

• Development of ACs with higher concentrations of nitrogen groups either by pre–

oxidation or utilization of a precursor material that has higher nitrogen content (such

as silk fibroin)

Page 93: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

80

• Investigation and improvement in the stability/longevity of activated carbon catalysts

in the presence of organic contaminants

Page 94: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

81

Appendix A

Supplemental Information to Chapter 4

Figure A-1: Example of RDE LSV data for bituminous coal based sample (B1) collected at rotation rates from 100 – 2100 RPM.

-1

-0.8

-0.6

-0.4

-0.2

0

0.2

-1 -0.8 -0.6 -0.4 -0.2 0 0.2

J (m

A/c

m2 )

E (V vs Ag/AgCl)

B1 100rpm

B1 350rpm

B1 600rpm

B1 1100rpm

B1 1600rpm

B1 2100rpm

Page 95: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

82

Figure A-2: Example of K-L analysis for bituminous coal based sample (B1) where the slope of the line is used to calculate n and the y-intercept is the inverse ik.

y = -84.9x - 6.1R² = 1.00

y = -76.8x - 4.0R² = 1.00

y = -71.0x - 3.3R² = 1.00

y = -70.7x - 2.6R² = 1.00

y = -77.3x - 1.5R² = 1.00

y = -80.4x - 1.2R² = 1.00

y = -86.9x - 11.8R² = 1.00

-45

-40

-35

-30

-25

-20

-15

-10

-5

00 0.1 0.2 0.3 0.4

I-1(m

A-1

)

ω-1/2 (rad-1/2s1/2)

Koutecky-Levich Analysis

B1 -0.3VB1 -0.4VB1 -0.5VB1 -0.6VB1 -0.8VB1 -1VB1 -0.2V

Page 96: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

83

Figure A-3: Potentiometric titration curves showing protons bound (positive Q) or released (negative Q). The isotherm data was then further analyzed using SAIEUS software to quantify the type (pKa) and quantity of acidic functional groups.

-0.8

-0.6

-0.4

-0.2

0

0.2

0.4

0.6

2 3 4 5 6 7 8 9 10 11 12

Q (m

mol

/g)

pH

Proton Binding Isotherm

P2P3P1C1C3C2B1R1W1

Page 97: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

84

Figure A-4: Inverse correlation between quantity of strong acid functional groups on the AC catalyst powder and the power production in an MFC using the AC cathode (a) with and (b) without the inclusion of the bituminous coal based sample (B1)

P2P3P1

B1

W1

C2

C1

C3

R1

R² = 0.521

0200400600800

10001200140016001800

0 0.1 0.2 0.3 0.4

Pow

er D

ensit

y [m

W/m

2 ]

Strong Acid Groups (pKa<8) [mmol/g]

a

P2P3P1

W1

C2

C1

C3

R1

R² = 0.8169

0200400600800

10001200140016001800

0 0.1 0.2 0.3 0.4

Pow

er D

ensit

y [m

W/m

2 ]

Strong Acid Groups (pKa<8) [mmol/g]

b

Page 98: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

85

Figure A-5: Cumulative pore volume distribution of AC catalyst powders measured by argon adsorption and DFT analysis

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1 10 100 1000 10000

Cum

ulat

ive

Por

e Vo

lum

e (m

l/g)

Pore Width (Å)

P3

P2

P1

C1

C3

C2

B1

R1

W1

micro meso macro

Page 99: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

86

Appendix B

Supplemental Information to Chapter 5

Figure B-1. (A) Power production increases with the additional stainless steel current collector. (B) Cathodes (open symbols) with the additional current collector operate at a higher potential. Anodes (filled symbols) perform the same with both cathode configurations. (100 mM Phosphate buffer, Data of MFC cathodes without the extra current collector is from Chapter 4.)

0

500

1000

1500

2000

2500

0.0 0.5 1.0 1.5

Pow

er D

ensit

y [m

W/m

2 ]

P-ssPB-ssBC-ssCW-ssWPtC

-0.5

-0.4

-0.3

-0.2

-0.1

0.0

0.1

0.2

0.3

0.4

0.0 0.5 1.0 1.5

Elec

trod

e Po

tent

ial (

V vs

Ag/

AgCl

)

I (mA/cm2)

Pt/CBPCWB-ssP-ssW-ssC-ss

B

Page 100: CHARACTERIZATION AND PERFORMANCE OF ACTIVATED CARBON

CURRICULUM VITAE

VALERIE J. WATSON

182 Kathryn Drive, Bellefonte, PA 16823, 814-880-1810, [email protected]

EDUCATION Ph.D. Environmental Engineering, May 2013 The Pennsylvania State University, University Park, PA Advisor: Bruce E. Logan, Evan Pugh Professor and Kappe Professor of Environmental Engineering M.S. Environmental Engineering, May 2009 The Pennsylvania State University, University Park, PA Advisor: Bruce E. Logan, Kappe Professor of Environmental Engineering M.M.M. Quality and Manufacturing Management, May 1998 The Pennsylvania State University, University Park, PA B.S. Chemical Engineering, May 1997 The Pennsylvania State University, University Park, PA

EXPERIENCE Graduate Research Fellow & Assistant, Environmental Engineering, Penn State, 2006-2013 ISO Project Manager, Armstrong World Industries, Floor Products Division, Lancaster, PA, 2001 – 2002 Project Chemical Engineer, Armstrong World Industries, Floor Products Division, Lancaster, PA, 1999 2001 Quality Engineer, AMP Incorporated, Harrisburg, PA, 1998 – 1999 Quality Engineer Intern, Carpenter Technology Corporation, Reading, PA, Summer 1997

PUBLICATIONS Watson, V.J. and B.E. Logan. 2011. Analysis of polarization methods for elimination of power overshoot in microbial fuel cells. Electrochem. Commun. 13(1):54-56. Watson, V.J., T. Saito, M.A. Hickner, and B.E. Logan. 2011. Polymer coatings as separator layers for microbial fuel cell cathodes. J. Power Sources. 196(6):3015-3025. Saito, T., M.D. Merrill, V.J. Watson, B.E. Logan, and M. A. Hickner. 2010. Investigation of ionic polymer cathode binders for microbial fuel cells. Electrochim. Acta. 55(9):3398-3403. Watson, V.J., and B.E. Logan. 2010. Power production in MFCs inoculated with Shewanella oneidensis MR-1 or mixed cultures. Biotechnol. Bioengin. 105(3):489-498. Logan, B.E., Cheng, S., Watson, V., and Estadt, G. (2007) “Graphite fiber brush anodes for increased power production in air-cathode microbial fuel cells.” Environmental Science & Technology, 41(9): p. 3341-3346

HONORS AND AWARDS Presentation Award Winner, North American Meeting of the International Society for Microbial Electrochemistry and Technology, Cornell University (2012) National Science Foundation Graduate Fellowship (2008-2011) Hydrogen Day at Penn State, Poster presentation honorable mention (2006) Cecil M Pepperman Memorial Graduate Fellowship, Penn State University (2006) General Electric’s First-Year Faculty for the Future Fellowship, Penn State University (2006) Manager’s award for team excellence, Armstrong World Industries (2002) Manager’s award for product development project (2000) QMM Scholarship for Academic Excellence, Penn State University (1998) Omega Chi Epsilon, Chemical Engineering Honor Society, Penn State University (1995-1997)