by william jason harrington - university of...

93
1 INVESTIGATION OF DIRECT METHANOL FUEL CELL VOLTAGE RESPONSE FOR METHANOL CONCENTRATION SENSING BY WILLIAM JASON HARRINGTON A THESIS PRESENTED TO THE GRADUATE SCHOOL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF MASTER OF SCIENCE UNIVERSITY OF FLORIDA 2012

Upload: others

Post on 17-Apr-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

1

INVESTIGATION OF DIRECT METHANOL FUEL CELL VOLTAGE RESPONSE FOR

METHANOL CONCENTRATION SENSING

BY

WILLIAM JASON HARRINGTON

A THESIS PRESENTED TO THE GRADUATE SCHOOL

OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT

OF THE REQUIREMENTS FOR THE DEGREE OF

MASTER OF SCIENCE

UNIVERSITY OF FLORIDA

2012

Page 2: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

2

© 2012 William Jason Harrington

Page 3: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

3

To my wife Ashley, my parents Bill & Xinia and brother Jesse, thanks for your enduring love

and support

Page 4: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

4

ACKNOWLEDGMENTS

I thank the members of my committee Dr. William E. Lear, Dr. James H. Fletcher and Dr.

David W. Mikolaitis for their support and guidance on this thesis. I must also thank Dr. Joseph

L. Campbell, Dr. Philip Cox and Dr. Oscar D. Crisalle for their wisdom and advice. Finally, I

would like to thank all of my fellow colleagues in the University of North Florida Fuel Cell

Laboratory and University of Florida Energy Park for their encouragement and comradery.

Page 5: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

5

TABLE OF CONTENTS

page

ACKNOWLEDGMENTS .............................................................................................................. 4

LIST OF TABLES .......................................................................................................................... 7

LIST OF FIGURES ........................................................................................................................ 8

ABSTRACT .................................................................................................................................. 11

CHAPTER

1 INTRODUCTION ................................................................................................................. 13

2 LITERATURE REVIEW ...................................................................................................... 17

Rechargeable Battery Technology Status .............................................................................. 17 Lithium-ion Battery Advantages .................................................................................... 17

Lithium-ion Battery Disadvantages ................................................................................ 17 Direct Methanol Fuel Cells .................................................................................................... 18

Effects of Methanol Concentration ................................................................................ 19 Methanol Sensing Technologies ............................................................................................ 21

Physical Property Type Methanol Sensing ..................................................................... 21

Capacitance-based sensors ...................................................................................... 21

Speed of sound-based sensors ................................................................................. 22 Refractive index-based sensors ............................................................................... 22 Infrared spectrum-based sensors ............................................................................. 23

Heat capacity-based sensors .................................................................................... 23 Viscosity-based sensors ........................................................................................... 24

Dynamic viscosity-based sensors ............................................................................ 24 Electrochemical Type Methanol Sensing ....................................................................... 25

3 EXPERIMENTATION AND DATA ANALYSIS ............................................................... 29

Test Station Description ........................................................................................................ 29 Fuel Cell Hardware ................................................................................................................ 31 Experimentation ..................................................................................................................... 32

Active Load Method .............................................................................................................. 33 Experimentation ............................................................................................................. 35 Results ............................................................................................................................ 36

Passive Load Method ............................................................................................................. 39 Experimentation ............................................................................................................. 39 Results ............................................................................................................................ 43

Page 6: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

6

4 VERIFICATION MODEL .................................................................................................... 52

Methanol Concentration Distribution .................................................................................... 52 Steady State Concentration Distribution ........................................................................ 52 Transient Concentration Distribution ............................................................................. 54

Transient Oxygen Concentration Distribution ....................................................................... 56 Modeling Summary ............................................................................................................... 57

5 SYSTEM INTEGRATION ................................................................................................... 58

Methanol Concentration Tracking ......................................................................................... 58 Methanol Consumption Model ....................................................................................... 58

Faradaic oxidation ................................................................................................... 59

Methanol crossover ................................................................................................. 60

Methanol Injection Model .............................................................................................. 61 Methanol Concentration Determination ......................................................................... 63

Brassboard Operation ............................................................................................................ 63

6 CONCLUSIONS ................................................................................................................... 65

APPENDIX

A DIFFUSION MODEL DEVELOPMENT ............................................................................. 68

Initial Conditions ................................................................................................................... 70 Anode Diffusion Layer (0 ≤ z ≤ zAD) ..................................................................................... 71

Anode Catalyst Layer (zAD ≤ z ≤ zAC) ................................................................................... 72 Membrane Layer (zAC ≤ z ≤ zM) ............................................................................................ 73

Cathode Catalyst Layer (zM ≤ z ≤ zCC) .................................................................................. 75 Homogenous Equations ......................................................................................................... 77 Steady State Equations .......................................................................................................... 80

B DIFFUSION MODEL MATLAB CODE ............................................................................. 84

LIST OF REFERENCES .............................................................................................................. 90

BIOGRAPHICAL SKETCH ........................................................................................................ 93

Page 7: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

7

LIST OF TABLES

Table page

3-1 OCV rise time at 0.60 M and 0.80 M. .............................................................................. 45

4-1 Summary of methanol crossover results from model for various feed methanol

concentrations with a current density of 150 mA/cm². ..................................................... 54

A-1 Matrix form of boundary equations for homogeneous equations. .................................... 80

A-2 Modified matrix of homogenous set of boundary equations. ........................................... 80

A-3 System of equations for homogeneous set of boundary conditions. ................................. 80

A-4 Boundary conditions for steady state, non-homogeneous boundary conditions............... 83

Page 8: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

8

LIST OF FIGURES

Figure page

2-1 The effect of methanol concentration on a typical DMFC at 75°C. ................................. 19

2-2 The effect of methanol concentration on a typical DMFC at low current densities. ........ 20

2-3 Transient behavior of DMFC with respect to methanol concentration ............................. 27

3-1 Screenshot of LabVIEW interface on University of North Florida Test Stations ............ 29

3-2 University of North Florida standard fuel cell test station. ............................................... 30

3-3 UNF Fuel Cell Test Station instantaneous methanol control attachment. ........................ 30

3-4 P&ID for University of North Florida standard test station and methanol control

attachment. ........................................................................................................................ 31

3-5 Compressed eight cell fuel cell stack. ............................................................................... 32

3-6 DMFC methanol concentration sensitivity with static loading......................................... 33

3-7 DMFC current density response with load change from 0.40 V to 0.35 V. ..................... 34

3-8 Transient current for a DMFC with load oscillations from 0.35 V to 0.40 V at 0.20 M

and 50°C. .......................................................................................................................... 35

3-9 Transient current for a DMFC with load oscillations from 0.35 V to 0.40 V at 0.60 M

and 50°C. .......................................................................................................................... 36

3-10 Transient current for a DMFC with load oscillations from 0.35 V to 0.40 V at 1.60 M

and 50°C. .......................................................................................................................... 36

3-11 Unfiltered fuel cell transient response when electrically loaded from 0.40 V to 0.35

V at 50°C. ......................................................................................................................... 37

3-12 Unfiltered fuel cell transient response when electrically loaded from 0.35 V to 0.40

V at 50°C. ......................................................................................................................... 37

3-13 Filtered and normalized fuel cell transient response when electrically loaded from

0.40 V to 0.35 V at 50°C. ................................................................................................. 38

3-14 Filtered and normalized fuel cell transient response when electrically loaded from

0.35 V to 0.40 V at 50°C. ................................................................................................. 38

3-15 Current Density Transient Response Power Curve Fit Coefficients for 0.35 V and

0.40 V at 50°C. ................................................................................................................. 39

Page 9: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

9

3-16 Representative fuel cell voltage response from loaded to unloaded operating point. ...... 40

3-17 Fuel cell air starve cycle with 0.60, 0.80 and 1.60 molar concentration at 50 °C. ........... 41

3-18 Typical load cycle for DMFC OCV decay slope testing. ................................................. 42

3-19 OCV decay slope at 120 mA/cm² at various methanol concentrations with error bars

indicating first standard deviation. .................................................................................... 44

3-20 OCV decay slope at various current densities with 50°C stack temperature. ................... 45

3-21 OCV Decay slope with constant current density and variable concentration and

temperature. ...................................................................................................................... 47

3-22 Max OCV rise time at constant current density with varying temperature and

methanol concentration. .................................................................................................... 48

3-23 OCV Rise Slope held at constant current density (40 mA/cm²) with variable

concentration and stack temperature with error bars indicating single standard

deviation. ........................................................................................................................... 49

3-24 OCV Rise Slope held at constant current density (120 mA/cm²) with variable

concentration and stack temperature with error bars indicating single standard

deviation. ........................................................................................................................... 50

4-1 Results from model for MEA methanol concentration distribution for 0.80 M feed

concentration at a current density of 150 mA/cm². ........................................................... 53

4-2 Results from model for MEA methanol concentration distribution for various feed

concentrations at a current density of 150 mA/cm². ......................................................... 53

4-3 Results from model for MEA transient concentration distribution from a current

density of 120 mA/cm² to 0 mA/cm² at a feed concentration of 0.80 M. ......................... 54

4-4 Results from model for MEA transient concentration distribution from a current

density of 120 mA/cm² to 0 mA/cm² at a feed concentration of 1.60 M. ......................... 55

4-5 Model results for transient methanol concentration response at cathode catalyst layer

from a current density of 120 mA/cm² to 0 mA/cm². ....................................................... 55

4-6 Model results for the mean transient oxygen content in the cathode. ............................... 57

5-1 UNF 20 W DP4 brassboard fuel cell system. ................................................................... 58

5-2 Simplified methanol consumption and injection model for DP4...................................... 59

5-3 Representative methanol crossover current density for DP4 stack at various stack

temperatures, methanol concentrations and current densities. .......................................... 60

Page 10: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

10

5-4 Performance Curve for Single µBase Pump ..................................................................... 62

5-5 Comparison of µBase Pump Performance versus various inlet pressures. ....................... 62

5-6 UNF DP4 brassboard operation using sensor-less methanol sensing techniques. ............ 64

6-1 Ideal methanol concentration distribution at various load conditions. ............................. 68

Page 11: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

11

Abstract of Thesis Presented to the Graduate School

of the University of Florida in Partial Fulfillment of the

Requirements for the Degree of Master of Science

INVESTIGATION OF DIRECT METHANOL FUEL CELL

VOLTAGE RESPONSE FOR METHANOL CONCENTRATION SENSING

By

William Jason Harrington

August 2012

Chair: William E. Lear, Jr.

Major: Mechanical Engineering

A Direct Methanol Fuel Cell (DMFC) was tested under various transient load conditions in

order to determine the sensitivity of response to methanol concentration. In addition to varying

load profiles, the DMFC was tested at several temperature and methanol concentration operating

conditions. The results demonstrated a strong correlation of open circuit voltage transient

response to methanol concentration with high repeatability and resolution in the methanol

concentration range of 0.60 - 1.60 M.

The findings and phenomena that were observed in the experiments were further studied

using a simple, 1-D, transient methanol diffusion model. The model represents the transient

methanol crossover within the membrane electrode assembly of the DMFC. The transient

methanol crossover values that were calculated using the model were used to approximate the

transient oxygen consumption for an open cathode system. The results generated by the

computer model exhibited similar timescales compared to what was observed during the DMFC

testing. This supports the theory of a cathode dominant response due to the change in methanol

crossover with various methanol feed concentrations.

Page 12: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

12

Finally, the measurements that were gathered during DMFC testing were applied in a

brassboard system. A table was generated allowing a DMFC open circuit transient voltage

response to be correlated to a methanol concentration. Due to the operation profile that a DMFC

must undergo, the open circuit transient voltage response could only be captured during

rest/rejuvenation cycles which typically occur every 10-20 minutes. Therefore, a secondary

model had to be created in order to track the methanol concentration between rest/rejuvenation

cycles by predicting the consumption (Faradaic, crossover) and addition (methanol injection) of

methanol in the system. Utilizing the transient open circuit voltage response with the methanol

concentration estimator allowed for over 20 hours of continuous operation in a brassboard

without the use of a secondary methanol sensor.

Page 13: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

13

CHAPTER 1

INTRODUCTION

The advances in multimedia (internet, social networking, multi-megapixel photos, high

definition video, high fidelity music, television programming, etc.) and wireless communications

(Wi-Fi, Bluetooth®, WWAN, 4G, LTE, 3G, etc.) have transformed the way people communicate

today. While the computing power in portable electronics such as laptops and cell phones

continues to double every two years [1], the energy density of the most common power source

found in these devices (lithium ion batteries) only doubles every thirteen years [2].Consumers

demand access to content and services at all times through devices with larger screens and faster

processors while achieving a lighter weight and a thinner profiles. The use of such devices with

existing battery technology has negatively impacted their run time. In addition to the deficiency

of the energy density found in lithium ion batteries, power density has also become a

technological limitation. As lithium ion batteries are pushed to higher power densities, the

combination of added heat generation and the limitations in manufacturing have resulted in an

increase of lithium ion battery related fires [3]. One of the most promising solutions that is being

considered as a potential replacement for battery technology is the direct methanol fuel cell

(DMFC) which offers advantages in both energy and power density.

Fuel cells are electrochemical devices that convert chemical energy into electrical energy.

The fuel cell is similar to a battery, as both operate through an electrochemical reaction; however

fuel cells have the advantage of storing the fuel and oxidant externally. This enables the fuel cell

to operate indefinitely provided that sufficient reactants are present. Because the energy

conversion process that takes place in a fuel cell is not based on the process that limits most

typical heat engines (Carnot cycle), the fuel cell is able to achieve high efficiencies at low

temperatures.

Page 14: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

14

The majority of PEM fuel cells accept hydrogen at the anode and oxygen at the cathode to

electrochemically produce power. A DMFC ultimately uses hydrogen at the anode, however a

methanol-water solution is electrochemically converted to hydrogen without the use of

intermediate steps or equipment. The use of methanol as a fuel has many advantages over

hydrogen including ease of storage, high availability and low cost. The DMFC has the distinct

advantage of higher energy density over a typical hydrogen-oxygen based fuel cell with low

power applications (< 100 W). The high energy density of methanol fuel and the ease of storage

and handling enable the DMFC to perform for longer periods of time given the same system

volume. This characteristic makes DMFC technology a prime candidate for portable electronic

applications.

1-1

1-2

1-3

The anode (1-1) and cathode (1-2) half reactions can be combined to form an overall

reaction (1-3) for the DMFC. Although there are many intermediate reactions that take place

before the overall reaction is completed, the half reactions are the most simplified way to

describe the processes that takes place within the fuel cell. At the anode, methanol and water are

electrochemically converted to hydrogen protons, electrons and carbon dioxide. In the cathode

reaction, oxygen and the hydrogen protons generated at the anode react to form water. One of the

most important operating parameters for a DMFC is the methanol concentration at the anode.

Operation of a DMFC using very low concentrations (less than 0.4 Molarity) of methanol can

result in reduced limiting current density, peak power and damage due to fuel starvation [4]. Due

Page 15: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

15

to the existing limitations of DMFC membrane technology, DMFCs should not operate using

equi-molar ratios of pure methanol and water as implied by Equation 1-1. When high

concentrations of methanol (greater than 1.0 Molarity) are exposed to the anode, excessive

permeation of methanol across the membrane occurs; this phenomenon has been termed

“methanol crossover”. The adverse effects of methanol crossover include depolarization at the

cathode and poor fuel utilization [5]. The methanol diffuses from the anode to the cathode across

the membrane, eventually reaching the platinum catalyst of the cathode, where it reacts with

oxygen from the air, creating waste heat. In order to minimize methanol crossover and maintain

reasonable DMFC output power levels, a DMFC will typically operate with an aqueous solution

of methanol (0.6-1.0 Molarity).

Due to the simplicity of operation at higher methanol concentrations, manufacturers are

working to mitigate issues encountered with methanol crossover by improving membrane

technology [6]. The essential task of producing a membrane that has a higher ion exchange

capacity while allowing less methanol to diffuse is difficult. PolyFuel Inc. developed a

hydrocarbon-based membrane to compete with DuPont’s widely used fluorocarbon based Nafion

117. PolyFuel claims that its family of membranes offers a 33 to 50 percent improvement in the

level of methanol crossover, water flux, and power density when compared to Nafion [7].

Although advancements in membrane technology are critical to improving DMFC power and

energy density, the methanol concentration will continue to be an essential factor for DMFC

performance.

In order to achieve optimum performance, the DMFC must operate in a tight methanol

concentration band. A number of technologies exist to measure the concentration of methanol,

however, few are able to meet the requirements (size, costs, weight, reliability, accuracy, etc.)

Page 16: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

16

and electrochemical sensing methods. Physical-type sensors correlate methanol concentration to

a physical property such as density or heat capacity. Some of these sensors are fairly robust;

however they usually do not package well for miniature applications and often require auxiliary

devices (pumps, heaters, complex sensing systems) which increase parasitic power loads on the

DMFC. Most electrochemical-type sensors work using the same principles found in DMFCs

where a signal can be interpreted based on electrochemical response. The most accessible

electrochemical sensor to integrate into a DMFC system is the fuel cell stack itself. With the

added benefit of reduced cost, weight and space, the stack is an ideal replacement for a methanol

sensor.

It is the goal of this thesis to correlate the transient voltage response of a DMFC to

methanol concentration in order to eliminate the requirement for a discrete methanol

concentration sensor in a DMFC system. Integration of a sensor-less (operation without a

methanol sensor) methanol sensing technique will significantly reduce system cost, weight and

complexity accelerating the movement of DMFC technology.

Page 17: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

17

CHAPTER 2

LITERATURE REVIEW

Rechargeable Battery Technology Status

Battery technology is the most prevalent form of portable power. Due to its convenience

and low operating cost, rechargeable batteries account for 85% of fiscal battery sales [9].

According to McAllister and Farrell, the annual electrical demand for the average household in

the state of California is 125 kWhr for rechargeable devices, which accounts for 2.3% of the total

electrical load [8]. Typically only 15% of the energy that is used to recharge batteries is stored,

while the remainder is wasted as heat [8]. The sales published by the Battery Association of

Japan show that lithium-ion battery technology is the most prevalent rechargeable battery

chemistry used today, accounting for 47% of the fiscal sales of all rechargeable batteries [9].

Lithium-ion Battery Advantages

Lithium-ion batteries have many advantages over other battery technologies. Lithium is the

lightest metal in the periodic table, therefore lithium-ion batteries are lighter than most other

battery chemistries. In addition to lithium’s light weight, the metal is also highly reactive with an

average open circuit potential of 3.7 Volts, resulting in one of the highest energy density battery

chemistries [10]. As published by Powers, lithium-ion battery technology also exceeds other

battery chemistries in maximum charge cycles and off-state discharge rates [10].

Lithium-ion Battery Disadvantages

However, lithium-ion battery technology has its drawbacks. Scrosati and Vetter both

observed accelerated degradation in the off state when storing lithium-ion batteries at elevated

temperatures [11, 12]. Shim published data for lithium ion battery testing at 60°C, where the

degradation was 15 times more than at 25°C (10% annually at 25°C) [13]. Researchers have

determined that the root cause of high temperature degradation for lithium ion batteries occurs

Page 18: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

18

within the morphology and composition of the solid electrolyte interface, which changes at

elevated temperatures [12].

In addition to degradation, elevated temperatures can adversely affect the stability and

safety of lithium ion batteries. To date, the Federal Aviation Administration has documented

over 28 lithium-ion related fires/explosions on commercial airline flights [14]. Due to the high

level of energy that can be stored in lithium-ion batteries, injury and/or death can occur due to

failure.

The major drawback for lithium-ion batteries is the rate at which they technologically

advance. Historically, lithium-ion technology advances at a rate of approximately 5% annually

while the microprocessors that most of these batteries power advance in processing speed at a

rate of 40% annually [1, 2]. This poses a major problem for the advancement of portable

electronics in general. A recent development in the energy density of batteries in portable

electronics has primarily been achieved by replacing off the shelf battery cells with custom fit,

non-user removable battery packs. Apple computers was one of the first major manufacturers to

incorporate this strategy resulting in a 30% increase in battery capacity [15]. Research institutes

such as Stanford University have reported advances (nanowire technology) in lithium-ion

technology, however the timeline for implementation of these breakthroughs is 5-10 years away

[16]. Without a power source that can meet the demands of tomorrow’s portable electronics, the

consumer will see a slowdown in the progression of portable technology that is available today.

Direct Methanol Fuel Cells

Direct methanol fuel cells have the potential to replace lithium-ion battery technology as a

power source in portable electronics. Direct methanol fuel cells (DMFCs) work on similar

principles to batteries. Both devices electrochemically convert fuel and oxidant into electricity;

however the fuel cell has the inherent advantage of storing its reactants externally. Instead of

Page 19: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

19

waiting for a battery to recharge, the fuel cell simply requires a new fuel cartridge of methanol to

continue producing power. The possibility of carrying multiple batteries is an option to extend

portable electronic runtimes; however the energy density of the liquid methanol that DMFCs use

exceeds all common battery technologies. It is this property that enables DMFCs to excel over

battery technology during extended durations (>10 hours). Direct methanol fuel cells are an

enabling technology which can give portable electronics manufacturers the flexibility to expand

the capabilities of portable devices without the risk of reducing portable operation duration.

Effects of Methanol Concentration

One of the major technological barriers that DMFCs must overcome are its sensitivity to

methanol concentration. The Nernst equation for the concentration polarization at the anode

suggests that the concentration losses can be minimized by increasing the limiting current

density. The limiting current density for the anode is directly related to the amount of methanol

(methanol concentration) that is present [17]. However, as shown in Figure 2-1, data presented

by Song suggests that the performance of a direct methanol fuel cell does not always increase

with increasing methanol concentration.

Figure 2-1. The effect of methanol concentration on a typical DMFC at 75°C [18].

Page 20: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

20

The effect of methanol concentration on DMFC performance indicates increased

performance with increasing methanol concentration until reaching a feed solution concentration

of 2.0 M. At high current densities (greater than 100 mA/cm²), the effect of methanol crossover

is less severe. Only until reaching 2.0 M do the benefits of reducing anode concentration losses

outweigh the voltage losses due to methanol crossover [18]. Primarily driven by concentration

and pressure gradients, as the concentration of methanol at the anode increases, more methanol

crosses the membrane depolarizing the cathode potential. Du, Zhao, and Yang report that the

resulting decline in cathode electrode performance is related primarily to the “poisoning” of the

cathode catalyst from methanol oxidation intermediates such as CO [19].

The effect of the feed methanol concentration on DMFC performance is most evident at

low current densities. As shown in Figure 2-2, Song presented data where a strong relationship

of fuel cell performance at low current densities with respect to feed methanol concentration was

established. With a low feed methanol concentration (0.25 M), the methanol crossover is reduced

allowing for the fuel cell to operate with a 12.5% increase in performance relative to a high

methanol concentration (4.0 M).

Figure 2-2. The effect of methanol concentration on a DMFC at low current densities [18].

Page 21: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

21

Methanol Sensing Technologies

There is strong evidence in the published literature that shows direct methanol fuel cell

performance is heavily dependent on the feed methanol concentration at the anode [4], [18], [23],

[20]. Therefore it is critical that the anode methanol concentration be kept at a level that is

optimized for DMFC power output and efficiency. In order to monitor the methanol

concentration in the DMFC, a methanol sensor can be used in the anode loop. A number of

different methanol sensing technologies exist each with its own advantages and disadvantages.

Most of the methanol sensors that are available today can be separated into physical property

type sensors or electrochemical type sensors.

Physical Property Type Methanol Sensing

Capacitance-based sensors

Doerner proposed a capacitance based methanol sensor utilizing impedance spectrum

analyzer electronics [21]. The sensor uses two planar sensing electrodes to measure the dielectric

constant (capacitance) of a test solution in order to determine the methanol concentration. The

sensor exhibited high signal to noise levels at low frequencies (>300 kHz) and a strong

relationship with respect to temperature. Capacitance-type sensors have been used in the past to

determine the concentration of methanol in gasoline-methanol fuel mixtures with reasonable

results due to the high disparity of dielectric constants for gasoline (2.0 [22]) and methanol (32.6

[22]). The distinction between the dielectric constants for water (80.4 [22]) and methanol (32.6

[22]) is much less. The resolution for measuring the dielectric constant of methanol solutions less

than 5.0 Molar is small. In addition to errors that may arise from measurement resolution, the

corrosion of electrodes, CO2 bubbles generated by the DMFC in the anode stream, or metallic

ions can severely impact capacitance measurements [23].

Page 22: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

22

Speed of sound-based sensors

A speed of sound-based methanol sensor was proposed in patent 6,748,793 by Rabinovich

and Tulimieri utilizing an ultrasonic sensor. According to the patent literature, the sound

propagation time is measured for a given distance in order to determine the speed of sound of the

mixture. Primarily based on the density and bulk modulus of elasticity of the test medium, the

speed of sound for methanol and water is 1580 m/s and 1150 m/s, respectively [23]. Like the

capacitance-type sensors, the disparity in sound velocities for methanol and water is not great

enough to provide high resolution measurements, particularly at elevated temperatures.

Temperature has a strong effect on the speed of sound of methanol-water solutions, therefore

Rabinovich and Tulimieri proposed a second measurement chamber for a calibration sample

(deionized water). This would allow the sensor to offset the temperature effects by compensating

with the output of the calibration sample. In addition to the low resolution provided by speed of

sound-type sensors, this type of sensor is not easily miniaturized and can also suffer from

measurement error due to anode CO2 generation.

Refractive index-based sensors

Longtin and Fan developed a refractive index-based concentration sensor that uses a one

mW 632.8 nm laser in conjunction with a semiconductor position sensor [24]. They were able to

achieve a highly accurate, small and inexpensive concentration sensor. However some

measurement error was observed attributed to vibration, air disturbances and laser fluctuations.

In addition, although the refractive index concept offers the most simple of designs, the change

in refractive index at low concentrations provides the least resolution among the discussed

concentration sensing methods. The refractive index for water is 1.333 [25] and the refractive

index for pure methanol is 1.329 [25].

Page 23: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

23

Infrared spectrum-based sensors

DuPont currently holds a patent for an infrared spectrum-based methanol sensor [26]. The

sensor works using similar principals found in the refractive index sensor. An IR source

transmits a non-visible light through a test medium where a photodetector analyzes the

transmitted light and a microprocessor converts the signal to a methanol concentration. A strong

relationship between absorption and methanol concentration can be determined from IR

wavelengths in the range of 9.8 to 9.9 µm. The infrared spectrum provides excellent resolution

for low methanol concentrations (less than 1.5 M), however measurement accuracy is still

effected by CO2 bubbles [23].

Heat capacity-based sensors

Siargo has developed a prototype heat capacity type methanol sensor utilizing their flow

measurement technology. The sensor requires a heat source, a constant flow rate and two

temperature sensors. The methanol solution temperature is measured using one of the

temperature sensors as heat is applied to the solution. The temperature of the solution

downstream of the heater is measured to determine the corresponding heat capacity. The isobaric

specific heat capacity for methanol and water is 78.81 J/mol·K [27] and 75.40 J/mol·K [27],

respectively. For aqueous methanol solutions, the heat capacity vs. methanol concentration curve

is non-linear. The heat capacity sharply increases until the molar concentration reaches 7.0

molarity, where it steadily starts to decrease with increasing methanol concentration. For a

solution flow rate of 100 mL/min, the difference in temperature rise for a solution changing from

0.5 Molarity to 1.0 Molarity is 2.0 °C [23]. Although this temperature difference is easily

measured, the accuracy of heat capacitance-type devices can be heavily influenced by

movement. At constant flow velocities, a convection coefficient can be determined for the

sensor. If the sensing device were to be abruptly shaken, the resulting convection coefficient

Page 24: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

24

would change resulting in a varied temperature rise across the heater. In addition to the

inaccuracies created by abrupt movements, the parasitic power from the heater and additional

pumping requirements must also be taken into consideration when dealing with portable power

applications.

Viscosity-based sensors

Siemens currently holds a US patent for a methanol sensor that measures the viscosity of a

methanol solution in order to determine its concentration [28]. The methanol solution is pumped

through a constriction and the pressure drop across the constriction is measured in order to

determine its viscosity using the Hagen-Poiseuille equation. At 20ºC, water is characterized with

a viscosity that is 1.5 times more than for methanol [29]. Interestingly, the viscosity of an

aqueous methanol solution increases with concentration until reaching a concentration of 12.0

molarity [23]. The large change in viscosity for low methanol concentrations provides excellent

resolution for correlating methanol concentration. Measuring the pressure drop across a flow

restriction provides one of the simplest methods for determining fluid viscosity.

Determination of methanol concentration using viscosity has its drawbacks. The CO2

bubbles that are introduced into the solution line by the fuel cell can create large errors in

viscosity determination. In addition, viscosity is a strong function of temperature. The viscosity

of water is nearly half at 50°C compared to its value at 20°C [29]. Therefore an accurate means

of temperature measurement would also be required.

Dynamic viscosity-based sensors

Integrated Sensing Systems (ISSYS) produces a methanol sensor that utilizes a micro-

machined resonating tube to measure kinematic viscosity [30]. Kinematic viscosity is composed

of density and the dynamic viscosity. As the density of the test solution changes, the effective

mass of the resonating tube shifts, therefore affecting the resonant frequency. Additionally, the

Page 25: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

25

damping effect on the resonator can be used to determine the dynamic viscosity. Kinematic

viscosity sensors work well with low flow applications such as fuel cells. However, these sensors

do not miniaturize well due to the requirement of a bulky counterbalance. In addition, previous

experience has shown that density-type sensors severely suffer from impact shock and

measurement drift over time.

Electrochemical Type Methanol Sensing

An electrochemical-based methanol sensor was developed at the Jet Propulsion Laboratory

based on the electro-oxidation of methanol to carbon dioxide on a platinum-ruthenium catalyst

[31]. The proposed methanol sensor operates on the same principles found in DMFCs. The anode

potential is set at a constant voltage and the oxidation current is measured. At lower methanol

concentrations, the oxidation current is limited by the transport of methanol to the electrode

surface. Electrochemical-based methanol sensors offer strong measurement resolution and are

less sensitive to CO2 compared to other methanol sensing technologies. In addition,

electrochemical-based methanol sensors can easily be constructed and miniaturized due to their

simplicity and similarities to fuel cells. However, electrochemical-based sensors suffer from

nearly all issues encountered by DMFCs including contamination, degradation, catalyst

deterioration and slow response time [23].

Other electrochemical-based methanol sensors have been developed including one from

the Institute of Nuclear Energy Research in Taoyuan County, Taiwan. The sensing method

utilizes the fuel cell stack by measuring an operating characteristic such as voltage, current or

power and applies a control strategy for the methanol injection pump accordingly. Based on the

performance of the stack, a certain amount of methanol is required in order for the methanol

concentration in the system to remain constant. If a slightly higher amount of methanol is

precisely metered into the system than what the system consumes, the methanol concentration

Page 26: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

26

will increase. Response of the fuel cell stack can be analyzed in order to determine whether the

system methanol concentration increased closer or further away from the optimal operating

point. One inherent advantage to this control strategy is that the methanol concentration can be

optimized for a stack that has suffered from performance degradation [32]. However, operating

close to the optimal methanol concentration poses a risk for fuel starvation due to the proximity

of the optimal methanol concentration operating point to the fuel starvation point.

One of the largest manufacturers of commercial DMFC systems is Smart Fuel Cells

Energy Inc. They have developed many systems utilizing sensorless technology. Although the

exact method for methanol concentration determination is undetermined, it is believed that a

lookup table is used in order to determine the methanol concentration within the system based on

key system parameters (fuel cell voltage, current, temperature, etc.). One of the major drawbacks

to using a lookup table in a DMFC is the measurement error that is introduced when the fuel cell

stack experiences a non-standard degradation mechanism. In addition, repeated start-up and shut-

downs can pose a major problem for methanol concentration determination due to the lack of

operation time resulting in minimal feedback of methanol concentration. This can lead to damage

of the stack due to excessive methanol concentration or fuel starvation [33].

Differences in electrochemical response with respect to methanol concentration can also be

seen in the transient behavior of direct methanol fuel cells. Essentially, the transient behavior of

the fuel cell is characterized by the transport properties of methanol through the anode diffusion

layer and membrane. The relationship of methanol crossover relative to feed methanol

concentration is nearly linear [34]. When the fuel cell is operated under constant load, a certain

quantity of methanol diffuses through the diffusion layer to the anode catalyst. The concentration

Page 27: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

27

gradient across the anode is different at varying load levels. When the load is abruptly changed,

it takes time for the concentration gradient to reach equilibrium.

Figure 2-3. Transient behavior of DMFC with respect to methanol concentration. [35]

As shown in Figure 2-3, the load on the DMFC oscillates from a loaded condition to open

circuit. When the load on the fuel cell is removed, less methanol is required for the reaction to

take place. At this very moment, when the load is changed, the local concentration at the cathode

catalyst is relatively low. The decreased load also allows for an increase in the amount of

methanol that is available at the anode catalyst. The combination of low methanol concentration

at the cathode (decreased methanol crossover) and high methanol concentration at the anode

(decreased anode concentration overpotential) is ideal. As a result, the transient voltage

performance when switching from a loaded to non-loaded condition sharply increases until the

diffusion of methanol is able to equilibrate. As shown in Figure 2-3, the dynamic behavior for

each methanol concentration varies. Analysis of the effect of methanol concentration on the

Page 28: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

28

transient electrochemical performance will be used in order to determine the feed solution

methanol concentration in a direct methanol fuel cell.

Page 29: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

29

CHAPTER 3

EXPERIMENTATION AND DATA ANALYSIS

Test Station Description

In order to analyze the transient behavior of a direct methanol fuel cell relative to methanol

concentration, a test station had to be developed that would provide accurate, repeatable data for

analysis. The University of North Florida acquired a number of disassembled, incomplete, Fuel

Cell Technologies, Inc. test stations. These test stations were refurbished utilizing the existing

electric load banks, temperature controllers and enclosures. All other components (data

acquisition, cathode mass flow controller, signal conditioning, anode solution heater and pump,

computer, etc.) were procured and selected based on the engineering requirements.

Figure 3-1. Screenshot of LabVIEW interface on University of North Florida Test Stations

As shown in the screenshot in Figure 3-1, a new LabVIEW interface was written in order

to control all of the applicable test station components. As shown in Figure 3-2, the University of

North Florida standard test station is fully equipped including:

Fuel cell load control

Fuel cell voltage and current measurements

Anode solution flow control

Anode solution temperature control and measurement

Page 30: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

30

Fuel cell temperature control and measurement

Fuel cell cathode flow rate control and measurement

Customized LabVIEW interface

Script enabled test operation

Figure 3-2. University of North Florida standard fuel cell test station.

In addition to the standard fuel cell test station, an attachment was developed in order to

precisely control the methanol feed concentration delivered to the fuel cell. As shown in Figure

3-3, a PDS-100 dual head precision piston pump was mounted to an aluminum substructure with

a USB enabled data acquisition controller.

Figure 3-3. UNF Fuel Cell Test Station instantaneous methanol control attachment.

Page 31: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

31

The use of the methanol control attachment with the standard test station (P&ID shown in

Figure 3-4) enabled repeatable, accurate data collection with the flexibility of instantaneous

methanol concentration control.

PDS-100 Pump Head A

PDS-100 Pump Head B

10% Methanol

Deionized Water

T

Thermocouple

Direct Methanol

Fuel Cell

Waste Container

T

Thermocouple

TThermocouple

Computer

DAQ

T-1

T-2

T-3

AR-1

AR-2

AR-6

AR-3

Solution

Heater AR-7

AR-8

AW-2

AW-1

AR-5

AR-4

FCV-1

CV-1

CV-2

CV-3

ACP-2AC ACP-1

Fuel Cell Test Stand

FCL-1 CR-1

ComputerCI-1

CI-2

Multimeter

FCI-1

P-1

Figure 3-4. P&ID for University of North Florida standard test station and methanol control

attachment.

Fuel Cell Hardware

The eight cell stack shown in Figure 3-5 with an active area of 15.5 cm² per cell was used

to analyze the variation in transient performance relative to the feed methanol concentration. The

MEA is composed of an in-house hydrocarbon based membrane with a catalyst loading of 3.7

mg/cm² on the anode and two mg/cm² on the cathode. The composition of the catalyst on the

anode is a 50/50 atomic ratio mixture of platinum/ruthenium, while the catalyst on the cathode is

a solitary platinum catalyst.

A flow channel plate is placed between each MEA in order to deliver various

concentrations of methanol to the anode and oxygen (air) to the cathode. The anode and cathode

Page 32: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

32

each have a unique flow pattern that has been optimized for reactant delivery. The anode side of

the flow channel plate is characterized with a serpentine flow channel, while the cathode has an

open straight channel design. The simplicity of the open cathode allows for easy heat removal,

however the cathode is typically subjected to operation at ambient conditions (pressure,

temperature, humidity). In order to acquire repeatable, consistent data, a PID temperature

controller was used with a solution heater to preheat the anode solution entering the stack so that

a uniform stack temperature could be achieved. For all of the testing that was conducted, the

minimum stoichiometric flow rates for the anode and cathode were ten and three respectively.

Figure 3-5. Compressed eight cell fuel cell stack.

Experimentation

Based on the data that is presented in Figure 3-6, the DMFC performance has a strong

correlation relative to the feed methanol concentration. However, the optimal methanol

concentration for these data is approximately 0.70 M. Without a reference of which side of the

optimal concentration that the fuel cell is operating at, it is very difficult to determine the

methanol concentration based on a static load measurement. Therefore, the transient response of

the DMFC was evaluated for sensitivity to methanol concentration.

Page 33: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

33

Figure 3-6. DMFC methanol concentration sensitivity with static loading.

Active Load Method

During the literature review, Argyropoulos, Scott and Taama [35] revealed a variation in

the cell voltage response with respect to methanol concentration when a DMFC was brought

from a loaded to an unloaded (OCV) condition. It is less than desirable to interrupt fuel cell

power production every time the methanol concentration needs to be determined. With the

intention of eliminating the unloaded condition to determine methanol concentration, an active

load approach was tested initially. In order to determine if a measurable correlation between

methanol concentration and the transient behavior of the fuel cell exists, the fuel cell was

operated with various transient electrical loads. Initially, the load was oscillated in constant

voltage mode at different frequencies and magnitudes. By comparison, the published literature

conducted their testing in constant current mode. In order to prevent cell reversal, constant

60.0

70.0

80.0

90.0

100.0

110.0

120.0

130.0

140.0

150.0

160.0

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8

Cu

rre

nt

De

nsi

ty (

mA

/cm

²)

Solution Concentration (Moles/Liter)

Current Density vs. Solution Concentration (Cv=0.35 V)

55.0 50.0 45.0

Page 34: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

34

voltage load steps were used. An exemplar transient load profile shown in Figure 3-7 identifies

the critical points and terminology for data analysis.

Figure 3-7. DMFC current density response with load change from 0.40 V to 0.35 V.

As shown in Figure 3-7, the fuel cell was electrically loaded from a constant voltage of

0.40 V to 0.35 V. The transient fuel cell current that corresponds to the voltage load steps was

analyzed. Immediately after the electrical load is changed, the fuel cell current experiences a

rapid transient. Due to the actuation speed of the electrical load bank, the initial peak, identified

as the peak cutoff in Figure 3-7, is more than likely an artifact of the electrical load bank

overshooting its target voltage. Therefore, any of the transient data prior to this peak was

disregarded for data analysis. In addition to the peak cutoff point, the maxima and minima for

each voltage step is labeled.

Page 35: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

35

Experimentation

An algorithm was developed in LabVIEW to determine the cutoff point, maximum and

minimum current for each electrical load change. The data that exists between these two points

was used to perform the analysis for methanol concentration sensitivity. During operation, the

DMFC is typically operated at a voltage between 0.35 and 0.40 V in order to maximize output

power and efficiency. Therefore, the load changes for experimentation were oscillated between

0.35 and 0.40 V. Furthermore, 10 seconds (5 seconds for each load step) were used for the load

oscillation period in order to obtain frequent concentration measurements. The DMFC was tested

at various methanol concentrations ranging from 0.20 M – 1.60 M in 0.20 M increments.

The DMFC was tested for ten minutes for each methanol concentration setpoint. As shown

in Figure 3-8, for low methanol concentrations (0.20 M), a very erratic behavior exists. Figure

3-9, with an operation methanol concentration of 0.60 M, indicates a substantial change in the

transient current density response relative to the 0.20 M sample. Finally, Figure 3-10 represents

the upper level of methanol concentration setpoints, with a more subtle change with respect to

the 0.60 M test case.

Figure 3-8. Transient current for a DMFC with load oscillations from 0.35 V to 0.40 V at 0.20 M

and 50°C.

Page 36: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

36

Figure 3-9. Transient current for a DMFC with load oscillations from 0.35 V to 0.40 V at 0.60 M

and 50°C.

Figure 3-10. Transient current for a DMFC with load oscillations from 0.35 V to 0.40 V at 1.60

M and 50°C.

Results

The data presented in Figure 3-11 and Figure 3-12 represent one instance of the unfiltered

(peak cutoff not removed) fuel cell current with load changes to 0.35 V and 0.40 V respectively.

Figure 3-13 and Figure 3-14 are the same data that was presented in Figure 3-11 and Figure 3-12,

however the data has been filtered and normalized in order to facilitate data analysis. Based on

an initial qualitative analysis, are large disparity forms between methanol concentration less than

Page 37: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

37

0.50 M and greater than 0.50 M. For the 0.35 V case, the test that were conducted with the lower

concentrations (<0.50 M) decreased, while the testing that was conducted with the higher

methanol concentrations increased. For the 0.40 V testing, the opposite relationship was

established. With both the 0.35 V and 0.40 V cases, the difference between the transient response

with methanol concentrations greater than 0.80 M was minimal.

Figure 3-11. Unfiltered fuel cell transient response when electrically loaded from 0.40 V to 0.35

V at 50°C.

Figure 3-12. Unfiltered fuel cell transient response when electrically loaded from 0.35 V to 0.40

V at 50°C.

60

80

100

120

140

160

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0

Cu

rre

nt

De

nsi

ty (

mA

/cm

²)

Elapsed Time (s)

Fuel Cell Current Transient Response at 0.35 V

0.2 M 0.4 M 0.6 M 0.8 M 1.0 M 1.2 M 1.4 M 1.6 M

0

50

100

150

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0

Cu

rre

nt

De

nsi

ty (

mA

/cm

²)

Elapsed Time (s)

Fuel Cell Current Transient Response at 0.40 V

0.2 M 0.4 M 0.6 M 0.8 M 1.0 M 1.2 M 1.4 M 1.6 M

Page 38: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

38

Figure 3-13. Filtered and normalized fuel cell transient response when electrically loaded from

0.40 V to 0.35 V at 50°C.

Figure 3-14. Filtered and normalized fuel cell transient response when electrically loaded from

0.35 V to 0.40 V at 50°C.

For each oscillation, a power curve fit was established of the form listed in Equation 3-1.

The exponential b-coefficient is representative of the slope of the curve. The power curve fit

coefficients for each methanol concentration cycle was average and is shown in Figure 3-15. The

active load measurement provides an acceptable resolution with concentrations ranging from 0.2

to 0.8 M, however this limited range for use in a DMFC is unsatisfactory for robust operation.

100.0102.5105.0107.5110.0112.5115.0117.5120.0122.5125.0

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0

Current Density Responses at 0.35 V

0.20 M 0.40 M 0.60 M 0.80 M 1.00 M 1.20 M 1.40 M 1.60 M

110.0

120.0

130.0

140.0

150.0

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0

Current Density Responses at 0.40 V

0.20 M 0.40 M 0.60 M 0.80 M 1.00 M 1.20 M 1.40 M 1.60 M

Page 39: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

39

3-1

Figure 3-15. Current Density Transient Response Power Curve Fit Coefficients for 0.35 V and

0.40 V at 50°C.

Passive Load Method

Based on the reviewed literature [35] and the results from the active method, it was

determined that the load profile that could provide the greatest resolution for methanol

concentration determination is when the stack is operated from a loaded (120-150 mA/cm²) to an

unloaded (OCV) operating point. This load profile is a realistic option for DMFC system

integration. In order to minimize on-state degradation due to cathode catalyst oxidation, a DMFC

is subjected to a periodic rest cycle during operation. This cycle involves removing the load from

the fuel cell stack, allowing the stack voltage to reach OCV and then reapplying the load with no

oxygen flow to the cathode so that the cathode potential is reduced.

Experimentation

In order to gain a better understanding of the voltage response, the stack was operated at

various methanol concentrations. The initial experiments revealed three distinct paths for

determining a correlation between the transient voltage response and methanol concentration. In

-1.0E-02-7.5E-03-5.0E-03-2.5E-030.0E+002.5E-035.0E-037.5E-031.0E-021.3E-021.5E-02

-3.E-02

-1.E-02

1.E-02

3.E-02

5.E-02

7.E-02

0.00 0.20 0.40 0.60 0.80 1.00 1.20 1.40 1.60 1.80

0.3

5 V

Po

we

r C

urv

e F

it C

oe

ffic

ien

t

0.4

0 V

Po

we

r C

urv

e F

it C

oe

ffic

ien

t

Current Density Transient Response Power Curve Fit Coefficients

0.40 V Power Curve Fit Coefficient 0.35 V Power Curve Fit Coefficient

Page 40: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

40

Figure 3-16, a representative voltage response is shown with the Max OCV, Max OCV rise time

and OCV decay slope identified.

Figure 3-16. Representative fuel cell voltage response from loaded to unloaded operating point.

As shown in Figure 3-17, a clear discrepancy is visible for methanol concentrations from

0.60 M to 1.60 M. Based on a qualitative analysis, the rise time provided the least amount of

resolution for determining methanol concentration. The time to reach max open circuit voltage

varied from three to 10 seconds. However, at higher methanol concentrations (CFeed > 1.0 M), the

difference between rise times was less significant providing less measurement resolution.

The Max OCV provides greater resolution than the MAX OCV rise time especially when

comparing concentrations from 0.80 M to 1.60 M. Unfortunately, the degradation that occurs

within the stack during operation has an intermediate effect on the open circuit voltage.

Page 41: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

41

Therefore, it was determined that Max OCV is not ideal for long term methanol concentration

determination. Finally, the OCV decay slope was chosen as a primary candidate for data

analysis. This method for methanol determination provides a distinct variation for the various

methanol concentrations providing great resolution.

Figure 3-17. Fuel cell air starve cycle with 0.60, 0.80 and 1.60 molar concentration at 50 °C.

The fuel cell stack was operated at eight different methanol concentrations in order to

determine a relationship with respect to OCV decay slope. For each concentration that was

tested, the stack was operated at three different temperatures and four different current densities.

The fuel cell stack was operated for ten cycles on an accelerated rest cycle for each configuration

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0 11.0 12.0 13.0 14.0 15.0 16.0 17.0 18.0

Ce

ll V

olt

age

(V

)

Elapsed Time (s)

DMFC OCV Response with No Air Flow

0.60 M 0.80 M 1.00 M 1.20 M 1.40 M 1.60 M

Page 42: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

42

of methanol concentration, temperature and current density. The abbreviated rest cycle is shown

in Figure 3-18, which consists of four distinct sub-cycles.

Figure 3-18. Typical load cycle for DMFC OCV decay slope testing.

The cycle begins with a constant current density step (sub-cycle 1) for two minutes. The

next step following sub-cycle one is the OCV response step (sub-cycle 2) with the oxygen flow

rate set to zero. The duration of this cycle varied based on the methanol concentration from 15 –

90 seconds. The data that was gathered during this sub-cycle was used for all of the data

analysis. During sub-cycle 3, the voltage on the stack is “pulled down” to a low voltage (<0.10

V) for thirty seconds in order to reduce the potential on the cathode. Sub-cycle four is where the

oxygen flow rate is turned back on and the voltage climbs back to open circuit voltage. The OCV

response in sub-cycle four has a different shape compared to the OCV response in sub-cycle 2.

Page 43: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

43

The presence of continuously flowing oxygen during sub-cycle four allows for the OCV to

continue climbing while the absence of oxygen flow during sub-cycle two reduces the cell

voltage. Finally the cycle returns back to the beginning at sub-cycle 1. The iteration of this cycle

occurred over one thousand times in order to collect data for the various operating conditions

(temperature, current density, feed methanol concentration).

Results

The variation in OCV response is believed to be driven primarily by the consumption of

oxygen at the cathode. Based on the concentration of methanol at the cathode catalyst layer, the

oxygen will be consumed proportionally and subsequently the cell voltage will decay. However,

if oxygen continues to flow to the cathode, similar to the second OCV sub-cycle, the oxygen will

never be entirely consumed resulting in a less dramatic signal for methanol concentration

determination. Therefore, the first OCV sub-cycle, where the flow of oxygen is brought to zero,

was used for the OCV decay slope analysis.

A strategy was developed in LabVIEW in order to determine the OCV decay slope. For

each OCV response sub-cycle, a linear regression was performed shortly after the peak OCV was

achieved. The linear regression is composed of two coefficients. The “b” coefficient from

Equation 3-2 defines the slope of the line, while the “c” coefficient defines the offset on the y-

axis. The “b” coefficient for any of the linear regressions that were performed will now be

referred to as the OCV decay slope.

3-2

In Figure 3-19, the OCV decay slope for methanol concentrations from 0.60 M to 2.0 M

are shown at a nominal load and operating temperature. The OCV decay slopes have been made

Page 44: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

44

positive for enhanced viewing on the logarithmic scale. The error bars display the first standard

deviation based on the ten cycles that were performed for each setpoint.

Figure 3-19. OCV decay slope at 120 mA/cm² at various methanol concentrations with error bars

indicating first standard deviation for the measurements that were conducted for each

cycle.

The OCV decay slope increases at nearly an order of magnitude with every 0.20 M change

in methanol concentration. Although less severe than what was observed for the active load

measurements, as the concentration approaches 1.60 M, the growth in the OCV decay slope

tapers off; the amount of methanol found at the cathode catalyst layer reaches a saturation point,

due primarily to the diffusion properties of methanol through the MEA.

1.0E-06

1.0E-05

1.0E-04

1.0E-03

1.0E-02

1.0E-01

1.0E+00

0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 2.2

Ab

solu

te V

alu

e o

f O

CV

De

cay

Slo

pe

(V

/s)

Solution Concentration (M)

OCV Decay Slope at 120 mA/cm² at 50°C

Page 45: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

45

Figure 3-20. OCV decay slope at various current densities with 50°C stack temperature.

Figure 3-20 indicates a weak function of OCV decay slope with respect to current density.

With the exception of the discrepancies at 0.80 M, the curves for all current densities follow a

similar trend. One of the key differences between the 40 and 80 mA/cm² current densities

compared to the 120 and 140 mA/cm² current densities at 0.80 M are the max OCV rise times.

As shown in Table 3-1, the OCV rise time for 120 and 140 mA/cm² is an order of magnitude

different when compared to the OCV rise time at 40, 80 mA/cm² and all of the data points

collected at 0.60 M.

Table 3-1. OCV rise time at 0.60 M and 0.80 M.

Current Density (mA/cm²) OCV Rise Time (s)

0.60 M 0.80 M

40 83.7 35.2

1.0E-05

1.0E-04

1.0E-03

1.0E-02

1.0E-01

1.0E+00

0.40 0.60 0.80 1.00 1.20 1.40 1.60 1.80 2.00 2.20

OC

V D

eca

y Sl

op

e (

V/s

)

Solution Concentration (M)

OCV Decay Slope at Various Current Densities at 50°C

40 mA/cm² 80 mA/cm² 120 mA/cm² 140 mA/cm²

Page 46: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

46

Current Density (mA/cm²) OCV Rise Time (s)

0.60 M 0.80 M

80 78.0 12.4

120 73.4 8.0

140 75.1 7.2

One potential explanation for the divergence in data points is localized flooding at the

cathode. With increasing methanol concentration, the amount of methanol crossover (internal

heating) also increases, therefore increasing the amount of cooling air required by the stack to

maintain constant temperature. In addition, with increasing current density, the cooling

requirements for the stack are also increased. At lower cathode flow rates, the stack is more

prone to flooding [36], resulting in blocked reaction sites at the cathode. Any reduction of

reaction sites would reduce the effective reactivity with oxygen and therefore the consumption

rate of oxygen at the cathode. At the prescribed concentrations and current densities, it is

believed that the DMFC is operating on a knife edge with the two competing effects of cathode

flooding and oxygen consumption.

The curves presented in Figure 3-21 display a similar trend to what has been observed in

Figure 3-19 and Figure 3-20, with the OCV decay slope increasing with methanol concentration.

The stack current density was held constant at 120 mA/cm² while varying the stack temperature

and concentration. A noticeable change in OCV decay slope is visible for varying temperature.

With increasing temperature, the OCV decay slope (diffusion) also increases. This agrees well

with other diffusion experiments that indicate methanol crossover’s high dependency on

temperature [37]. Similar to what was described for Figure 3-19, at higher methanol

concentrations, the diffusion of methanol through the MEA appeared to reach a maximum at

1.80 M irrespective of temperature.

The use of OCV decay slope provides a repeatable, accurate measurement for the

determination of methanol concentration in a DMFC. In addition, this measurement technique

Page 47: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

47

can easily be integrated into any existing DMFC system with little system reconfiguration. In

order to reduce on-state degradation, a rest (air starve) is conducted multiple times per hour to

remove any buildup of oxides on the cathode catalyst layer [38]. This would be an ideal time to

determine methanol concentration because generally the stack voltage is allowed to go to OCV

before entering the air starved load condition. However, at reduced temperatures and methanol

concentrations, the OCV decay slope can require as much as 100 seconds to be determined. In

order to implement such a methanol sensing technique, the OCV rise time should not be greater

than 15 seconds for most operating conditions. This time is typical for DMFC OCV rest cycles.

As shown in Figure 3-22, the max OCV rise time falls in an acceptable range for methanol

concentrations 1.0 M and greater. In order to operate with rest cycles less than 15 seconds,

another technique must be used in order to determine the methanol concentration at lower

temperatures or methanol concentrations.

Figure 3-21. OCV Decay slope with constant current density and variable concentration and

temperature.

1.0E-06

1.0E-05

1.0E-04

1.0E-03

1.0E-02

1.0E-01

1.0E+00

0.40 0.60 0.80 1.00 1.20 1.40 1.60 1.80 2.00 2.20

OC

V D

eca

y Sl

op

e (

V/s

)

Solution Concentration (M)

OCV Decay Slope at 120 mA/cm²

45 °C 50 °C 55 °C

Page 48: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

48

Figure 3-22. Max OCV rise time at constant current density with varying temperature and

methanol concentration.

For all previous measurements, the OCV decay slope was the method used to determine

methanol concentration. However, due to the unacceptable rise time for Max OCV at low

temperatures and methanol concentrations, the rise slope OCV was evaluated as an alternative to

measuring OCV decay slope. For the experiment, the maximum OCV rise time was defined at 15

seconds. If the max OCV was not reached within 15 seconds, the OCV rise slope was determined

for the last three seconds of OCV. For each temperature, methanol concentration and current

density configuration, the OCV rise slope was evaluated ten times.

As shown in Figure 3-23, the OCV rise slope at 40 mA/cm² is characterized with a linear

relationship relative to methanol concentration. However, as indicated by the error bars, the first

standard deviation for the sample is considerably higher than the measured OCV decay slopes. In

addition, the point with a solution concentration of 1.0 M and a stack temperature of 45°C

0

10

20

30

40

50

60

70

80

90

100

110

120

0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2

Max

OC

V R

ise

Tim

e (

s)

Solution Concentration (M)

Max OCV Rise Time at 80 mA/cm²

45°C 50°C 55°C

Page 49: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

49

exceeds an order of magnitude of the linear trend that is established by the existing points. As

previously discussed, these anomalies could very well be caused by the fuel cell stack operating

on a knife edge. At this condition, the stack could potentially be experiencing flooding, which

would affect OCV response.

Figure 3-23. OCV Rise Slope held at constant current density (40 mA/cm²) with variable

concentration and stack temperature with error bars indicating single standard

deviation.

The OCV rise slope at a current density of 40 mA/cm² provides much uncertainty in

methanol concentration determination. At a current density of 120 mA/cm², the uncertainty in

methanol concentration determination increased with a weaker correlation between methanol

concentration and OCV rise slope. The combination of weak correlation, with respect to

methanol concentration and high uncertainty, make the OCV rise slope an unsatisfactory method

for determining methanol concentration.

0.0E+00

5.0E-04

1.0E-03

1.5E-03

2.0E-03

2.5E-03

3.0E-03

0.5 0.6 0.7 0.8 0.9 1.0 1.1

OC

V R

ise

Slo

pe

(V

/s)

Solution Concentration (M)

OCV Rise Slope at 40 mA/cm²

45°C 40°C 35°C

Page 50: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

50

Figure 3-24. OCV Rise Slope held at constant current density (120 mA/cm²) with variable

concentration and stack temperature with error bars indicating single standard

deviation.

The failure of the OCV decay slope to provide reliable measurements for methanol

concentration determination leaves only a few methods for improving OCV decay slope response

time. One possibility is to operate the fuel cell stack at an elevated temperature just before rest.

Operation at an elevated temperature (55-60°C) can significantly shorten the max OCV rise time

to durations that would be acceptable for methanol concentration determination in a DMFC

system.

Another method for improving the max OCV rise time is to operate at an elevated

methanol concentration. At an elevated methanol concentration (CFeed > 1.0 M), the ratio of

methanol to oxygen at the cathode catalyst layer would exist at a considerably higher level than

for 0.60 M or 0.8 M. As previous data suggests, the OCV rise time would be significantly less.

However, with an increase in methanol concentration, the stack would operate at a less efficient

point resulting in less net power and poor fuel utilization.

0.0E+00

1.0E-03

2.0E-03

3.0E-03

4.0E-03

5.0E-03

6.0E-03

0.5 0.6 0.7 0.8 0.9 1.0 1.1

OC

V R

ise

Slo

pe

(V

/s)

Solution Concentration (M)

OCV Rise Slope at 120 mA/cm²

45°C 40°C 35°C

Page 51: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

51

The last method for improving the max OCV rise time at reduced concentrations and

temperatures would involve the modification of the MEA. The supporting theory for long

duration max OCV rise times is the lack of oxygen consumption at the cathode catalyst due to

poor diffusion and flooding. A single MEA on the stack could be optimized for oxygen transport,

methanol crossover and reduced flooding effects. This would promote the consumption of

oxygen at the cathode catalyst layer with reduced max OCV rise times.

Page 52: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

52

CHAPTER 4

VERIFICATION MODEL

A basic 1-D transient model was created in order to gain a better understanding of the

diffusion phenomena and transients measured with the fuel cell test station. The direct methanol

fuel cell has many side reactions and transport dependencies that take place within the MEA,

therefore many of the complex, secondary factors (membrane swelling, two phase flow, electro

osmotic drag, catalyst loading, etc.) that contribute to DMFC performance have been neglected

in order to simplify the model. The formulation of the model and all supporting boundary

conditions and assumptions are outlined in Appendix A.

The formulas that have been compiled to solve the 1-D non-homogeneous transient model

were entered into MATLAB. This model enabled the prediction of methanol and oxygen

concentration transient behavior through the MEA. The MATLAB code for the model can be

found in Appendix B.

Methanol Concentration Distribution

Steady State Concentration Distribution

The model was initially assessed in a steady state configuration. As shown in Figure 4-1,

the model was tested at a fuel cell current density of 150 mA/cm² with a feed concentration of

0.80 M. The model reveals a non-linear methanol concentration gradient across the anode

catalyst layer. The non-linear portion is a result of the methanol that is consumed for electrical

power production by the fuel cell at 150 mA/cm². At the cathode catalyst layer, the methanol

concentration approaches zero. The methanol crossover can be calculated based on the methanol

concentration gradient across the cathode catalyst layer or membrane layer.

In Figure 4-2, the model was executed for various feed concentration at 150 mA/cm². For a

feed concentration of 0.40 or lower, the methanol concentration at the anode catalyst layer is

Page 53: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

53

below 0, which would result in a fuel starvation condition due to insufficient methanol. If the

effects of relatively low stoichiometric ratios were taken into account, a fuel starvation condition

would be realized at the end of the cell due to reduced feed methanol concentration relative to

the beginning of the cell. The variation of methanol concentration at the entrance of the cathode

catalyst layer is an indication of the variation in methanol crossover for various concentrations.

The methanol crossover for each methanol concentration is summarized in Table 4-1.

Figure 4-1. Results from model for MEA methanol concentration distribution for 0.80 M feed

concentration at a current density of 150 mA/cm².

Figure 4-2. Results from model for MEA methanol concentration distribution for various feed

concentrations at a current density of 150 mA/cm².

0 0.005 0.01 0.015 0.02 0.025 0.03 0.0350

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

MEA Position (cm)

Meth

an

ol C

on

cen

trati

on

(M

)

MEA Methanol Concentration Distribution for 0.80 M Feed Concentration at 150 mA/cm²

0 0.005 0.01 0.015 0.02 0.025 0.03 0.0350

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

MEA Position (cm)

Meth

an

ol C

on

cen

trati

on

(M

)

MEA Methanol Concentration Distribution for Various Feed Concentrations at 150 mA/cm²

0.40 M

0.60 M

0.80 M

1.00 M

1.20 M

1.40 M

1.60 M

Anode Diffusion Layer

Anode Catalyst Layer

Membrane Layer

Cathode Catalyst Layer

Page 54: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

54

Table 4-1. Summary of methanol crossover results from model for various feed methanol

concentrations with a current density of 150 mA/cm². Feed Concentration (M) Crossover Current Density (mA/cm²)

0.6 8.2

0.8 20.1

1.0 32.0

1.2 43.9

1.4 55.8

1.6 67.7

Transient Concentration Distribution

The next set of test conditions that were executed using the model evaluated the transient

response of the methanol concentration distribution. The model was used to determine the

methanol concentration response with a load profile where the load is instantly removed. As

shown in Figure 4-3, the methanol concentration distribution requires approximately 75 seconds

to reach equilibrium. At the cathode catalyst layer, the non-linear concentration gradient is

evident when the load is still engaged at t = 0 s. Once the load is removed (t > 0), the methanol

concentration quickly increases and a linear concentration gradient is visible.

Figure 4-3. Results from model for MEA transient concentration distribution from a current

density of 120 mA/cm² to 0 mA/cm² at a feed concentration of 0.80 M.

The transient methanol concentration distribution for relatively high feed methanol

concentrations is shown in Figure 4-4. For a feed methanol concentration of 1.60 M, the

0 0.005 0.01 0.015 0.02 0.025 0.03 0.0350

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

MEA Position (cm)

Meth

an

ol C

on

cen

trati

on

(M

)

MEA Methanol Concentration Distribution from 120 mA/cm² to 0 mA/cm² at 0.80 M

0.0 s (120 mA/cm²)

15.0 s (0 mA/cm²)

30.0 s (0 mA/cm²)

45.0 s (0 mA/cm²)

60.0 s (0 mA/cm²)

75.0 s (0 mA/cm²)

Page 55: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

55

methanol consumption for electrical power production is less significant. Therefore, the

methanol concentration at the cathode catalyst is similar to that for the unloaded points and is

relatively high.

Figure 4-4. Results from model for MEA transient concentration distribution from a current

density of 120 mA/cm² to 0 mA/cm² at a feed concentration of 1.60 M.

The response of the methanol concentration at the cathode catalyst layers are shown in

Figure 4-5. For each feed concentration, the response is similar in shape however the magnitude

of the concentration is more than double for a feed concentration of 1.6 compared to 0.80 M.

Figure 4-5. Model results for transient methanol concentration response at cathode catalyst layer

from a current density of 120 mA/cm² to 0 mA/cm².

0 0.005 0.01 0.015 0.02 0.025 0.03 0.0350

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

MEA Position (cm)

Meth

an

ol C

on

cen

trati

on

(M

)

MEA Methanol Concentration Distribution from 120 mA/cm² to 0 mA/cm² at 1.60 M

0.0 s (120 mA/cm²)

15.0 s (0 mA/cm²)

30.0 s (0 mA/cm²)

45.0 s (0 mA/cm²)

60.0 s (0 mA/cm²)

75.0 s (0 mA/cm²)

0

0.02

0.04

0.06

0.08

0.1

0.12

0 10 20 30 40 50 60 70 80

Me

than

ol C

on

cen

trat

ion

(M

)

Elapsed Time (s)

Transient Methanol Concentration Response at the Cathode Catalyst Layer from 120 mA/cm² to 0 mA/cm²

0.80 M 1.60 M

Page 56: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

56

Transient Oxygen Concentration Distribution

Due to the relatively slow response of the methanol concentration distribution, the effect of

the transient methanol concentration distribution alone is not a viable means for methanol

concentration determination. However, if the flow of cathode air is removed from the fuel cell,

the oxygen on the cathode would quickly be consumed due to the methanol crossover. Using the

methanol crossover model that has been developed, a basic oxygen consumption model was

established in order to determine if the transient response of the cathode would provide a more

rapid response. Based on the methanol concentration response and the diffusion coefficient at the

cathode catalyst layer, the transient diffusion of methanol (methanol crossover) can be

calculated. Assuming that for each mole of methanol that is consumed at the cathode catalyst

layer, 1.5 moles of oxygen is consumed; the mean oxygen content at the cathode can be

calculated.

Even though an active air source is not present, the oxygen that is consumed at the cathode

can be replenished by means of natural convection or diffusion, due to the open cathode design

of the modeled fuel cell. A basic linear model was used to accommodate for the oxygen that is

replenished based on open cathode passive fuel cells. In addition to the replenished oxygen, the

amount of oxygen stored in the cathode flow channels was also taken into account. As shown in

Figure 4-6, the oxygen that is consumed at the cathode catalyst layer due to methanol crossover

occurs relatively fast. With increasing methanol concentration, the consumption rate of oxygen

on the cathode increases, resulting in shorter durations of high oxygen content. However, the

timescales for oxygen consumption for an open cathode are long compared to a closed cathode

where replenished oxygen would not be available. In order to emphasize the transient oxygen

concentration response for all methanol concentrations, a logarithmic time scale was used on the

x-axis.

Page 57: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

57

Figure 4-6. Model results for the mean transient oxygen content in the cathode.

The cathode potential on a DMFC is strongly dependent on the concentration of oxygen at

the cathode [36]. The higher the partial pressure of oxygen at the cathode, the higher the cathode

potential. Therefore with a transient response in the concentration of oxygen at the cathode, the

voltage of the DMFC will also experience a dramatic transient.

Modeling Summary

Data provided by the model suggests that the dynamic behavior of the methanol

concentration distribution is not rapid enough for practical use in a DMFC as a sensor-less

measurement technique. Therefore, the air flow was removed from the cathode in the model in

order to accelerate the effects of methanol on the DMFC performance. The data from the

transient 1-D model agrees with the data that was collected experimentally and highlights the

importance of the removal of an active oxygen supply at the cathode. Furthermore, the model

showed that the feed methanol concentration has a greater influence on the anode methanol

concentration distribution than the operating current density.

0.0%

5.0%

10.0%

15.0%

20.0%

25.0%

0.1 1 10 100

Oxy

gen

Co

nce

ntr

atio

n

Elapsed Time (s)

Transient Mean Cathode Oxygen Concentration

1.6 M 1.4 M 1.2 M 1.0 M 0.8 M 0.6 M 0.4 M

Page 58: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

58

CHAPTER 5

SYSTEM INTEGRATION

The University of North Florida has the capability to operate an unpackaged DMFC

system in a brassboard configuration. The brassboard platform shown in Figure 5-1 enables

maximum flexibility for in-situ testing of components and system control strategies. The OCV

decay slope method was implemented into the brassboard system in order to determine methanol

concentration during operation. The DP4 (demonstration prototype 4) brassboard was designed

to operate on ten minute rest cycles. Assuming the methanol concentration can be determined

during each rest cycle, the methanol concentration must be determined during non-rest operation.

Figure 5-1. UNF 20 W DP4 brassboard fuel cell system.

Methanol Concentration Tracking

Methanol Consumption Model

As highlighted in Figure 5-2, the methanol consumed by the stack to produce electrical

current and the methanol consumed at the cathode catalyst due to methanol crossover are the two

major contributors to methanol consumption in the DP4 system. The minor sources of methanol

consumption include leakage through the CO2 gas liquid separator and the solution storage tank.

Page 59: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

59

However, these sources of methanol consumption are negligible relative to the rates of methanol

consumption due to crossover and power production.

A

N

O

D

E

M

E

M

B

R

A

N

E

C

A

T

H

O

D

E

Figure 5-2. Simplified methanol consumption and injection model for DP4.

Faradaic oxidation

The calculation of methanol consumption due to electrical current is a straightforward

calculation based on the anode half reaction. Equation 5-1 states that for every six moles of

electrons that are used, one mole of methanol is consumed. Using Faraday’s constant, the

conversion from amperes to moles of electrons can be made. Using these two factors, the final

conversion from amperes to methanol consumption equals 1.73E-6 Moles of

CH3OH/(s·Ampere·cell).

5-1

M

E

T

H

A

N

O

L

C

R

O

S

S

O

V

E

R

CH3OH

Page 60: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

60

Methanol crossover

The methanol consumption due to methanol crossover was calculated based on the amount

of CO2 that was measured on the cathode exhaust using a CO2 analyzer. Carbon dioxide is a

product of the methanol that crosses over to the cathode catalyst and the oxygen that is delivered

to the cathode for the fuel cell reaction. For every mole of methanol that is oxidized at the

cathode catalyst, one mole of CO2 is released into the cathode stream. The methanol

consumption is converted to an equivalent crossover current density in order to simplify

comparison to the stack current density.

Figure 5-3. Representative methanol crossover current density for DP4 stack at various stack

temperatures, methanol concentrations and current densities.

15.0

25.0

35.0

45.0

55.0

65.0

75.0

0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150 160

Stac

k C

urr

en

t D

en

sity

(m

A/c

m²)

Crossover Current Density (mA/cm²)

Methanol Crossover at Various Stack Temperature, Concentration and Current Densities

0.8 M at 45 °C 1.0 M at 45 °C 1.5 M at 45 °C 0.8 M at 50 °C 1.0 M at 50 °C

1.5 M at 50 °C 0.8 M at 55 °C 1.0 M at 55 °C 1.5 M at 55 °C

Page 61: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

61

Methanol crossover is primarily a function of current density, temperature and methanol

concentration. Therefore, these parameters were varied in order to determine the methanol

crossover for the fuel cell stack. The crossover current measurements in Figure 5-3 exhibit a

strong linear function with respect to stack current density. As the stack current density

approaches 0 mA/cm² (OCV), the availability of methanol at the cathode catalyst reaches a

maximum resulting in a maximum of crossover current density. With increased temperature, the

diffusion coefficient for methanol across the MEA also increases resulting in higher methanol

levels at the cathode catalyst layer. The last variable that was tested in order to determine the

amount of methanol crossover rates in the MEA was the sensitivity to methanol concentration.

With increasing methanol concentrations at the anode, the available methanol at the cathode

catalyst would only increase, resulting in a higher methanol crossover rate. The values from

Figure 5-3 were compiled in a 2-D matrix where they could be used in a lookup table.

Methanol Injection Model

In order to accurately track the methanol concentration within the brassboard, the amount

of methanol that is injected into the anode stream must be accounted for in addition to the

methanol that is consumed. Methanol is injected into the brassboard using a single piezoelectric

pump controlled by pulse width modulation (PWM). The pump was characterized at several duty

cycles using a mass balance. As shown in Figure 5-4, the pump has a strong linear relationship

with respect to PWM duty cycle.

During preliminary testing, the accuracy of the methanol injection pump appeared to vary

based on the fuel reservoir level. The fuel reservoir where the methanol is stored is a 500 mL

container. The difference in pressure head for when the fuel reservoir is full, compared to when it

is empty, can be as much as six inches of methanol. In order to more accurately predict the

amount of methanol injected into the system, the pump was characterized at two additional fuel

Page 62: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

62

levels. As shown in Figure 5-5, the pump is strongly affected by the pressure head that is applied

to the inlet.

Figure 5-4. Performance Curve for Single µBase Pump

Figure 5-5. Comparison of µBase Pump Performance versus various inlet pressures.

0.0

0.2

0.4

0.6

0.8

0 5 10 15 20 25 30 35 40 45 50

Me

than

ol I

nje

ctio

n P

um

p F

low

Rat

e

(mL/

min

)

Pump Duty Cycle (%)

µBase Pump Performance

0.0

0.2

0.4

0.6

0.8

0 5 10 15 20 25 30 35 40 45 50

Me

than

ol I

nje

ctio

n P

um

p F

low

Rat

e

(mL/

min

)

Pump Duty Cycle (%)

µBase Pump Performance vs. Inlet Pressures

No Inlet Head 4" of Negative Inlet Head

Page 63: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

63

Methanol Concentration Determination

The empirical data that was collected for methanol crossover and the methanol injection

pump were used in conjunction with the methanol consumption model based on electrical current

in order to determine a net mass balance of methanol for the system. To determine the methanol

concentration within the system, an accurate model must be used to determine both the quantity

of methanol and water in the system.

The anode solution reservoir tank for the system features a tank level sensor that utilizes

twelve points of conduction in order to determine the amount of solution in the reservoir tank. In

addition to the reservoir tank, anode solution is stored in the fuel cell stack, gas liquid separator,

methanol sensor and the silicone lines that are used to connect each of the components. The fuel

cell stack generates CO2 gas in the anode stream during power production, therefore the

displacement of gas in the components must be taken into consideration in order to accurately

account for the solution volume in the system.

Based on the liquid inventory , the previous methanol concentration ( , the methanol

consumption due to electrical current ( ̇ and methanol crossover ( ̇ , the methanol injected

by the feed pump ( ̇ , the molar mass of methanol and the amount of time

between the concentration measurements, the new calculated methanol concentration can

be calculated.

( ̇ ̇ ̇ )

5-2

Brassboard Operation

The results shown in Figure 5-6 are operation data from the brassboard using the sensor-

less OCV decay slope methanol sensing technique. The system was able to maintain a constant

methanol concentration without methanol excursions greater than 0.15 M for greater than

Page 64: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

64

eighteen hours. This data suggests that sensor-less methanol sensing techniques are a feasible

method for continuous, reliable DMFC operation.

Figure 5-6. UNF DP4 brassboard operation using sensor-less methanol sensing techniques.

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

0 2 4 6 8 10 12 14 16 18 20

Me

than

ol C

on

cen

trat

ion

(M

)

Elapsed Time (Hrs)

Actual Concentration (M) Sensorless Concentration (M)

Page 65: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

65

CHAPTER 6

CONCLUSIONS

A new method for determining methanol concentration in a direct methanol fuel cell

(DMFC) system was evaluated at the University of North Florida Fuel Cell Laboratory. In this

study, a multilevel approach was used to move from concept to implementation. Initially, a

literature review revealed the sensitivity of DMFC transient open circuit voltage (OCV) response

with respect to methanol concentration.

Initial testing was performed to evaluate various parameters of the DMFC transient OCV

and their sensitivity to feed methanol concentration. Preliminary testing revealed that the OCV

decay slope offered the greatest resolution for methanol concentrations from 0.60 M to 1.60 M.

All future testing used OCV decay slope as a metric for methanol concentration determination.

The OCV decay slope was mapped for eight different concentrations, four different current

densities and three different operating temperatures, resulting in 96 unique operating points. For

each operating point, the DMFC was cycled 10 times in approximately three minute cycles. The

testing revealed a high resolution and repeatability in the OCV decay slope for methanol

concentrations ranging from 0.60 M to 2.00 M. For each change in methanol concentration of

0.20 M, the OCV decay slope changed by nearly one order of magnitude, while the first standard

deviation of the sample set remained relatively low, with less than 5% measurement error.

The sensitivity of OCV decay slope with respect to current density was quite low with the

exception of the points at reduced current densities and methanol concentrations. It is believed

that the irregularity in the OCV decay response is a result of localized flooding in the cathode. At

lower current densities and feed methanol concentrations, the amount of cathode cooling air that

is required to maintain the same operating temperature is less, therefore increasing the likelihood

of localized flooding at the cathode.

Page 66: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

66

The sensitivity of OCV decay slope with respect to operating temperature was much higher

when compared to the sensitivity to current density. With increasing temperature, the OCV

decay slope also increased due to the higher level of diffusion.

The use of OCV decay slope provides a repeatable, accurate measurement for the

determination of methanol concentration in a DMFC. However, the acquisition time for OCV

decay slope can be as much as 90 seconds for low methanol concentrations (CFeed < 0.80 M) at

reduced operating temperatures (T < 50°C). The OCV rise slope was used as an alternative

measurement technique to reduce the acquisition time for operation at low methanol

concentrations with reduced stack temperatures. Unfortunately, there was a very weak

correlation between methanol concentration and OCV rise slope with very poor repeatability.

Other techniques were recommended, however they were not tested. They included changing the

DMFC operating profile to higher methanol concentrations and/or temperature during sensor-less

methanol detection, or optimization of the MEA cathode for OCV decay measurements.

A simplified 1-D transient model was developed to approximate the transient methanol

concentration distribution and verify the findings that were observed in previous experiments.

The initial model revealed a sluggish response time for the methanol concentration distribution

in the MEA. In order to better capture the phenomena that were occurring with respect to

methanol concentration, the computer model was modified to account for an open cathode with

the active air supply removed. The lack of an active air supply accelerated the depletion of

oxygen on the cathode due to methanol crossover resulting in timescales that were compliant

with the DMFC open circuit voltage response times measured. These results agreed well with the

measured data.

Page 67: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

67

The final level of testing was implementation of the sensor-less OCV decay technique into

a 20 W DMFC brassboard system. Integration of OCV decay measurements into a DMFC

system is trivial. During operation, the DMFC must undergo a periodic rest to remove the oxides

that have accumulated on the cathode catalyst. This is achieved by removing the active air supply

from the cathode and pulling the voltage down on the DMFC. The OCV decay measurement is

in essence a free signal that can be gathered during an essential part of the DMFC’s operation

profile. The only drawback to this measurement strategy is the frequency that the DMFC enters a

rest period. Typically the time period between rest cycles is 10-20 minutes, therefore the

methanol concentration must be tracked between rests.

Methanol is consumed by the fuel cell stack through electrical power production and

methanol crossover. The methanol that is consumed for electrical power production can be

calculated based on the DMFC anode half reaction while the methanol that is consumed through

methanol crossover was measured at various methanol concentrations, temperatures and current

densities in order to develop an empirical model to predict methanol concentration. In addition to

the methanol that is consumed, a model was also developed to predict the amount of methanol

that is injected into the system via the methanol injection pump. One last model was created to

monitor the amount of solution within the system. By determining the net methanol intake for the

DMFC system and the level of anode solution, the methanol concentration can be tracked with

reasonable accuracy. The combination of the transient OCV decay measurement technique and

the methanol consumption model enabled the 20 W DMFC brassboard to successfully operate

without a methanol sensor for over 18 hours with methanol concentration excursions less than

0.15 M. Integration of this sensor-less methanol sensing technique will significantly reduce

system cost, weight and complexity accelerating the movement of DMFC technology.

Page 68: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

68

APPENDIX A

DIFFUSION MODEL DEVELOPMENT

Figure A-1. Ideal methanol concentration distribution at various load conditions.

The MEA is typically composed of five distinct layers. The anode electrode is

characterized with a diffusion and catalyst layer. At the diffusion layer, the methanol-water

solution diffuses in the direction of the membrane, while the CO2 gas generated by the anode

reaction exits in the opposite direction. The catalyst layer at the anode is where the majority of

the methanol is consumed [20]. However, some methanol continues to migrate across the

membrane to the cathode catalyst layer. In an oxygen and catalyst rich environment, the

methanol is quickly consumed at the cathode catalyst layer [39]. Due to the assumption of

complete methanol oxidation at the cathode catalyst layer, the effects of the cathode diffusion

layer were neglected. In addition, due to the limitations of the simplified 1-D model, the effects

due to stoichiometric ratios between methanol consumption rates and methanol feed rate are

neglected.

The following assumptions were used to create the unsteady diffusion model:

1. Uniform material and reactant properties.

2. Negligible diffusion contact resistance.

z-axis

Met

han

ol c

on

cen

trat

ion

Page 69: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

69

3. Consumption of methanol at the anode catalyst layer based purely on

electrochemical conversion.

4. Complete consumption of methanol at the cathode catalyst layer regardless of

concentration.

5. Concentration at the beginning (z = 0) of the anode diffusion layer is equal to the

feed methanol concentration.

6. Methanol concentration at the interface between the anode diffusion and catalyst

layers (z = zAD) is equal.

7. Diffusion of methanol at the interface between the anode diffusion and anode

catalyst layers (z = zAD) is equal.

8. Methanol concentration at the interface between the anode catalyst and membrane

layers (z = zAC) is equal.

9. Diffusion of methanol at the interface between the anode catalyst and membrane

layers (z = zAC) is equal.

10. Methanol concentration at the interface between the membrane and cathode catalyst

layers (z = zM) is equal.

11. Diffusion of methanol at the interface between the membrane and cathode catalyst

layers (z = zM) is equal.

12. Methanol concentration at the end of the cathode catalyst layer (z = zCC) is zero.

13. Negligible effects from cathode diffusion layer.

The foundation of any diffusion model starts with Fick’s First Law of Diffusion. The

concentration flux (J) is equal to the diffusion coefficient (D) times the concentration gradient,

.

A-1

In addition to Fick’s First Law of Diffusion, Fick’s Second Law of Diffusion provides a

means for solving unsteady diffusion problems.

A-2

At the anode and cathode catalyst layers, methanol is consumed through an oxidation

reaction. At the anode catalyst layer, the methanol is electrochemically consumed based on the

amount of electrical current that is generated by the fuel cell. At the cathode catalyst layer, the

crossover methanol from the anode and the oxygen from the cathode are readily consumed in the

catalyst enriched environment.

Page 70: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

70

A-3

The reaction term “r”, which accounts for the methanol that is consumed

electrochemically, was added to Fick’s Second Law resulting in a second order, non-

homogeneous, linear, partial differential equation. In order to simultaneously solve for the

methanol concentration distribution across all four layers, the method of separation of variables

and the orthogonal expansion technique were used. For each layer, the differential equations

were divided into a homogeneous and non-homogeneous set of equations as shown in Equation

A-4. Equations A-7 through A-53 summarize the mass balance, boundary and initial condition

equations.

A-4

The model that is presented is an attempt to simulate the methanol concentration

distribution for varying fuel cell load levels. The simplified, ideal, 1-D model is used primarily

for data verification and phenomena understanding. The model assumes that the fuel cell has

been instantaneously loaded from an initial condition with a current density io to a final load

condition with a current density of if.

Initial Conditions

Initial Condition 1

The methanol concentration in all of the MEA layers at t=0 is equal to the steady state

methanol concentration at the initial current density (io), solution feed concentration and

temperature. The steady state equation is only a function of space, methanol feed concentration,

temperature and current density. Therefore the homogenous equation inherits the entire portion

of the initial condition.

Page 71: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

71

A-5

A-6

Anode Diffusion Layer (0 ≤ z ≤ zAD)

The anode diffusion layer is where the feed methanol solution enters. No reaction is

present, therefore at steady state, a linear concentration gradient occurs across the diffusion layer.

For the anode diffusion layer, the formation of the homogeneous and non-homogeneous

equations are summarized in equations A-7 through A-10.

Mass Balance Equations

A-7

A-8

A-9

A-10

The equations that apply to the boundary condition at the interface at the beginning of the

anode diffusion layer are shown in equations A-11 through A-14.

Boundary Conditions

Boundary condition 1

The methanol concentration at the anode diffusion layer inlet is equal to the feed methanol

concentration. The boundary condition for the homogeneous set of equations is equal to zero.

A-11

Page 72: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

72

A-12

A-13

A-14

Anode Catalyst Layer (zAD ≤ z ≤ zAC)

The anode catalyst layer is where the majority of the methanol in the fuel cell is consumed.

The amount of methanol that is consumed at the anode catalyst layer (rAC) is based entirely on

the amount of electrical current that the fuel cell is producing. The mass balance equation for this

region is defined in equations A-15 through A-18. A non-linear distribution can be expected at

the anode catalyst layer at steady state due to the methanol that is consumed at the anode catalyst

layer.

Mass Balance Equations

A-15

A-16

A-17

A-18

Steady State Boundary Equations

The boundary condition for the interface between the anode diffusion layer and the anode

catalyst layer are defined in Equations A-19 through A-26.

Page 73: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

73

Boundary condition 2

The methanol concentration at the anode diffusion layer exit is equal to the methanol

concentration at the inlet of the anode catalyst layer.

A-19

A-20

A-21

A-22

Boundary condition 3

The methanol concentration flux at the exit of the anode diffusion layer is equal to the

methanol flux at the inlet anode catalyst layer.

|

|

A-23

|

|

|

|

A-24

|

|

A-25

|

|

A-26

Membrane Layer (zAC ≤ z ≤ zM)

The membrane layer is what separates the anode and cathode reactions. However,

crossover methanol migrates across this layer and subsequent layers. It is assumed that the

Page 74: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

74

consumption of methanol at the membrane is zero, therefore a linear methanol concentration

gradient can be expected after steady state is reached.

Mass Balance Equations

A-27

A-28

A-29

A-30

Boundary Equations

The boundary condition for the interface between the anode catalyst layer and the

membrane layer are defined in Equations A-31 through A-38.

Boundary condition 4

The methanol concentration at the exit of the anode catalyst layer is equal to the methanol

concentration at the inlet of the membrane.

A-31

A-32

A-33

A-34

Page 75: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

75

Boundary condition 5

The methanol concentration flux at the exit of the anode catalyst layer is equal to the

methanol flux at the inlet of the membrane.

|

|

A-35

|

|

|

|

A-36

|

|

A-37

|

|

A-38

Cathode Catalyst Layer (zM ≤ z ≤ zCC)

The cathode catalyst layer is not the final layer found in a typical MEA, however it has

been assumed that due to the presence of the cathode catalyst, the methanol is entirely oxidized

with the oxygen from the cathode. The consumption of the methanol in the cathode catalyst layer

is defined by rcc, where the rate of methanol consumption is heavily dependent upon the local

concentration at the cathode catalyst layer. The resulting steady state distribution of methanol

concentration is non-linear.

Mass Balance Equations

A-39

Page 76: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

76

A-40

A-41

A-42

Boundary Equations

The boundary conditions six and seven are defined at the interface between the membrane

and cathode catalyst layers, while boundary equation eight defines the condition at the exit of the

cathode catalyst layer.

Boundary condition 6

The concentration at the exit of the membrane layer is equal to the inlet at the cathode

catalyst layer.

A-43

A-44

A-45

A-46

Boundary condition 7

The concentration flux at the exit of the membrane is equal to the inlet at the cathode

catalyst layer.

|

|

A-47

Page 77: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

77

|

|

|

|

A-48

|

|

A-49

|

|

A-50

Boundary condition 8

The concentration of methanol at the exit of the cathode catalyst layer is equal to zero.

A-51

A-52

A-53

Homogenous Equations

The boundary conditions and the differential equations for the steady state and

homogeneous portions have been defined. The homogenous equations were solved using the

orthogonal expansion technique. The general solution for each layer is defined in Equations A-54

through A-57 where βn is the eigen-function and the coefficients A and B for each equation

represent constants.

(

) (

) A-54

Page 78: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

78

(

) (

) A-55

(

) (

) A-56

(

) (

) A-57

In order to solve for the eight unknowns, eight equations were defined using the boundary

equations. The eight boundary conditions are defined in general solution form in Equations A-58

through A-65.

Boundary condition 1

(

) (

) A-58

Boundary condition 2

(

) (

) (

) A-59

Boundary condition 3

√ (

)

√ (

) √ (

)

A-60

Page 79: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

79

Boundary condition 4

(

) (

) (

)

(

)

A-61

Boundary condition 5

√ (

) √ (

)

√ (

) √ (

)

A-62

Boundary condition 6

(

) (

) (

) (

) A-63

Boundary condition 7

√ (

) √ (

)

√ (

) √ (

)

A-64

Boundary condition 8

(

) (

) A-65

Boundary condition equations one through eight are defined in matrix form under Table A-

1. In order to avoid a trivial solution, the matrix from Table A-1 was modified assuming AAD is

equal to unity resulting in Table A-2.

Page 80: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

80

Table A-1. Matrix form of boundary equations for homogeneous equations. (

) (

) (

) 0 0 0 0 0

√ (

) √ (

) √ (

) 0 0 0 0 0

0 (

) (

) (

) (

) 0 0 0

0 √ (

) √ (

) √ (

) √ (

) 0 0 X = 0

0 0 0 (

) (

) (

) (

) 0

0 0 0 √ (

) √ (

) √ (

) √ (

) 0

0 0 0 0 0 (

) (

) 0

Table A-2. Modified matrix of homogenous set of boundary equations. (

) (

) (

) 0 0 0 0 0

√ (

) √ (

) √ (

) 0 0 0 0 0

0 (

) (

) (

) (

) 0 0 0

0 √ (

) √ (

) √ (

) √ (

) 0 0 X = 0

0 0 0 (

) (

) (

) (

) 0

0 0 0 √ (

) √ (

) √ (

) √ (

) 0

0 0 0 0 0 (

) (

) 0

The resulting system of equations is shown in Table A-3. The unique solution for each

system of equations was solved for the first thirty eigenvalues.

Table A-3. System of equations for homogeneous set of boundary conditions. (

) (

) 0 0 0 0 (

)

√ (

) √ (

) 0 0 0 0 √ (

)

(

) (

) (

) (

) 0 0

X

=

0

√ (

) √ (

) √ (

) √ (

) 0 0 0

0 0 (

) (

) (

) (

) 0

0 0 √ (

) √ (

) √ (

) √ (

) 0

Steady State Equations

Previously, the homogeneous equations were solved using the orthogonal expansion

technique. The other half of the overall solution is characterized with non-homogeneous

boundary conditions. These steady state equations were solved using the differential

relationships defined by Equations A-7 through A-53. The equations labeled A-66 through A-69

Page 81: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

81

are characterized with eight unknowns. Therefore eight equations were defined using the

boundary conditions.

A-66

A-67

A-68

A-69

The boundary equations for the steady state, non-homogeneous conditions are listed as

Equations A-70 through A-84.

Boundary condition 1

A-70

Boundary condition 2

A-71

A-72

Boundary condition 3

A-73

A-74

Boundary condition 4

A-75

Page 82: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

82

A-76

Boundary condition 5

A-77

A-78

Boundary condition 6

A-79

A-80

Boundary condition 7

A-81

A-82

Boundary condition 8

A-83

A-84

In order to facilitate calculations using MATLAB software, the boundary equations were

formatted into a matrix as shown in Table A-4.

Page 83: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

83

Table A-4. Boundary conditions for steady state, non-homogeneous boundary conditions.

0 1 0 0 0 0 0 0

1 -1

0 0 0 0

0 0 0 0 0 0

0 0 1 -1

0 0 x

=

0 0 0 0 0 0

0 0 0 0 1 √

0 0 0 0 0 √

0 0 0 0 0 0 √

Page 84: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

84

APPENDIX B

DIFFUSION MODEL MATLAB CODE

clear

T=50; %Cell Temperature (°C) C_FEED = 0.4; %Feed methanol concentration (Moles/Liter) i_o = 120; %Initial Current Density (mA/cm²) i_f = 0; %Final Current Density (mA/cm²)

T=T+273.15; %Convert from ºC to K. C_FEED=C_FEED/1000; %Convert from Moles/Liter to Moles/cm³

Beta_n_precision= 10; %Number of decimal places to apply to eigenvalues (14

is the max). n_max = 30; %Max number of n iterations for the summation of orthogonal

functions. t_max = 75; %The number of seconds to evaluate the function for. (s) t_d = 6; %The number of divisions for the time space specified by t_max. z_d =200; %The number of z divisions for the thickness of the MEA.

D_AD = 1.1175e-005; %Anode Diffusion Layer Diffusion Coefficient (cm²/s) t_AD = 0.015; %Anode Diffusion Layer Thickness (cm)

D_AC = 2.8*10^-5*exp(2436*(1/353-1/T)); %Anode Catalyst Layer Diffusion

Coefficient (cm²/s) t_AC = 0.0023; %Anode Catalyst Layer Thickness (cm) r_AC_o = i_o/1000/96485/6/t_AC; %Anode Catalyst Layer Reaction Coefficient

before load change. [Moles of methanol/(cm³*s), where i is the current

density in mA/cm²] r_AC_f = i_f/1000/96485/6/t_AC; %Anode Catalyst Layer Reaction Coefficient

after load change. [Moles of methanol/(cm³*s), where i is the current density

in mA/cm²]

D_M = (4.9*10^-6*exp(2436*(1/333-1/T))); %Membrane Layer Diffusion

Coefficient (cm²/s) "Determination of methanol diffusion and electro-

osmotic drag coefficients in proton-exchange-membranes for DMFC" t_M = 0.018; %Membrane Layer Thickness (cm) [Alex's Model]

D_CC = 2.8*10^-5*exp(2436*(1/353-1/T)); %Cathode Catalyst Layer Diffusion

Coefficient (cm²/s) t_CC = 0.0023; %Cathode Catalyst Layer Thickness (cm) [Alex's Model] r_CC = 70/1000/96485/6/t_CC; %Cathode Catalyst Layer Reaction Coefficient

z_AD = t_AD; %Distance in the z-direction to end of the anode diffusion layer

starting from the anode side. (cm) z_AC = t_AD+t_AC; %Distance in the z-direction to to end of the anode

catalyst layer starting from the the anode side. (cm) z_M = t_AD+t_AC+t_M; %Distance in the z-direction to to end of the membrane

layer starting from the the anode side. (cm) z_CC = t_AD+t_AC+t_M+t_CC; %Distance in the z-direction to to end of the

cathode catalyst layer starting from the the anode side. (cm)

Page 85: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

85

A_S = [0, 1, 0, 0, 0, 0, 0, 0; %Boundary Condition at Beginning of Anode

Diffusion Layer (Concentration) z_AD, 1, -z_AD, -1, 0, 0, 0, 0; %Boundary Condition between Anode

Diffusion and Catalyst Layers (Concentration) D_AD, 0, -D_AC, 0, 0, 0, 0, 0; %Boundary Condition between Anode

Diffusion and Catalyst Layers (Diffusion of Methanol) 0, 0, z_AC, 1, -z_AC, -1, 0, 0; %Boundary Condition between Anode

Catalyst Layer and Membrane Layer (Concentration) 0, 0, D_AC, 0, -D_M, 0, 0, 0; %Boundary Condition between Anode Catalyst

Layer and Membrane Layer (Diffusion of Methanol) 0, 0, 0, 0, z_M, 1, -exp(sqrt(r_CC/D_CC)*z_M), -exp(-

sqrt(r_CC/D_CC)*z_M); %Boundary Condition between Membrane Layer and Cathode

Catalyst Layer (Concentration) 0, 0, 0, 0, D_M, 0, -D_CC*sqrt(r_CC/D_CC)*exp(sqrt(r_CC/D_CC)*z_M),

D_CC*sqrt(r_CC/D_CC)*exp(-sqrt(r_CC/D_CC)*z_M); %Boundary Condition between

Membrane Layer and Cathode Catalyst Layer (Diffusion of Methanol) 0, 0, 0, 0, 0, 0, exp(sqrt(r_CC/D_CC)*z_CC), exp(-sqrt(r_CC/D_CC)*z_CC)];

%Boundary Condition at exit of cathode catalyst layer. (Concentration)

%Non_Homogeneous Conditions for Steady State ODEs at the initial condition. B_o_S = [C_FEED; %Boundary Condition at Beginning of Anode Diffusion Layer

(Concentration) r_AC_o*z_AD^2/2/D_AC; %Boundary Condition between Anode Diffusion and

Catalyst Layers (Concentration) r_AC_o*z_AD; %Boundary Condition between Anode Diffusion and Catalyst

Layers (Diffusion of Methanol) -r_AC_o*z_AC^2/2/D_AC; %Boundary Condition between Anode Catalyst Layer

and Membrane Layer (Concentration) -r_AC_o*z_AC; %Boundary Condition between Anode Catalyst Layer and

Membrane Layer (Diffusion of Methanol) 0; %Boundary Condition between Membrane Layer and Cathode Catalyst Layer

(Concentration) 0; %Boundary Condition between Membrane Layer and Cathode Catalyst Layer

(Diffusion of Methanol) 0]; %Boundary Condition at exit of cathode catalyst layer.

(Concentration)

%Non_Homogeneous Conditions for Steady State ODEs at the final condition. B_f_S = [C_FEED; %Boundary Condition at Beginning of Anode Diffusion Layer

(Concentration) r_AC_f*z_AD^2/2/D_AC; %Boundary Condition between Anode Diffusion and

Catalyst Layers (Concentration) r_AC_f*z_AD; %Boundary Condition between Anode Diffusion and Catalyst

Layers (Diffusion of Methanol) -r_AC_f*z_AC^2/2/D_AC; %Boundary Condition between Anode Catalyst Layer

and Membrane Layer (Concentration) -r_AC_f*z_AC; %Boundary Condition between Anode Catalyst Layer and

Membrane Layer (Diffusion of Methanol) 0; %Boundary Condition between Membrane Layer and Cathode Catalyst Layer

(Concentration) 0; %Boundary Condition between Membrane Layer and Cathode Catalyst Layer

(Diffusion of Methanol) 0]; %Boundary Condition at exit of cathode catalyst layer.

(Concentration)

Page 86: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

86

k_o = A_S\B_o_S; %Steady State Coefficients for concentration gradient before

load change. k_f = A_S\B_f_S; %Steady State Coefficients for concentration gradient after

load change.

syms z

C_AD_S = sym((k_f(1)*z+k_f(2))); %Steady state concentration gradient at

anode diffusion layer after load change. C_AC_S = sym((r_AC_f/2/D_AC*z^2+k_f(3)*z+k_f(4))); %Steady state

concentration gradient at anode catalyst layer after load change. C_M_S = sym((k_f(5)*z+k_f(6))); %Steady state concentration gradient at

membrane layer after load change. C_CC_S = sym(k_f(7)*exp(sqrt(r_CC/D_CC)*z)+k_f(8)*exp(-sqrt(r_CC/D_CC)*z));

F_AD = sym((k_o(1)*z+k_o(2)))-C_AD_S; %Initial condition for Anode Diffusion

Layer. F_AC = sym((r_AC_o/2/D_AC*z^2+k_o(3)*z+k_o(4)))-C_AC_S; %Initial condition

for Anode Catalyst Layer. F_M = sym((k_o(5)*z+k_o(6)))-C_M_S; %Initial condition for Membrane Layer. F_CC = sym(k_o(7)*exp(sqrt(r_CC/D_CC)*z)+k_o(8)*exp(-sqrt(r_CC/D_CC)*z))-

C_CC_S; %Inital condition for Cathode Catalyst Layer.

syms Beta_n A_AD A_AC B_AC A_M B_M A_CC B_CC

A_H = [sin(Beta_n*z_AD/D_AC^0.5), cos(Beta_n*z_AD/D_AC^0.5), 0, 0, 0, 0; Beta_n*D_AC^0.5*cos(Beta_n*z_AD/D_AC^0.5), -

Beta_n*D_AC^0.5*sin(Beta_n*z_AD/D_AC^0.5), 0, 0, 0, 0; sin(Beta_n*z_AC/D_AC^0.5), cos(Beta_n*z_AC/D_AC^0.5), -

sin(Beta_n*z_AC/D_M^0.5), -cos(Beta_n*z_AC/D_M^0.5), 0, 0; Beta_n*D_AC^0.5*cos(Beta_n*z_AC/D_AC^0.5), -

Beta_n*D_AC^0.5*sin(Beta_n*z_AC/D_AC^0.5), -

Beta_n*D_M^0.5*cos(Beta_n*z_AC/D_M^0.5),

Beta_n*D_M^0.5*sin(Beta_n*z_AC/D_M^0.5), 0, 0; 0, 0, sin(Beta_n*z_M/D_M^0.5), cos(Beta_n*z_M/D_M^0.5), -

sin(Beta_n*z_M/D_CC^0.5), -cos(Beta_n*z_M/D_CC^0.5); 0, 0, Beta_n*D_M^0.5*cos(Beta_n*z_M/D_M^0.5), -

Beta_n*D_M^0.5*sin(Beta_n*z_M/D_M^0.5), -

Beta_n*D_CC^0.5*cos(Beta_n*z_M/D_CC^0.5),

Beta_n*D_CC^0.5*sin(Beta_n*z_M/D_CC^0.5)];

B_H = [sin(Beta_n*z_AD/D_AD^0.5); Beta_n*D_AD^0.5*cos(Beta_n*z_AD/D_AD^0.5); 0; 0; 0; 0];

Psi_AD_n = sym(A_AD*sin(Beta_n*z/D_AD^0.5)); Psi_AC_n = sym(A_AC*sin(Beta_n*z/D_AC^0.5)+B_AC*cos(Beta_n*z/D_AC^0.5)); Psi_M_n = sym(A_M*sin(Beta_n*z/D_M^0.5)+B_M*cos(Beta_n*z/D_M^.5)); Psi_CC_n = sym(A_CC*sin(Beta_n*z/D_CC^0.5)+B_CC*cos(Beta_n*z/D_CC^0.5));

k_H = simple(A_H\B_H);

Page 87: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

87

syms Beta_n_t

%Homogenous Equations for Unsteady PDEs in matrix form to solve for

Eigenvalues. A_H_E = [sin(Beta_n_t*z_AD/D_AD^0.5), -sin(Beta_n_t*z_AD/D_AC^0.5), -

cos(Beta_n_t*z_AD/D_AC^.5), 0, 0, 0, 0; Beta_n_t*D_AD^0.5*cos(Beta_n_t*z_AD/D_AD^0.5), -

Beta_n_t*D_AC^0.5*cos(Beta_n_t*z_AD/D_AC^0.5),

Beta_n_t*D_AC^0.5*sin(Beta_n_t*z_AD/D_AC^0.5), 0, 0, 0, 0; 0, sin(Beta_n_t*z_AC/D_AC^0.5), cos(Beta_n_t*z_AC/D_AC^0.5), -

sin(Beta_n_t*z_AC/D_M^0.5), -cos(Beta_n_t*z_AC/D_M^0.5), 0, 0; 0, Beta_n_t*D_AC^0.5*cos(Beta_n_t*z_AC/D_AC^0.5), -

Beta_n_t*D_AC^0.5*sin(Beta_n_t*z_AC/D_AC^0.5), -

Beta_n_t*D_M^0.5*cos(Beta_n_t*z_AC/D_M^0.5),

Beta_n_t*D_M^0.5*sin(Beta_n_t*z_AC/D_M^0.5), 0, 0; 0, 0, 0, sin(Beta_n_t*z_M/D_M^0.5), cos(Beta_n_t*z_M/D_M^0.5), -

sin(Beta_n_t*z_M/D_CC^0.5), -cos(Beta_n_t*z_M/D_CC^0.5); 0, 0, 0, Beta_n_t*D_M^0.5*cos(Beta_n_t*z_M/D_M^0.5), -

Beta_n_t*D_M^0.5*sin(Beta_n_t*z_M/D_M^0.5), -

Beta_n_t*D_CC^0.5*cos(Beta_n_t*z_M/D_CC^0.5),

Beta_n_t*D_CC^0.5*sin(Beta_n_t*z_M/D_CC^0.5); 0, 0, 0, 0, 0, sin(Beta_n_t*z_CC/D_CC^0.5), cos(Beta_n_t*z_CC/D_CC^0.5)];

P = simple(det(A_H_E));

%Solve for Eigenvalues Beta_n_t = 1/10^Beta_n_precision;

for n = 1:1:n_max n_max-n+1 delta = 1;

while delta >= 1/10^Beta_n_precision Test_b = subs(P) >= 0; Beta_n_t = Beta_n_t + delta; Test_a = subs(P) >= 0;

if Test_a ~= Test_b Beta_n_t = Beta_n_t-delta; delta = delta/10; end end Beta_n_array(n) = Beta_n_t-delta/2*10; Beta_n_t = Beta_n_t + delta*10; end

t_array = linspace(0,t_max,t_d); z_array = linspace(0,z_CC,z_d);

for n=1:1:n_max Beta_n = Beta_n_array(n);

n_max-n

A_AD = 1;

Page 88: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

88

A_AC = subs(k_H(1)); B_AC = subs(k_H(2)); A_M = subs(k_H(3)); B_M = subs(k_H(4)); A_CC = subs(k_H(5)); B_CC = subs(k_H(6));

N_n =

subs(int(Psi_AD_n^2,0,z_AD))+subs(int(Psi_AC_n^2,z_AD,z_AC))+subs(int(Psi_M_n

^2,z_AC,z_M))+subs(int(Psi_CC_n^2,z_M,z_CC)); tail =

subs(int(Psi_AD_n*F_AD,0,z_AD))+subs(int(Psi_AC_n*F_AC,z_AD,z_AC))+subs(int(P

si_M_n*F_M,z_AC,z_M))+subs(int(Psi_CC_n*F_CC,z_M,z_CC));

for j=1:1:t_d t=t_array(j) nose = subs(1/N_n*exp(-Beta_n^2*t)); for i=1:1:z_d z=z_array(i);

if j == 1

if z <= z_AD C_S(i) = subs(C_AD_S);

elseif z <= z_AC & z_AD < z C_S(i) = subs(C_AC_S);

elseif z <= z_M & z_AC < z C_S(i) = subs(C_M_S);

elseif z <= z_CC & z_M < z C_S(i) = subs(C_CC_S); end end

if z <= z_AD C_H_n(i,j,n) = subs(nose*Psi_AD_n*tail);

elseif z <= z_AC & z_AD < z C_H_n(i,j,n) = subs(nose*Psi_AC_n*tail);

elseif z <= z_M & z_AC < z C_H_n(i,j,n) = subs(nose*Psi_M_n*tail);

elseif z <= z_CC & z_M < z C_H_n(i,j,n) = subs(nose*Psi_CC_n*tail); end

if n == n_max C_H(i,j) = sum(C_H_n(i,j,:)); C(i,j) = C_H(i,j) + C_S(i); end end

Page 89: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

89

end end

plot(z_array,C*1000,'linewidth',2) axis([0 z_CC 0 1.8]) set(gca,'FontSize',16) xlabel('MEA Position (cm)','FontSize',16,'FontWeight', 'Bold') ylabel('Methanol Concentration (M)','FontSize',16,'FontWeight', 'Bold') Chart_Title = sprintf('MEA Methanol Concentration Distribution from %d mA/cm²

to %d mA/cm² at %0.2f M',i_o, i_f,C_FEED*1000); title(Chart_Title,'FontSize',24,'FontWeight','Bold') leg1=sprintf('%0.1f s (%d mA/cm²)',0*t_max/(t_d-1),i_o); leg2=sprintf('%0.1f s (%d mA/cm²)',1*t_max/(t_d-1),i_f); leg3=sprintf('%0.1f s (%d mA/cm²)',2*t_max/(t_d-1),i_f); leg4=sprintf('%0.1f s (%d mA/cm²)',3*t_max/(t_d-1),i_f); leg5=sprintf('%0.1f s (%d mA/cm²)',4*t_max/(t_d-1),i_f); leg6=sprintf('%0.1f s (%d mA/cm²)',5*t_max/(t_d-1),i_f); legh=legend(leg1,leg2,leg3,leg4,leg5,leg6); set(legh, 'Position', [.733,.663,.145,.22]) line([z_AC z_AC],[0 1.8],'linestyle','--','color','black') line([z_AD z_AD],[0 1.8],'linestyle','--','color','black') line([z_M z_M],[0 1.8],'linestyle','--','color','black')

format short g Total_CH3OH=(C_FEED-C(80,1))/z_array(80)*D_AD*96485*6*1000 XO_CH3OH=(C(188,1)-C(200,1))/(z_array(200)-z_array(188))*D_CC*96485*6*1000 CH3OH_Ratio=XO_CH3OH/((C(188,2)-C(200,2))/(z_array(200)-

z_array(188))*D_CC*96485*6*1000) format long poo=[C_FEED; i_o; D_AD; D_AC; r_AC_o; D_M; D_CC; r_CC; Total_CH3OH; XO_CH3OH;

CH3OH_Ratio] format short

Page 90: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

90

LIST OF REFERENCES

[1] Moore's Law: Made real by Intel® innovation: 2009; Available at:

http://www.intel.com/technology/mooreslaw/. Accessed 07/21, 2009.

[2] M. Broussely, G. Archdale, J.Power Sources, 136 (2004) 386-394.

[3] http://www.faa.gov/about/office_org/headquarters_offices/ash/ash_programs/hazmat/aircarri

er_info/media/Battery_incident_chart.pdf

[4] J.G. Liu, T.S. Zhao, R. Chen, C.W. Wong, Electrochemistry Communications, 7 (2005) 288-

294.

[5] A. Heinzel, V.M. Barragán, J.Power Sources, 84 (1999) 70-74.

[6] M. Walker, K.-M. Baumgärtner, M. Kaiser, J. Kerres, A. Ullrich, E. Räuchle, J Appl Polym

Sci, 74 (1999) 67-73.

[7] ATP Project Brief - 00-00-7744: Available at:

http://jazz.nist.gov/atpcf/prjbriefs/prjbrief.cfm?ProjectNumber=00-00-7744. Accessed

07/23, 2009.

[8] J.A. McAllister, A.E. Farrell, Energy, 32 (2007) 1177-1184.

[9] BAJ Website | Total battery production statistics: 2011; Available at:

http://www.baj.or.jp/e/statistics/01.html. Accessed 08/11, 2011.

[10] R.A. Powers, Proceedings of the IEEE, 83 (1995) 687-693.

[11] B. Scrosati, Electrochim.Acta, 45 (2000) 2461-2466.

[12] J. Vetter, P. Novák, M.R. Wagner, C. Veit, K.-. Möller, J.O. Besenhard, M. Winter, M.

Wohlfahrt-Mehrens, C. Vogler, A. Hammouche, J.Power Sources, 147 (2005) 269-281.

[13] J. Shim, R. Kostecki, T. Richardson, X. Song, K.A. Striebel, J.Power Sources, 112 (2002)

222-230.

[14] J. McLaughlin, (2008).

[15] Ackerman D. New Apple MacBooks demystified. 06/08/09; Available at:

http://news.cnet.com/8301-17938_105-10260001-1.html?tag=rb_content;contentMain.

Accessed 08/19, 2011.

[16] Stanford University. New Nanowire Battery Holds 10 Times The Charge Of Existing Ones.

2007; Available at: http://www.sciencedaily.com/releases/2007/12/071219103105.htm.

Accessed 08/19, 2011.

Page 91: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

91

[17] K. Scott, W.M. Taama, S. Kramer, P. Argyropoulos, K. Sundmacher, Electrochim.Acta, 45

(1999) 945-957.

[18] S. Song, W. Zhou, W. Li, G. Sun, Q. Xin, S. Kontou, P. Tsiakaras, Ionics, 10 (2004) 458-

462.

[19] C.Y. Du, T.S. Zhao, W.W. Yang, Electrochim.Acta, 52 (2007) 5266-5271.

[20] B. Gurau, E.S. Smotkin, J.Power Sources, 112 (2002) 339-352.

[21] S. Doerner, T. Schultz, T. Schneider, K. Sundmacher, P. Hauptmann, Sensors, 2004.

Proceedings of IEEE, (2004) 639-641 vol.2.

[22] Dielectric Constants of Materials: 2008; Available at:

http://clippercontrols.com/info/dielectric_constants.html#D. Accessed 07/24, 2009.

[23] H. Zhao, J. Shen, J. Zhang, H. Wang, D.P. Wilkinson, C.E. Gu, J.Power Sources, 159

(2006) 626-636.

[24] J.P. Longtin, C.H. Fan, Microscale Thermophysical Engineering, 2 (1998) 261-272.

[25] CRC handbook of chemistry and physics, CRC Press; 1978, pp. 8-70.

[26] A. Rabinovich, E. Diatzikis, J. Mullen, D. Tulimieri, US patent 6,815,682 (2003).

[27] G.C. Benson, P.J. D'Arcy, Journal of Chemical & Engineering Data, 27 (1982) 439-442.

[28] M. Baldauf, W. Preidel, US Patent 6,536,262

[29] F.M. White, Fluid mechanics, 6th ed., McGraw-Hill Higher Education, Boston, MA; 2008,

pp. 864.

[30] D. Sparks, R. Smith, V. Cruz, N. Tran, A. Chimbayo, D. Riley, N. Najafi, Sensors and

Actuators A: Physical, 149 (2009) 38-41.

[31] S.R. Narayanan, T.I. Valdez, W. Chun, Electrochem.Solid-State Lett., 3 (2000) 117-120.

[32] C.L. Chang, C.Y. Chen, C.C. Sung, D.H. Liou, J.Power Sources, 164 (2007) 606-613.

[33] J. Cristiani, N. Sifer, E. Bostic, P. Fomin, D. Reckar, Annual Meeting of the American

Institute of Chemical Engineers, (2005).

[34] E. Antolini, R.R. Passos, E.A. Ticianelli, J.Appl.Electrochem., 32 (2002) 383-388.

[35] P. Argyropoulos, K. Scott, W.M. Taama, Electrochim.Acta, 45 (2000) 1983-1998.

[36] Q. Ye, T.S. Zhao, H. Yang, J. Prabhuram, Electrochem.Solid-State Lett., 8 (2005) A52-A54.

Page 92: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

92

[37] A. Casalegno, P. Grassini, R. Marchesi, Appl.Therm.Eng., 27 (2007) 748-754.

[38] C. Eickes, P. Piela, J. Davey, P. Zelenay, J.Electrochem.Soc., 153 (2006) A171-A178.

[39] F. Liu, C. Wang, J.Electrochem.Soc., 154 (2007) B514-B522.

Page 93: BY WILLIAM JASON HARRINGTON - University of Floridaufdcimages.uflib.ufl.edu/UF/E0/04/44/18/00001/HARRINGTON_W.pdf · By William Jason Harrington August 2012 Chair: William E. Lear,

93

BIOGRAPHICAL SKETCH

William “Jason” Harrington was born in Quepos, Costa Rica where at the age of one year

he moved with his family to Ormond Beach, Florida. In the spring of 2006, Jason graduated with

his Bachelor of Science in mechanical engineering from the University of North Florida. There,

he developed an interest in clean and renewable energy while working under Dr. James Fletcher

as a laboratory assistant. In 2010, he began working as a systems engineer for direct methanol

fuel cells with the University of North Florida. Jason graduated with his Master of Science in

mechanical engineering in the summer of 2012.