using microscale flow cells to study the … · araz zarkhah, chris, timberg, karl siil, and saman...

132
USING MICROSCALE FLOW CELLS TO STUDY THE ELECTROCHEMICAL REDUCTION OF CARBON DIOXIDE BY MICHAEL RICHARD THORSON DISSERTATION Submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in Chemical Engineering in the Graduate College of the University of Illinois at Urbana-Champaign, 2012 Urbana, Illinois Doctoral Committee: Professor Paul J. A. Kenis, Chair Professor Charles F. Zukoski Assistant Professor Charles M. Schroeder Professor Andrew A. Gewirth Professor Thomas B. Rauchfuss

Upload: others

Post on 05-Aug-2020

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

USING MICROSCALE FLOW CELLS TO STUDY THE

ELECTROCHEMICAL REDUCTION OF CARBON DIOXIDE

BY

MICHAEL RICHARD THORSON

DISSERTATION

Submitted in partial fulfillment of the requirements

for the degree of Doctor of Philosophy in Chemical Engineering

in the Graduate College of the

University of Illinois at Urbana-Champaign, 2012

Urbana, Illinois

Doctoral Committee:

Professor Paul J. A. Kenis, Chair

Professor Charles F. Zukoski

Assistant Professor Charles M. Schroeder

Professor Andrew A. Gewirth

Professor Thomas B. Rauchfuss

Page 2: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

ii

Abstract

An electrochemical process to reduce CO2 has potential to store electrical energy from

renewable sources such as wind and solar in chemical form via the production of chemical fuels

and feedstocks. When combined with a “green” renewable source, this technology provides a

means of (a) reducing dependence on foreign oil via enhancing the penetration of renewable

technologies into the transportation sector and (b) reducing CO2 emissions by moving away from

fossil fuels (Chapter 1). The goal of this research is to use a microfluidic platform to study the

influence of electrolytes, reactor conditions and catalysts to address the three primary

challenges for the electrochemical reduction of CO2 to CO: (1) low reaction energetic efficiency;

(2) low reaction rate; and (3) poor reaction selectivity.

This dissertation reports an investigation of roles electrolytes, operating conditions, and

catalysts can play in the reduction of CO2. The anion and cation size, along with pH, were found

to drastically influence both the current density and selectivity of the product distribution

(Chapter 2). Specifically, small anions and large cations were found to favor the production of

CO and hinder the evolution of H2.

The microfluidic reactor was used to look at the performance of a larger cation,

EMIM BF4, an ionic liquid. With the ionic liquid-based electrolyte, an early cathode onset

potential was observed at the expense of poor anode performance. The poor anode performance

was overcome via the addition of a secondary electrolyte stream. Using the modified reactor, in

the presence of an EMIM BF4 electrolyte solution, CO production was achieved at a reactor

potential of only 1.5 V, which constitutes a maximum cathode overpotential of 0.17 V

(Chapter 3). This low onset potential constitutes a drastic reduction in the onset potential for

CO production in an arrangement that achieves excellent selectivity for CO. While the energetic

Page 3: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

iii

efficiency in the dual electrolyte setup was drastically improved, the achieved current density

was quite low because the EMIM+ cation poisoned the cathode. Because the poor anode

performance was from the BF4 anion, the membrane could be eliminated by substituting a BF4

anion in the ionic liquid electrolyte solution for either a OH or Cl anion. The substitution of the

BF4 anion for either a OH or Cl anion enabled drastic improvements in the partial current density

for CO when using either an EMIM OH or an EMIM Cl electrolyte solution (Chapter 4).

Further improvements were achieved via bolstering the conductivity with the addition of a

secondary salt, KOH.

Amine-based novel organometallic complexes have been developed for the

electrochemical reduction of CO2 to CO (Chapter 5). Specifically, several silver-based

organometallic catalysts, AgDAT, AgPc, and AgPz, have comparable or better current densities

with increased selectivity for CO as compared to Ag nanoparticle-based catalysts. Furthermore,

when comparing the performance relative to the silver loading, the amine-based organometallic

catalysts outperform the silver catalyst by more than 20 fold. Furthermore, when using copper-

based organometallic catalysts (i.e., CuDAT), the product selectivity changed. This resultant

change in selectivity and current density demonstrates that organometallic complexes have

potential to “tune” catalysts to produce a wide range of desired products.

The electrochemical CO2 conversion reactor used in the work described here (Chapters 2

through 5) is based on a microfluidic fuel cell developed earlier in the Kenis group. Chapter 6

reports on the influence of boundary layer depletion on performance in an alkaline, air-breathing

laminar flow fuel cell as a function of electrode length, electrode arrangement, and flow rates.

Higher current densities were achieved when using shorter electrodes. Furthermore, alternative

Page 4: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

iv

electrode arrangements (aspect ratios and multiple electrodes in series) enable higher fuel

utilization and power output per catalyst area.

Looking forward, these studies will shed insight in regards to optimal electrolyte

composition and catalyst development with the aim of further improving the current density and

energetic efficiency of the reaction.

Page 5: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

v

To Candi Thorson

Page 6: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

vi

Acknowledgements

I am extremely grateful for the many people who have helped to make my dissertation

possible. First, I would like to thank my advisor, Dr. Paul J.A. Kenis for providing critical

feedback and much needed advice in my dissertation. I am very fortunate to have had an advisor

who was supportive of me finding an area of research that I was passionate about. I am

appreciative of my dissertation committee, Professors Thomas Rauchfus, Andrew Gewirth,

Charles Schroeder, and Charles Zukoski for the insight they provided and time they invested to

be part of my committee.

In addition to my mentors, I have been fortunate to have excellent co-workers within the

Kenis group. First, I would like to thank Dr. Fikile Brushett and Dr. Devin Whipple, who

provided mentorship. Specifically, I would like to thank Sachit Goyal, Matt Naughton,

Dr. Adam Hollinger, Molly Jhong, and Sichao Ma who I worked most closely with. Finally, I

am grateful for the other members of the Kenis group for their scientific insight, and comradery,

particularly, Dr. Benjamin Schudel, Sudipto Guha, Dr. Sarah Perry, Dr. Amit Desai, Josh Tice,

Ritika Mohan, Dr. Ashtamurthy Pawate, Vivek Kumar, Matthew Byrne, Dr. Matthew Cole, and

Dr. Pedro Lopez-Montesinos. In addition, thanks to the hard working undergraduate students,

Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition

and whom I was given the privilege to mentor.

I am grateful for my collaborators, without whom much of this work would not have been

possible. Specifically, I am grateful for my collaborators at Dioxide Materials, Dr. Richard

Masel, Dr. Amin Salehi-Khojin, Dr. Wei Zhu, and Brian Rosen whom I worked alongside in the

testing of the ionic liquid, EMIM BF4. I am also very grateful for the assistance of Claire

Page 7: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

vii

Tornow and Dr. Andrew Gewirth from the Chemistry Department at the University of Illinois for

catalyst synthesis, testing, and analysis.

I appreciate the generous funding for this work, which was made possible by the NSF,

DOE, I2CNER, and the Harry G. Drickamer fellowship program.

I especially thank my wife, Candi Thorson for her emotional support throughout my PhD.

I appreciate my parents for supporting me and instilling in me the perceptive that there is no

substitute for hard work.

Page 8: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

viii

Table of Contents

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

1.1 Grand Challenges .................................................................................................................1

1.2 Electrochemical Reduction of CO2 ....................................................................................10

1.3 Challenges for the Electrochemical Reduction of CO2......................................................12

1.4 Economic Analysis ............................................................................................................16

1.5 Remaining Challenges .......................................................................................................17

1.6 References ..........................................................................................................................19

Chapter 2: Effect of Ions on the Electrochemical Conversion of CO2 to CO ........................25

2.1 Introduction ........................................................................................................................25

2.2 Experimental ......................................................................................................................27

2.3 Results and Discussion ......................................................................................................30

2.4 Conclusions ........................................................................................................................38

2.5 References ..........................................................................................................................40

Chapter 3: Amine Based Electrolytes for the Electrochemical Reduction of CO2 ................44

3.1 Introduction ........................................................................................................................44

3.2 Experimental ......................................................................................................................46

3.3 Results and Discussion ......................................................................................................49

3.4 Conclusions ........................................................................................................................56

3.5 References ..........................................................................................................................57

Chapter 4: Increased Current Densities for Amine-Based EMIM+ Electrolytes ..................60

4.1 Introduction ........................................................................................................................60

4.2 Experimental ......................................................................................................................61

4.3 Results and Discussion ......................................................................................................61

4.4 Conclusions ........................................................................................................................65

4.5 References ..........................................................................................................................66

Chapter 5: Organometallic Catalysts for the Electrochemical Reduction of CO2 ................68

5.1 Introduction ........................................................................................................................68

5.2 Experimental ......................................................................................................................70

Page 9: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

ix

5.3 Results and Discussion ......................................................................................................74

5.4 Conclusions ........................................................................................................................89

5.5 References ..........................................................................................................................90

Chapter 6: Design Rules for Electrode Arrangement in an Air-breathing Alkaline Direct

Methanol LFFC ............................................................................................................................93

6.1 Introduction ........................................................................................................................93

6.2 Experimental ......................................................................................................................95

6.3 Results and Discussion ......................................................................................................99

6.4 Conclusions ......................................................................................................................107

6.5 References ........................................................................................................................108

Chapter 7: Summary of Accomplishments and Future Directions .......................................112

7.1 Summary of Accomplishments ........................................................................................112

7.2 Future Directions .............................................................................................................115

7.3 References ........................................................................................................................117

Appendix A: Supplementary Section to Chapter 6 ................................................................119

Page 10: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

1

Chapter 1

Introduction

1.1 Grand Challenges

Carbon dioxide is nature’s chemical of choice for energy storage. Most plants store solar

energy by converting CO2 from the air into glucose, which is temporarily stored in leaves, roots,

and stems [1]. Later, the glucose is turned into starch and plant matter [1]. Here we propose to

use CO2 as a storage medium for energy from renewable sources, such as the sun, to enable

“artificial photosynthesis”. Specifically, we propose that coupling renewable energy with a

process for the electrochemical conversion of CO2 can enable penetration of non-fossil fuel

sources into the transportation sector and thereby both decrease CO2 emissions and meet the

growing demand for liquid fuels within the transportation sector. The major technical challenge

to improving the efficiency of the process for the electrochemical conversion of CO2 is achieving

high current densities and high energetic efficiencies while being selective for the desired

product.

1.1.1 Increasing Demand for Oil

Figure 1.1 shows several projections predicting an increase in demand for petroleum for

the foreseeable future. Over the last 25 years, the demand for oil has steadily increased.

Furthermore, as seen in Figure 1.1, estimates for the growth rate in demand of petroleum vary

between 1.0 and 1.9 % per year, resulting in the predicted demand for petroleum to increase 35

to 76 % over the next 30 years [2]. Much of the historical and expected demand for petroleum

stems from a growing worldwide population and an increase in the percentage of the world

population living in petroleum reliant, developed communities [2]. The world population is

Page 11: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

2

expected to grow from roughly 7 billion people to roughly 9 billion people between 2012 and

2030 [3]. However, most of the demand for petroleum comes from people living in developed

countries. The Organization for Economic Co-operation and Development (OECD), which

represents part of the developed world, uses half of the world’s energy to produce half of the

world’s GDP [4]. The high demand for energy of the OECD (50 % of worldwide demand) is

significant because it constitutes only a small fraction of the world’s population [4]. The fraction

of the world population living within developed countries is expected to increase from 50 to

80 % by 2030 [2]. Because both the worldwide population is growing, and a larger fraction of

the world population will rely on oil for their daily lives, the number of people worldwide

dependent on oil is expected to double, effectively doubling the demand for oil by 2030.

1.1.2 Limited Supply of Oil

The increasing demand for oil becomes complicated by limitations in supply. Current

Figure 1.1: Historical and Projected world energy demand for oil between 1980 and 2030. © National Petroleum

Council 2007, reprinted with permission from “Hard Truths: Facing the Hard Truths About Energy: A

Comprehensive View to 2030 of Global Oil and Natural Gas” [2].

Page 12: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

3

petroleum capacity, in terms of oil in established wells, is expected to be mostly depleted by

2030, however, reserves and other advanced oil recovery means are expected to provide a

petroleum supply for the foreseeable future [5]. Nonetheless, because most of the oil reserves

which will be used over the next several decades have yet to be identified, all predictions of

when the world’s oil supplies will be depleted have significant uncertainty [5,6]. Furthermore,

as traditional wells become depleted, less efficient, unconventional means of oil recovery will

have to be used, which increases the cost due to the increased technology and equipment

necessary to recovery the oil. In some cases, unconventional means of oil recovery come with

increased risks (e.g., spills from deep water drilling), which must be considered in an energy

source analysis [7].

Even if petroleum is available for the foreseeable future, other factors such as geopolitical

influences, emissions of greenhouse gasses, and energy security issues can limit the extent of oil

Figure 1.2: Production rates of oil from the leading worldwide producers of petroleum. © National Petroleum

Council 2011 reprinted with permission from “Prudent Development: Realizing the Potential of North America's

Abundant Natural Gas and Oil” [4].

Page 13: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

4

availability or drive up the price of petroleum. For example, political tensions or embargos can

disturb the free market and thereby, hinder the availability of the petroleum supply. This need

for free trade may explain some of the motivation for the United States to declare war against

Iraq. If this was indeed motivated by the availability of oil, funding this war provides quite

significant “hidden costs” to acquiring oil, and consequently, the cost of war should be

considered when considering the price of petroleum.

A major complication, which must be considered when securing future oil supplies, is the

prevention of potential energy security issues. As Figure 1.2 shows, the United States and

Europe rely heavily on a few countries within the Middle East, and to a lesser extent, Russia and

northwestern Africa, for the majority of their imported oil. Furthermore, this trend is expected to

become exaggerated through 2030 [2,8]. Being heavily reliant on a few countries for oil imports

Figure 1.3: Historical and Projected worldwide oil imports for 2005 and 2030. © National Petroleum Council 2007,

reprinted with permission from “Hard Truths: Facing the Hard Truths About Energy: A Comprehensive View to

2030 of Global Oil and Natural Gas” [2].

Page 14: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

5

can rapidly affect Europe’s and United States’ economies if political situations change for the

worse between these countries. This potential energy security issue led the National Petroleum

Council to recommend that the United States diversify its imports, increase modernization, and

expand oil production [2]. However, as Figure 1.3 shows, the United States is already the

world’s 3rd

largest producer of oil and has realistic limitations regarding the vastness of its wells

[2]. The United States has large reserves of Natural Gas and Coal, which are much more

extensive than the United States’ oil reserves; however, at present, penetration of these sources

into the transportation sector is minimal.

1.1.3 Linking Supply and Demand

Figure 1.4 shows the current penetration of the most common sources of energy into

various sectors of energy demand [9]. Of note, the transportation sector relies almost entirely on

Figure 1.4: Primary U.S. energy consumption by source and sector for 2008. Reprinted (adapted) with permission

from: D. T. Whipple and P. J. A. Kenis, “Prospects of CO2 Utilization via Direct Heterogeneous Electrochemical

Reduction” Journal of Physical Chemistry Letters, 2010. 1 (24): p. 3451-3458.. Copyright (2010) American

Chemical Society [9].

Page 15: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

6

petroleum as its energy source. The reliance of the transportation sector on petroleum

compounded with the aforementioned increasingly concentrated supply of petroleum exacerbates

potential national security issues because a shortage of oil immediately limits the extent of

transportation possible. Consequently, a need exists for technology that enables penetration of

other sources of energy into the transportation sector.

A simple means to enable penetration of other energy sources into the transportation

sector relies on improving automobile efficiency through the development of hybrid or all-

electric cars that use electricity from the electric grid. While electric powered cars are in early

stages of market penetration (e.g., Chevy Volt, Nissan Leaf) the majority of the transportation

infrastructure is already engineered to support automobiles that run on liquid fuels [10].

Consequently, liquid fuels have an inherent economic advantage with the current infrastructure.

Two methods that can enable penetration of natural gas or coal to the transportation

sector include natural gas to liquid (GTL) [11-15] and coal to liquid (CTL) [14-17] technology,

which yield syngas, a mixture of hydrogen and carbon monoxide, which can be further converted

to liquid fuels through an established technology, the Fischer-Tropsch process [18-23]. While

CTL and GTL are capable of addressing issues in the foreseeable future with limited fuel supply

and energy security, they are contributing towards global warming through the emission of CO2.

Carbon Capture and Sequestration (CCS) has been proposed as a means for long term CO2

storage in underground geologic formations away from the atmosphere [24,25]. However, issues

regarding long-term viability of this technology (e.g., leakage) could potentially negate the

positive effects of this technology [26].

Page 16: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

7

1.1.4 Global Warming

A primary advantage of

alternative fuel sources, such as

wind and solar, is their low

emission levels of CO2. This is

particularly important because the

concentration of CO2 in the

atmosphere has been increasing

since the industrial revolution, as

shown in Figure 1.5 [27-29].

Increasing CO2 concentrations in the

atmosphere have been directly linked to

an increase in global temperature,

which has been theorized to lead to

catastrophic effects on the ecosystem

[30,31]. Furthermore, as shown in

Figure 1.6, the emission levels for CO2

are expected to increase roughly 40 %

over the next 20 years [2].

1.1.5 Alternative Energy Sources

The need to reduce emissions

has renewed interest in a means to

expand penetration of “low emission”

Figure 1.5: Historical CO2 concentrations in the atmosphere from

trapped CO2 in Ice (Ice Dome) [29] and direct observation of CO2 in the

atmosphere at Mauna Loa Observatory (Mauna Loa, Hawaii). Data

from Francey, 1999 and NOAA/ESRL 2012 [28].

270

290

310

330

350

370

390

1000 1250 1500 1750 2000

CO

2C

on

cen

trat

ion

(p

pm

)

Year (AD)

Mauna Loa

Ice Dome

Figure 1.6: Historical and predicted CO2 emissions between

1980 and 2030. © National Petroleum Council 2007, reprinted

with permission from “Hard Truths: Facing the Hard Truths

About Energy: A Comprehensive View to 2030 of Global Oil

and Natural Gas” [2].

Page 17: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

8

renewable energy sources into the transportation sector. Several potential renewable energy

sources are making progress regarding improvements to the overall efficiency, including, biofuel

production form algae or biomass [32,33]. Additionally, wind, solar, and tidal power

technologies continue to become more efficient and less expensive per kW produced [34-36].

Furthermore, along with improved efficiencies, a favorable political environment for wind and

solar technology has resulted in large scale implementation of these technologies throughout the

United States [37].

While the efficiency, politics and economics of wind and solar have drastically improved

over the last century, the intermittent nature of the technology limits the extent of their

implementation. Specifically, the DOE recommended that the combined extent of wind and

solar implementation will max-out at 15 % of the United States’ electrical energy consumption

due to intermittent nature of these technologies [38]. Figure 1.7 shows examples from (a) a wind

turbine [39] and (b) a utility-scale photovoltaic system [40] to demonstrate typical fluctuations in

the power outputs from these sources.

While the electrical grid is the largest potential market for wind and solar, the electrical

grid does not respond well to intermittency. Electrical demand depends on many factors, such as

time of day, sun intensity and outside temperature. Consequently, traditional power plants must

scale their operation to match demand. This has several effects: First, power plants must be

designed to meet max-power demand, or else, entire cities will experience power “blackouts”

when demand for power is high. Second, the price of electricity is exceptionally high when

demand is at a maximum because power plants operate less efficiently to match high demand.

Third, power is often produced at a loss in off-peak times because power plants struggle to cut

back production and consequently receive a poor return from the market due an oversupply of

Page 18: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

9

power. Figure 1.7 demonstrates how the aforementioned effects influenced Ontario’s power

prices in July of 2011 [41]. Specifically, Ontario’s power prices fluctuated between roughly

18 ¢/kWh and roughly -7 ¢/kWh [41].

The addition of renewable sources of electricity, such as wind and solar, can potentially

help in “peak-power” demand times. However, because renewable sources of electricity, such as

wind and solar, are intermittent by nature [39,40], traditional fossil-fuel or nuclear power plants

still need to be built to meet max power demand, unless a means of storage is developed. Also,

much of the power generated from renewable sources in off-peak periods, when demand is low,

receive a minimal, if not negative, financial return for the power provided. Because the grid

does not respond well to intermittent power and because the renewable power sources discussed

above (wind and solar) are intermittent by nature, in many cases these renewable energy sources

are only economically viable with external stimuli.

Figure 1.7: (a) Power output for a wind turbine over 10 days (860,000 seconds). Reprinted from “The Spectrum of

Power from Wind Turbines”, 169, Jay Apt, 369-374, Copyright (2007) [39], with permission from Elsevier.

(b) power output for a utility-scale photovoltaic system over 6 days. Reprinted from The character of power output

from utility-scale photovoltaic systems, 16, Aimee E. Curtright, Jay Apt, 241-247, Copyright (2007), with

permission from John Wiley and Sons [40].

(a) (b)

Page 19: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

10

1.2 Electrochemical Reduction of CO2

1.2.1 Motivation for the Electrochemical Reduction of CO2

Here, we propose exploring the electrochemical reduction of CO2 as a means of

producing chemical fuels and feedstocks to store electrical energy from renewable sources such

as wind and solar in chemical form. When combined with a “green” renewable source, such as

wind or solar, this technology provides a means of (a) reducing our dependence on foreign oil

via enabling penetration of renewable technologies into the transportation sector and (b) reducing

CO2 emissions by moving away from fossil fuels.

An electrochemical process for CO2 reduction can be tied to renewable power sources in

several ways. First, an electrochemical reactor for chemical production can use energy produced

from a renewable power source in a continuous fashion. A second approach is to use a transient

process governed by the immediate economics in the following way: The energy from a

Figure 1.8: Historical electricity prices in Ontario for July 2011. Image courtesy of ieso “Power to Ontario. On

Demand” [40].

Page 20: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

11

renewable source is fed to an electrochemical reduction reactor only in off-peak times. Storing

electrical energy from intermittent power sources in chemical form only when the demand (and

price) for electricity is low has potential to improve the overall process economics for renewable

energy sources and thereby enable wider penetration of renewable technology (well beyond

15 %). Lastly, an electrochemical reactor could be tied to a fuel cell to (a) store electrical energy

in chemical form in off-peak times and (b) convert the chemicals back to electricity when the

price of electricity is once again high.

1.2.2 Explanation of the Electrochemical Reduction of CO2

The electrochemical conversion of CO2 relies on an electrochemical reactor that balances

electrocatalytic reduction and oxidation reactions. Figure 1.9 depicts the overall reaction

process. Specifically, on the anode (e.g., Pt as the catalyst), water is oxidized to make oxygen

Figure 1.9: (Left) Schematic of an electrochemical reactor for CO2 reduction. (Right) Standard Reduction

Potentials for the common cathode and anode reactions.

Electrolyte

CO

2

H+

CO

, C

H4,

C2H

4,

H2

(Pro

du

cts

)

H+

H+

e-

O2

O2

O2

O2

O2

O2

O2

Anode (e.g., Pt) Reactions

H2O→2H++2e-+½O2

Cathode (e.g., Ag, Cu) Reactions

2H++2e-→H2

2H++2e-+CO2→H2O+CO

2H++2e-+CO2→HCOOH

8H++8e-+CO2→2H2O+CH4

12H++12e-+2CO2→4H2O+C2H4

Gas Diffusion Electrodes

Eo (vs. SHE)*

-0.41

-0.52

-0.61

-0.25

-0.34

Eo (vs. SHE)*

+0.82

*at pH 7

Page 21: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

12

and protons. The evolved oxygen escapes the reactor via diffusion through micro-pores in the

anode gas diffusion electrode (GDE). The protons formed on the anode cross either a conductive

electrolyte stream or a conductive membrane to react on the cathode (e.g., Ag, Cu as the

catalyst). On the cathode, the protons can react to evolve H2, or, when reacting with CO2, to

form both water and either carbon monoxide, formic acid, methane, ethylene, or other C1, C2, or

C3 carbon products [42,43]. The carbon dioxide reaches the cathode via diffusion through the

cathode GDE to reach the cathode catalyst. Unlike a fuel cell, the reaction does not proceed

without an applied potential because the process is endothermic. Consequently, a potential must

be applied between the anode and cathode to overcome the activation energy barrier and generate

the potential necessary for the reaction to proceed. The theoretical potential necessary for the

reaction is the difference between the standard potentials for both the anode and cathode

reactions as shown in Figure 1.9 [44]. In reality, a higher potential must be applied, commonly

referred to as an overpotential, to overcome activation barriers to the reaction, often due to the

formation of high energy intermediates.

1.3 Challenges for the Electrochemical Reduction of CO2

As mentioned above, the electrochemical reduction of CO2 has potential to expand the

use of renewable energy sources and enable penetration of other sources into the transportation

sector. However, for large scale implementation of this technology, improvements are needed

regarding: (a) energetic efficiency, (b) reaction rate, and (c) product selectivity.

While all three of the aforementioned challenges will benefit from improved catalysts,

better catalysts are critical to improving the energetic efficiency of the reaction, as catalysts are

the only way to reduce the overpotentials for the reaction. Specifically for CO2 reduction, the

intermediate formed after the first electron transfer, CO2- requires an overpotential of -1.9 V vs.

Page 22: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

13

SHE [45]. While transition metal catalysts reduce the overpotential for this reaction, as shown

schematically in Figure 1.10, further reduction in the overpotential can be achieved via either

improved catalysts of co-catalyst electrolytes. Specifically, Hori explored an extensive list of

transition catalysts and reported their product selectivity and onset potentials [42]. Furthermore,

Bocarsly et al. demonstrated that Pyridine can serve as a co-catalyst with Cu electrodes to reduce

the onset potential for methanol production [46-49].

Improving the reaction rate is another critical challenge for CO2 reduction as reaction rate

influences capital expenditures for both the reactor and the catalyst. Like the reaction

overpotential, the reaction rate is also drastically influenced by catalysis, however, other factors

such as electrolyte interactions, operation conditions, and reactor design influence the reaction

rate. Lister et al. demonstrated that both the partial pressure of CO2 and the reactor temperature

influence the partial reaction rates for both hydrogen and carbon monoxide formation [50].

Figure 1.10: Schematic of the overpotential for CO2 reduction. The red curve is a depiction of how a catalyst can

be used to lower the overpotential for the formation of a CO2 radical intermediate compared to the scenario without

a catalyst (red). Reprinted with permission from: D. T. Whipple and P. J. A. Kenis, “Prospects of CO2 Utilization

via Direct Heterogeneous Electrochemical Reduction” Journal of Physical Chemistry Letters, 2010. 1 (24): p. 3451-

3458.. Copyright (2010) American Chemical Society [9].

Page 23: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

14

From this, an optimal temperature can be identified, beyond which and below which, the reaction

rate for CO formation decreases. Kaneco and others showed that electrolyte composition

drastically influences the partial reaction rates of multiple carbon products on mercury or copper

electrodes [51-53]. Whipple et al. showed that the electrolyte pH drastically influences the

partial reaction rate for hydrogen and formic acid formation on a tin electrode [54].

Furthermore, several reactor designs have been developed to improve the reaction rates [55-61].

The last requirement for large scale implementation of this technology is optimization of

the selectivity for a desired product. Many of the same factors influencing reaction rate

influence product selectivity. First, catalyst choice more so than virtually any other factor,

Figure 1.11: Cathode potentials, current densities and Faradiac efficiencies on various transition metal electrodes in

a 0.1 M KHCO3 electrolyte. Reprinted from “Electrocatalytic process of CO selectivity in electrochemical reduction

of CO2 at metal electrodes in aqueous media”, 39, Yoshio Hori, Hidetoshi Wakebe, Toshio Tsukamoto, Osamu

Koga, 1833-1839, Copyright (1994), with permission from Elsevier [62].

Page 24: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

15

controls product selection. Figure 1.11 shows product distributions from the electrochemical

reduction of CO2 on a wide range of catalysts as tabulated by Hori et al. [62]. Four classes of

catalysts exist for the production of carbon products via the electrochemical reduction of CO2.

First, copper has been shown to be selective for the production of methanol [63,64], methane

[51,53,65-67], ethylene [51,67-70], and sometimes larger carbon products, such as propane,

propanol, etc. [43]. Ag, Au, Zn, Pd, and Ga are selective for the production of carbon monoxide

[42,55-57,71]. Pb, Hg, In, Sn, Cd, and Ti are selective for the production of formic acid

[9,42,72-79]. Lastly, Ni, Fe, Pt and Ti are selective for the production of H2.

While Hori tabulated the selectivity of a wide range of catalysts in a 0.1 M KHCO3

solution, the selectivity can be tuned by changing the electrolyte composition. Several

Figure 1.12: Current densities and Faradaic efficiencies of a copper electrode with a variety of electrolyte solutions.

Reprinted (adapted) with permission from: S. Kaneco, K. Iiba, H. Katsumata, T. Suzuki and K. Ohta,

“Electrochemical Reduction of High Pressure CO2 at a Cu Electrode in Cold Methanol.” Electrochimica Acta, 2006.

51 (23): p. 4880-4885. Kaneco 2006. Copyright (2006) American Chemical Society [53].

Page 25: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

16

researchers have shown that various alkali cations influence the selectivity for methane and

ethylene on a copper electrode [53,80] as well as formic acid and hydrogen on a mercury

electrode [52]. Figure 1.12 shows the effect of a wide range of electrolyte solutions on product

selection with a copper electrode [51,53]. Additionally, pyridine in the electrolyte solution was

found to change the product distribution of a copper electrode from methane, ethylene, hydrogen

and carbon monoxide to methane [46,49,81].

1.4 Economic Analysis

The power required to produce a given mass of a product via electrochemical reduction

of CO2 can be calculated based on the applied reactor potential using the following equation:

(1.1)

where V is the potential applied to the reactor, F is Faraday’s constant, ci is the number of

electrons required for the reaction, and MW is the product molecular weight. At a typical reactor

potential of 2.5 V, the power required for the electrochemical reduction of CO2 to CO is

4,342 kWhr/ton. The price of CO on the market is roughly $1,000 per ton [82]. If ignoring

capital costs, transportation costs, and all other costs besides the direct electricity cost for the

reduction process, the financial return for producing a chemical per unit of electricity consumed

is the break-even price of electricity. The break-even price of electricity could be calculated

using the following equation:

(1.2)

Using this equation, the break-even cost of electricity is: 23 ¢/kWhr, which is greater than

the maximum price of electricity shown earlier in Ontario for an expensive summer month. As

long as electricity stays below this price, an electrochemical process for CO2 reduction could be

Page 26: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

17

economically favorable with the other expenditures (e.g., capital, transportation, storage)

remaining low.

1.5 Remaining Challenges

Better catalysts and optimized reaction conditions have improved reaction kinetics and

thereby enabled current densities greater than 600 mA and energetic efficiencies greater than

70 % for the electrochemical reduction of CO2 to CO; however, an economical process requires

both high energetic efficiencies and current densities [9]. Figure 1.13 shows a comparison of

data in the literature for the electrochemical reduction of CO2 into formic acid, carbon monoxide,

and larger hydrocarbons, e.g., CH4, C2H4, with regards to energetic efficiency and current

Figure 1.13: Comparison of the energy efficiencies and partial current densities for CO2 reduction to formic acid,

syngas, and hydrocarbons (methane and ethylene) reported in literature with those of water electrolyzers.

Efficiencies of electrolyzers are total system efficiencies, while the CO2 conversion efficiencies only include

cathode losses and neglect anode and system losses. Reprinted with permission from: D. T. Whipple and P. J. A.

Kenis, “Prospects of CO2 Utilization via Direct Heterogeneous Electrochemical Reduction” Journal of Physical

Chemistry Letters, 2010. 1 (24): p. 3451-3458.. Copyright (2010) American Chemical Society [9].

Page 27: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

18

density [9]. Figure 1.13 classifies the catalysts into three categories based on their products.

First, the reactions selective for larger hydrocarbons achieve high current densities at the expense

of energetic efficiency. While the direct conversion of CO2 to hydrocarbons is attractive due to

the immediate application of the product, the overall efficiency is far too low for

commercialization. Second, the reactions selective for formic acid achieve intermediate current

densities and intermediate energetic efficiencies. Third, the reactions selective for CO achieve

high energetic efficiencies at low current densities. Here, we chose to study the electrochemical

reaction for CO as it has the highest energetic efficiencies compared to other reactions in the

literature. An important consideration is how this technology compares to a more traditional

technology, water electrolysis. Figure 1.13 shows that the energetic efficiencies and current

densities for water electrolysis are much higher than those achieved for CO reduction. However,

using hydrogen as a fuel has inherent disadvantages due to challenges with both storage and

transportation of the compressed hydrogen gas. While transportation and storage issues exist

with water electrolysis, the efficiency and current density of a CO2 reduction process should be

competitive with water electrolyzers.

In summary, the electrochemical reduction of CO2 has tremendous promise as a means of

energy storage and enabling penetration of electrical energy into other energy sectors, such as

transportation. However, improvements are necessary primarily regarding the energetic efficiency

and current density of the reaction. The purpose of this dissertation is to improve the energetic

efficiency and current density for the electrochemical reduction of CO2 to CO using a microfluidic

reactor.

Chapter 2 details the influence of electrolytes on the electrochemical reduction of CO2 on a

Ag electrode. Chapter 3 details the development of an electrochemical platform to study the use of

an ionic liquid as a co-catalyst. Chapter 4 demonstrates a means to achieve high current densities

Page 28: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

19

with ionic liquids. Chapter 5 explores the use of organometallic catalysts. Chapter 6 presents a study

of the role of local geometry on a methanol laminar flow fuel cell, a reactor configuration that can

also be applied to electrochemical CO2 conversion. Chapter 7 summarizes the key findings and

conclusions obtained from this work.

1.6 References

1. J.B. Reece, M.L. Cain, L.A. Urry, P.V. Minorsky and S.A. Wasserman, Campbell

Biology, Pearson Benjamin Cummings, 2010.

2. National Petroleum Council. Committee on Global Oil and Gas., Hard truths: Facing the

Hard Truths About Energy: A Comprehensive View to 2030 of Global Oil and Natural

Gas, National Petroleum Council, Washington, DC, 2007.

3. United Nations. Department of Economic Affairs: Population Division., World

Population to 2300: Economic & Social Affairs, United Nations, New York, NY, 2012.

4. Organization for Economic Co-orporation and Development (OECD): International

Energy Agency, World Energy Outlook, International Energy Agency, Paris, France,

2010.

5. R. W. Bentley, Global oil & gas depletion: an overview, Energy Policy, 2002, 30 (3), p.

189-205.

6. R. Guseo, A. Dalla Valle and M. Guidolin, World Oil Depletion Models: Price effects

compared with strategic or technological interventions, Technological Forecasting and

Social Change, 2007, 74 (4), p. 452-469.

7. J. S. Harvey, B. P. Lyons, T. S. Page, C. Stewart and J. M. Parry, An assessment of the

genotoxic impact of the Sea Empress oil spill by the measurement of DNA adduct levels

in selected invertebrate and vertebrate species, Mutation Research/Genetic Toxicology

and Environmental Mutagenesis, 1999, 441 (1), p. 103-114.

8. National Petroleum Council. Committee on Global Oil and Gas., Prudent Development:

Realizing the Potential of North America's Abundant Natural Gas and Oil, National

Petroleum Council, Washington, D.C., 2011.

9. D. T. Whipple and P. J. A. Kenis, Prospects of CO2 Utilization via Direct Heterogeneous

Electrochemical Reduction, Journal of Physical Chemistry Letters, 2010, 1 (24), p. 3451-

3458.

10. S. Beggs, S. Cardell and J. Hausman, Assessing the potential demand for electric cars,

Journal of Econometrics, 1981, 17 (1), p. 1-19.

11. R. Krishna, J. Ellenberger and S. T. Sie, Reactor development for conversion of natural

gas to liquid fuels: A scale-up strategy relying on hydrodynamic analogies, Chemical

Engineering Science, 1996, 51 (10), p. 2041-2050.

12. N. D. Parkyns, C. I. Warburton and J. D. Wilson, Natural gas conversion to liquid fuels

and chemicals: Where does it stand?, Catalysis Today, 1993, 18 (4), p. 385-442.

Page 29: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

20

13. Buping Bao, Mahmoud M. El-Halwagi and Nimir O. Elbashir, Simulation, integration,

and economic analysis of gas-to-liquid processes, Fuel Processing Technology, 2010, 91

(7), p. 703-713.

14. Mikael Höök and Kjell Aleklett, A review on coal-to-liquid fuels and its coal

consumption, International Journal of Energy Research, 2010, 34 (10), p. 848-864.

15. Dragomir Bukur, Dr Mahmood Amani and Bashar Hassna, Reactors and Catalysts for

Gas-to-Liquids (GTL) and Coal-to-Liquids (CTL) Technologies, QNRS Repository,

2011, (2011), p. 3954.

16. Delanie Lamprecht, Reinier Nel and Dieter Leckel, Production of On-Specification Fuels

in Coal-to-Liquid (CTL) Fischer−Tropsch Plants Based on Fixed-Bed Dry Bottom Coal

Gasification, Energy & Fuels, 2009, 24 (3), p. 1479-1486.

17. X. Hao, G. Dong, Y. Yang, Y. Xu and Y. Li, Coal to Liquid (CTL): Commercialization

Prospects in China, Chemical Engineering & Technology, 2007, 30 (9), p. 1157-1165.

18. C. Oloman and H. Li, Electrochemical Processing of Carbon Dioxide, ChemSusChem,

2008, 1 (5), p. 385-391.

19. C. M. Sanchez-Sanchez, V. Montiel, D. A. Tryk, A. Aldaz and A. Fujishima,

Electrochemical Approaches to Alleviation of the Problem of Carbon Dioxide

Accumulation, Pure and Applied Chemistry, 2001, 73 (12), p. 1917-1927.

20. S. Choi, J. H. Drese and C. W. Jones, Adsorbent Materials for Carbon Dioxide Capture

from Large Anthropogenic Point Sources, ChemSusChem, 2009, 2 (9), p. 796-854.

21. X. Hao, M. E. Djatmiko, Y. Y. Xu, Y. N. Wang, J. Chang and Y. W. Li, Simulation

Analysis of a GTL Process Using ASPEN Plus, Chemical Engineering & Technology,

2008, 31 (2), p. 188-196.

22. M. Sudiro and A. Bertucco, Production of Synthetic Gasoline and Diesel Fuel by

Alternative Processes Using Natural Gas and Coal: Process Simulation and

Optimization, Energy, 2009, 34 (12), p. 2206-2214.

23. M. Sudiro, A. Bertucco, F. Ruggeri and A. Fontana, Improving Process Performances in

Coal Gasification for Power and Synfuel Production, Energy & Fuels, 2008, 22 (6), p.

3894-3901.

24. Soren Anderson and Richard Newell, Prospects for carbon capture and storage

technologies, 2003.

25. Heleen de Coninck, Jennie C. Stephens and Bert Metz, Global learning on carbon

capture and storage: A call for strong international cooperation on CCS demonstration,

Energy Policy, 2009, 37 (6), p. 2161-2165.

26. Michael L. Szulczewski, Christopher W. MacMinn, Howard J. Herzog and Ruben

Juanes, Lifetime of carbon capture and storage as a climate-change mitigation

technology, Proceedings of the National Academy of Sciences, 2012, 109 (14), p. 5185-

5189.

27. Charles D. Keeling, Robert B. Bacastow, Arnold E. Bainbridge, Carl A. Ekdahl, Peter R.

Guenther, Lee S. Waterman and John F. S. Chin, Atmospheric carbon dioxide variations

at Mauna Loa Observatory, Hawaii, Tellus, 1976, 28 (6), p. 538-551.

28. NOAA/ESRL (www.esrl.noaa.gov/gmd/ccgg/trends/) and Dr. Ralph Keeling Dr. Pieter

Tans, Scripps Institution of Oceanography (scrippsco2.ucsd.edu/).

Page 30: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

21

29. R. J. Francey, C. E. Allison, D. M. Etheridge, C. M. Trudinger, I. G. Enting, M.

Leuenberger, R. L. Langenfelds, E. Michel and L. P. Steele, A 1000-year high precision

record of δ13C in atmospheric CO2, Tellus B, 1999, 51 (2), p. 170-193.

30. Terry L. Root, Jeff T. Price, Kimberly R. Hall, Stephen H. Schneider, Cynthia

Rosenzweig and J. Alan Pounds, Fingerprints of global warming on wild animals and

plants, Nature, 2003, 421 (6918), p. 57-60.

31. Malte Meinshausen, Nicolai Meinshausen, William Hare, Sarah C. B. Raper, Katja

Frieler, Reto Knutti, David J. Frame and Myles R. Allen, Greenhouse-gas emission

targets for limiting global warming to 2[thinsp][deg]C, Nature, 2009, 458 (7242), p.

1158-1162.

32. C. Knocke and J. Vogt, Biofuels - Challenges & Chances: How Biofuel Development Can

Benefit from Advanced Process Technology, Engineering in Life Sciences, 2009, 9 (2), p.

96-99.

33. P. J. L. Williams and L. M. L. Laurens, Microalgae as Biodiesel & Biomass Feedstocks:

Review & Analysis of the Biochemistry, Energetics & Economics, Energy &

Environmental Science, 2010, 3 (5), p. 554-590.

34. D. Abbott, Keeping the Energy Debate Clean: How Do We Supply the World's Energy

Needs?, Proceedings of the Ieee, 2010, 98 (1), p. 42-66.

35. O. Langhamer, K. Haikonen and J. Sundberg, Wave Power-Sustainable Energy or

Environmentally Costly? A Review with Special Emphasis on Linear Wave Energy

Converters, Renewable & Sustainable Energy Reviews, 2010, 14 (4), p. 1329-1335.

36. M. Lenzen, Current State of Development of Electricity-Generating Technologies: A

Literature Review, Energies, 2010, 3 (3), p. 462-591.

37. Holly Klick and Eric R. A. N. Smith, Public understanding of and support for wind

power in the United States, Renewable Energy, 2010, 35 (7), p. 1585-1591.

38. J.M. Duthie and H.W. Whittington, Securing Renewable Energy Supplies Through

Carbon Dioxide Storage in Methanol, Power Engineering Society Summer Meeting,

2002 IEEE, 2002, 1 p. 145-150.

39. Jay Apt, The spectrum of power from wind turbines, Journal of Power Sources, 2007, 169

(2), p. 369-374.

40. Aimee E. Curtright and Jay Apt, The character of power output from utility-scale

photovoltaic systems, Progress in Photovoltaics: Research and Applications, 2008, 16 (3),

p. 241-247.

41. Anonymous, Monthly Market Report: July 2011. ieso: Power to Ontario. On Demand.,

2011.

42. Y. Hori, H. Wakebe, T. Tsukamoto and O. Koga, Electrocatalytic Process of CO

Selectivity in Electrochemical Reduction of CO2 at Metal-Electrodes in Aqueous-Media,

Electrochimica Acta, 1994, 39 (11-12), p. 1833-1839.

43. Kendra P. Kuhl, Etosha R. Cave, David N. Abram and Thomas F. Jaramillo, New insights

into the electrochemical reduction of carbon dioxide on metallic copper surfaces, Energy

& Environmental Science, 2012, 5 (5), p. 7050-7059.

44. Y. Hori, Electrochemical CO2 Reduction on Metal Electrodes, Modern Aspects of

Electrochemistry, 2008, 42 p. 89-189.

Page 31: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

22

45. Etsuko Fujita, Photochemical carbon dioxide reduction with metal complexes,

Coordination Chemistry Reviews, 1999, 185–186 (0), p. 373-384.

46. Emily Barton Cole, Prasad S. Lakkaraju, David M. Rampulla, Amanda J. Morris, Esta

Abelev and Andrew B. Bocarsly, Using a One-Electron Shuttle for the Multielectron

Reduction of CO2 to Methanol: Kinetic, Mechanistic, and Structural Insights, Journal of

the American Chemical Society, 2010, 132 (33), p. 11539-11551.

47. E. B. Cole, P. S. Lakkaraju, D. M. Rampulla, A. J. Morris, E. Abelev and A. B. Bocarsly,

Using a One-Electron Shuttle for the Multielectron Reduction of CO2 to Methanol:

Kinetic, Mechanistic, and Structural Insights, Journal of the American Chemical Society,

2010, 132 (33), p. 11539-11551.

48. A. J. Morris, R. T. McGibbon and A. B. Bocarsly, Electrocatalytic Carbon Dioxide

Activation: The Rate-Determining Step of Pyridinium-Catalyzed CO2 Reduction,

Chemsuschem, 2011, 4 (2), p. 191-196.

49. G. Seshadri, C. Lin and A. B. Bocarsly, A New Homogeneous Electrocatalyst for the

Reduction of Carbon-Dioxide to Methanol at Low Overpotential, Journal of

Electroanalytical Chemistry, 1994, 372 (1-2), p. 145-150.

50. Eric Dufek, Tedd Lister and Michael McIlwain, Bench-scale electrochemical system for

generation of CO and syn-gas, Journal of Applied Electrochemistry, 2011, 41 (6), p. 623-

631.

51. S. Kaneco, K. Iiba, H. Katsumata, T. Suzuki and K. Ohta, Electrochemical Reduction of

High Pressure CO2 at a Cu Electrode in Cold Methanol, Electrochimica Acta, 2006, 51

(23), p. 4880-4885.

52. Y. Hori and S. Suzuki, Electrolytic Reduction of Carbon-Dioxide at Mercury-Electrode

in Aqueous-Solution, Bulletin of the Chemical Society of Japan, 1982, 55 (3), p. 660-665.

53. S. Kaneco, H. Katsumata, T. Suzuki and K. Ohta, Electrochemical Reduction of CO2 to

Methane at the Cu Electrode in Methanol with Sodium Supporting Salts and its

Comparison with Other Alkaline Salts, Energy Fuels, 2006, 20 (1), p. 409-414.

54. D. T. Whipple, E. C. Finke and P. J. A. Kenis, Microfluidic Reactor for the

Electrochemical Reduction of Carbon Dioxide: The Effect of pH, Electrochemical and

Solid State Letters, 2010, 13 (9), p. D109-D111.

55. T. Yamamoto, D. A. Tryk, A. Fujishima and H. Ohata, Production of Syngas Plus

Oxygen from CO2 in a Gas-Diffusion Electrode-Based Electrolytic Cell, Electrochimica

Acta, 2002, 47 (20), p. 3327-3334.

56. C. Delacourt, P. L. Ridgway, J. B. Kerr and J. Newman, Design of an Electrochemical

Cell Making Syngas (CO+H2) from CO2 and H2O Reduction at Room Temperature,

Journal of the Electrochemical Society, 2008, 155 (1), p. B42-B49.

57. F. Bidrawn, G. Kim, G. Corre, J. T. S. Irvine, J. M. Vohs and R. J. Gorte, Efficient

Reduction of CO2 in a Solid Oxide Electrolyzer, Electrochemical and Solid-State Letters,

2008, 11 (9), p. B167-B170.

58. S. D. Ebbesen and M. Mogensen, Electrolysis of Carbon Dioxide in Solid Oxide

Electrolysis Cells, Journal of Power Sources, 2009, 193 (1), p. 349-358.

Page 32: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

23

59. S. H. Jensen, X. F. Sun, S. D. Ebbesen, R. Knibbe and M. Mogensen, Hydrogen and

Synthetic Fuel Production Using Pressurized Solid Oxide Electrolysis Cells, International

Journal of Hydrogen Energy, 2010, 35 (18), p. 9544-9549.

60. V. Kaplan, E. Wachtel, K. Gartsman, Y. Feldman and I. Lubomirsky, Conversion of CO2

to CO by Electrolysis of Molten Lithium Carbonate, Journal of the Electrochemical

Society, 2010, 157 (4), p. B552-B556.

61. Z. L. Zhan and L. Zhao, Electrochemical Reduction of CO2 in Solid Oxide Electrolysis

Cells, Journal of Power Sources, 2010, 195 (21), p. 7250-7254.

62. Yoshio Hori, Hidetoshi Wakebe, Toshio Tsukamoto and Osamu Koga, Electrocatalytic

process of CO selectivity in electrochemical reduction of CO2 at metal electrodes in

aqueous media, Electrochimica Acta, 1994, 39 (11–12), p. 1833-1839.

63. G. Arai, T. Harashina and I. Yasumori, Selective Electrocatalytic Reduction of Carbon-

Dioxide to Methanol on Ru-Modified Electrode, Chemistry Letters, 1989, (7), p. 1215-

1218.

64. D. P. Summers, S. Leach and K. W. Frese, The Electrochemical Reduction of Aqueous

Carbon-Dioxide to Methanol at Molybdenum Electrodes with Low Overpotentials,

Journal of Electroanalytical Chemistry, 1986, 205 (1-2), p. 219-232.

65. S. Kaneco, K. Iiba, H. Katsumata, T. Suzuki and K. Ohta, Electrochemical Reduction of

High Pressure Carbon Dioxide at a Cu Electrode in Cold Methanol with CsOH

Supporting Salt, Chemical Engineering Journal, 2007, 128 (1), p. 47-50.

66. S. Kaneco, R. Iwao, K. Iiba, K. Ohta and T. Mizuno, Electrochemical Conversion of

Carbon Dioxide to Formic Acid on Pb in KOH/Methanol Electrolyte at Ambient

Temperature and Pressure, Energy, 1998, 23 (12), p. 1107-1112.

67. R. L. Cook, R. C. Macduff and A. F. Sammells, High-Rate Gas-Phase CO2 Reduction to

Ethylene and Methane Using Gas-Diffusion Electrodes, Journal of the Electrochemical

Society, 1990, 137 (2), p. 607-608.

68. S. Kaneco, H. Katsumata, T. Suzuki and K. Ohta, Electrochemical Reduction of Carbon

Dioxide to Ethylene at a Copper Electrode in Methanol Using Potassium Hydroxide and

Rubidium Hydroxide Supporting Electrolytes, Electrochimica Acta, 2006, 51 (16), p.

3316-3321.

69. T. Mizuno, K. Ohta, M. Kawamoto and A. Saji, Electrochemical Reduction of CO2 on Cu

in 0.1 M KOH-methanol, Energy Sources, 1997, 19 (3), p. 249-257.

70. J. Yano and S. Yamasaki, Pulse-mode Electrochemical Reduction of Carbon Dioxide

Using Copper and Copper Oxide Electrodes for Selective Ethylene Formation, Journal of

Applied Electrochemistry, 2008, 38 (12), p. 1721-1726.

71. N. Furuya and S. Koide, Electroreduction of Carbon-Dioxide by Metal Phthalocyanines,

Electrochimica Acta, 1991, 36 (8), p. 1309-1313.

72. M. Azuma, K. Hashimoto, M. Hiramoto, M. Watanabe and T. Sakata, Electrochemical

Reduction of Carbon-Dioxide on Various Metal-Electrodes in Low-Temperature Aqueous

KHCO3 Media, Journal of the Electrochemical Society, 1990, 137 (6), p. 1772-1778.

73. N. Furuya, T. Yamazaki and M. Shibata, High Performance Ru-Pd Catalysts for CO2

Reduction at Gas-Diffusion Electrodes, Journal of Electroanalytical Chemistry, 1997, 431

(1), p. 39-41.

Page 33: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

24

74. M. N. Mahmood, D. Masheder and C. J. Harty, Use of Gas-Diffusion Electrodes for

High-Rate Electrochemical Reduction of Carbon-Dioxide .1. Reduction at Lead, Indium-

Impregnated and Tin-Impregnated Electrodes, Journal of Applied Electrochemistry,

1987, 17 (6), p. 1159-1170.

75. Y. Akahori, N. Iwanaga, Y. Kato, O. Hamamoto and M. Ishii, New Electrochemical

Process for CO2 Reduction to Formic Acid from Combustion Flue Gases,

Electrochemistry, 2004, 72 (4), p. 266-270.

76. F. Koleli, T. Atilan, N. Palamut, A. M. Gizir, R. Aydin and C. H. Hamann,

Electrochemical Reduction of CO2 at Pb- and Sn-electrodes in a Fixed-bed Reactor in

Aqueous K2CO3 and KHCO3 Media, J. of Appl. Electrochem., 2003, 33 (5), p. 447-450.

77. H. Li and C. Oloman, Development of a Continuous Reactor for the Electro-reduction of

Carbon Dioxide to Formate - Part 2: Scale-up, Journal of Applied Electrochemistry,

2007, 37 (10), p. 1107-1117.

78. K. Subramanian, K. Asokan, D. Jeevarathinam and M. Chandrasekaran, Electrochemical

Membrane Reactor for the Reduction of Carbon Dioxide to Formate, Journal of Applied

Electrochemistry, 2007, 37 (2), p. 255-260.

79. B. Innocent, D. Liaigre, D. Pasquier, F. Ropital, J. M. Leger and K. B. Kokoh, Electro-

reduction of Carbon Dioxide to Formate on Lead Electrode in Aqueous Medium, Journal

of Applied Electrochemistry, 2009, 39 (2), p. 227-232.

80. A. Murata and Y. Hori, Product Selectivity Affected by Cationic Species in

Electrochemical Reduction of CO2 and CO at a Cu Electrode, Bulletin of the Chemical

Society of Japan, 1991, 64 (1), p. 123-127.

81. Amanda J. Morris, Robert T. McGibbon and Andrew B. Bocarsly, Electrocatalytic

Carbon Dioxide Activation: The Rate-Determining Step of Pyridinium-Catalyzed CO2

Reduction, ChemSusChem, 2011, 4 (2), p. 191-196.

82. Narasi Sridhar, Carbon Dioxide Utilization: Electrochemical Conversion of CO2 -

Opportunities and Challenges. Det Norske Veritas, 2011. 07 1-20.

Page 34: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

25

Chapter 2

Effect of Ions on the Electrochemical Conversion of CO2 to CO1

2.1 Introduction

Carbon dioxide (CO2) concentrations in the atmosphere have been linked to climate

change [1]. A reduction in CO2 emissions and/ or an increase in CO2 capture may be needed to

stabilize CO2 concentrations in the atmosphere, and consequently, minimize the negative effects

of CO2 as a greenhouse gas [2-6]. Additionally, to reduce global dependence on fossil fuels,

approaches to ensure further penetration of alternative energy sources need to be developed.

Furthermore, uneven global distribution of oil reserves present potential energy security issues

for oil importing countries (e.g., US and Europe) [7]. Natural gas to liquid (GTL) and coal to

liquid (CTL) technologies are being developed to reduce dependence on imported petroleum [7].

While GTL and CTL can provide a means of decreasing dependence on imported oil, these

technologies do not address the non-renewable nature of fossil fuels. To reduce atmospheric

CO2 emissions, one technology being implemented is carbon capture and sequestration [5,8-10].

Additionally, carbon neutral energy sources such as wind and solar are becoming increasingly

important sources of energy. However, in the current infrastructure, the potential to replace

traditional non-renewable sources with the aforementioned renewable sources is limited due to

their intermittent nature. For example, in the US, the use of these energy sources will be limited

to a maximum of ~15 % of the US electricity supply without improved power output leveling

(smart grid) [11]. Conversion of CO2 to useable fuels provides a means of decreasing

dependence on foreign oil via the penetration of renewable energy sources into the transportation

1 Part of this work has been submitted for publication: Thorson, M. R, Siil K.I., Kenis, P. J. A., “Effect of Cations on

the Electrochemical Conversion CO2 to CO.” 2012

Page 35: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

26

sector, while expanding the use of intermittent, renewable power sources (wind and solar). Since

transportation is responsible for about 33 % of CO2 emissions in the United States, introducing a

carbon neutral fuel would have a drastic impact on CO2 concentrations in the atmosphere [12].

Multiple methods (e.g., photochemical, thermochemical, and electrochemical) are being

explored for the conversion of CO2 into useful fuels. Specifically, electrochemical reduction of

CO2 combines electrical energy from a carbon neutral source with CO2 to yield many products,

including: methanol [13-15], formic acid [15-21], methane [15,22-25], ethylene [15,24-28], and

carbon monoxide (CO) [15,29-32]. A broad spectrum of metal catalysts has been tested for the

production of various fuels and feedstocks from CO2 reduction [15,16,21,25]. Carbon

monoxide, which is primarily produced using Ag and Au as the catalysts, is particularly

interesting as it can be combined with H2 to yield syngas, which can be further reacted in a

Fischer-Tropsch process to fuels such as methanol or “Fischer-Tropsch diesel” [33-35].

Consequently, several reactor designs have been explored for CO production [30-32,36-39]. As

we reported earlier, the primary hurdle to the further development of electrochemical CO2

reduction processes lies in the challenge to simultaneously maximize energetic efficiency and

throughput [7].

Previously, we reported a flow reactor for electrochemical CO2 conversion to formic acid

[18]. Using a similar reactor, we showed that amine based ionic liquid electrolytes can

drastically reduce the overpotential necessary for the electrochemical reduction of CO2 to CO,

albeit at the expense of throughput [40]. Pyridine has been shown to have similar effects

[41,42]. Additionally, several researchers have shown that various alkali cations influence the

selectivity for methane and ethylene on a copper electrode [23,43] as well as formic acid and

hydrogen on a mercury electrode [44]; however, little work has been published regarding the

Page 36: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

27

influence of electrolyte choice with the most prominent catalysts for CO2 reduction, Ag or Au.

The study of electrolytes is further complicated by effects from the electrolyte solutions on

solution properties, such as conductivity, solubility and pH, which need to be de-convoluted.

To date, most efforts for the electrochemical reduction of CO2 to CO have focused on

researching different catalysts to optimize energetic efficiency and/or current density. Here we

investigate in detail the influence of alkali cations (i.e., Li+, Na

+, K

+, Rb

+, and Cs

+) with two

halogens (i.e., Cl- and I

-) and hydroxide (OH

-) on the electrochemical reduction of CO2 to CO

with regards to current density and selectivity.

2.2 Experimental

2.2.1 Electrochemical cell

A schematic of the electrochemical reactor used in this study is shown in Figure 2.1.

Stainless steel plates (6 x 3 cm) served as current collectors, which held the fuel cell together and

provided electrical contact between the gas diffusion electrode (GDE) and an external

potentiostat (PGSTAT-30, EcoChemie). The potentiostat was connected to the cell via banana

clips which plugged directly into precisely machined 3/16” holes in the current collectors. Two

1-mm thick poly(methyl methacrylate) (PMMA) windows with precisely machined 0.5-cm wide

by 2.0-cm long channels provided the electrolyte flow fields within the electrochemical reactor.

Figure 2.1: Schematic of an electrochemical reactor used to study influence of various electrolyte solutions on the

electrochemical reduction of CO2 to CO.

Page 37: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

28

A 212 Nafion membrane (DuPont) separated the two PMMA windows. On the cathode side, the

catalyst was applied only to the first 1.5 cm and a polytetrafluoroethylene (PTFE) filter (10 µm,

Pall Life Sciences) was added to cover the last 0.5 cm. This facilitated the removal of bubbles

from oxygen evolution on the anode that entered the electrolyte stream and could otherwise

collect and block electrolyte contact with the catalyst. The cathode current collector had a

precisely machined 0.5-cm wide by 2.0-cm long by 0.25-cm deep window behind the GDEs to

allow for the flow of gases in and out of the reactor. The anode was left open to air for oxygen to

escape. Four insulated stainless steel bolts held the entire setup together.

2.2.2 Electrode preparation

The electrodes were prepared as previously reported [18] using Sigracet “35BC”-type

GDEs. In short, a suspension of catalyst and Nafion binder was prepared with a 50/50 mixture of

water and isopropyl alcohol, which was then sonicated and subsequently painted on a GDE

followed by hot-pressing. The cathodes consisted of 5 mg/cm2 Ag and 0.1 mg/cm

2 Nafion. The

anode was 2 mg/cm2 Pt black with 0.1 mg/cm

2 Nafion. All electrodes were hot-pressed at 130°C

and 2000 kPa for 15 min, similar to the procedure previously reported [18].

2.2.3 Electrolyte composition

All salts were purchased and used as provided. The following electrolyte solutions were

prepared in ultrapure water (18.2 MΩ∙cm) such that the final concentrations were 1.0 M: LiCl

(Product Number: L121; Purity: 98.5 %; Provider: Fischer Scientific), NaCl (S271; 99 %;

Fischer Scientific), KCl (P1597; 99 %; Sigma Aldrich), CsCl (289329; 99.9 %; Sigma Aldrich),

LiI (518018; 99.9 %; Sigma Aldrich), NaI (383112; 99.5 %; Sigma Aldrich), KI (P410; 99 %;

Fischer Scientific), CsI (202134; 99.9 %; Sigma Aldrich), NaOH (221465; 97 %; Sigma

Aldrich), KOH (44016; 90 %; Sigma Aldrich), RbOH (243698; 99 %; metal basis Sigma

Page 38: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

29

Aldrich), and CsOH (C8518; 99.5 %; metal basis Sigma Aldrich). For the pH studies, a 1 M KCl

solution and the pH was adjusted with dilute solutions of KOH or HCl to achieve the desired

final pH levels.

2.2.4 Cell testing

An Autolab potentiostat (PGSTAT-30, EcoChemie) controlled the cell potential and

measured the resulting current as previously reported [18,45]. The individual electrode

potentials were measured using multimeters connected to each electrode and a Ag/AgCl

reference electrode (RE-5B, BASi) in the exit stream. The cell was allowed to reach steady state

for 200 s, after which, the gas flowed into a gas chromatograph (GC) (Trace GC, Thermo) for

composition analysis of H2 and CO in the affluent gas streams. The current was averaged for an

additional 150 s before stepping to the next potential. All experiments were run at ambient

conditions. The electrode potentials were not corrected for IR drop. A mass flow controller

(32907-80, Cole Parmer) was used to flow CO2 from a cylinder at 25 sccm Also, a syringe pump

(PHD 2000, Harvard Apparatus) supplied the electrolyte(s) at 0.5 mL/min. In the case of iodide

salt solutions (i.e., LiI, NaI, KI, CsI), the iodide electrolyte solutions were pumped between the

nafion membrane and the cathode at 0.5 mL/min and a 1.0 M solution of the corresponding

chloride salt solutions (i.e., LiCl, NaCl, KCl, CsCl), were pumped at 0.5 mL/min between the

nafion membrane and the anode. In the case of the chloride (i.e., LiCl, NaCl, KCl, CsCl) and

hydroxide (i.e., NaOH, KOH, RbOH, CsOH) salt solutions, the nafion membrane was removed

as well as one of the 1 mm flow channels and the electrolyte was flowed directly between the

cathode and the anode. The affluent gas stream flowed directly into a gas chromatograph

(Thermo Finnegan Trace GC) operating in the thermal conductivity detection (TCD) mode,

Page 39: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

30

which uses a Carboxen 1000 column (Supelco) and a He carrier gas at a flow rate of 20 SCCM.

The column was held at 150°C and the TCD detector was held at 200°C.

2.3 Results and Discussion

2.3.1 Cation Size Effects

We first studied the effect of the cation size on the Faradaic yield of CO2 reduction and

on the partial current densities of the products, CO and H2. Specifically, we tested electrolyte

solutions with a variety of alkali cations (i.e., Li+, Na

+, K

+, Rb

+, Cs

+) for three anions (i.e., Cl

-, I

-,

Figure 2.2: Partial current densities for (a) H2, and (b) CO, as well as (c) Faradaic yield for CO in 1 M (1) chloride,

(2) hydroxide, and (3) iodide solutions, as a function of cathode potential. All conditions tested represent reactor

potentials between 2 and 3 V.

Page 40: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

31

OH-) at a concentration of 1 M in an electrochemical reactor. Figures 2a and 2b show,

respectively, the partial current densities for H2 (iH2), the unwanted product, and for CO, the

desired product, as a function of cathode potential for applied cell voltages between 2.0 and

3.0 V for chloride, iodide, and hydroxide salts. First, we discuss the effects of different salts on

hydrogen evolution (Figure 2.2a). In the case of chloride salts (Figure 2.2a1), the smallest

cation (Li+) favored H2 evolution at large potentials. For example, in the presence of LiCl, iH2

was 11 mA/cm2 at a cathode potential of -1.85 V vs. Ag/AgCl. Whereas, with electrolyte

solutions with larger cations, (i.e., NaCl, KCl, CsCl), substantially less H2 was produced across

the range of potentials tested. For example, in the presence of NaCl, KCl, or CsCl, iH2 was

6 mA/cm2 at a cathode potential of -1.85 V vs. Ag/AgCl. In the case of iodide salts (Figure

2.2a2), the influence of the cation size was less pronounced; however, we observe the same trend

that the smallest cation (Li+) and the largest cation (Cs

+) have the highest and lowest iH2,

respectively. In the case of hydroxide salts (Figure 2.2a3), rather than testing the Li+ cation,

which was found to favor H2 evolution, we tested a larger cation (Rb+), which we expected to

favor CO production. Again, we observe the same trend that the smallest cation (Na+) and the

largest cation (Cs+) have the highest and lowest iH2, respectively. In summary, for the three sets

of electrolytes tested, we observed a strong trend that an increase in cation size led to a decrease

in iH2.

We also analyzed the effect of different electrolyte compositions on CO formation

(Figure 2.2b). In the case of chloride salts (Figure 2.2b1), a maximum iCO of 26 mA was

observed in the presence of the smallest cation (Li+); however, with larger cations, much more

CO is produced at the same potentials. In particular, with the largest cation (Cs+), an iCO of

60 mA was observed at a cathode potential of -1.85 V vs. a Ag/AgCl reference electrode. In the

Page 41: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

32

case of iodide salts (Figure 2.2b2), the same trend was observed that the partial current density

of CO increased with an increase in cation size. In the case of hydroxide salts (Figure 2.2b3),

iCO increased as a function of increasing cation size for Na+, K

+, and Rb

+ whereas in the presence

of Cs+ iCO dropped relative to the iCO of Rb

+. The iCO of 175 mA/cm

2 at a cathode potential of -

2.0 V vs a Ag/AgCl reference electrode observed when using 1 M RbOH as the electrolyte is the

highest iCO reported in the literature at room temperature at this cell potential. In summary, we

observed a strong trend of larger cations increasing iCO without influencing the onset potential.

Figure 2.2c shows the Faradaic yield (selectivity) for CO as a function of cathode

potentials in the presence of chloride, iodide, and hydroxide salts. The high (or low) iCO and low

(or high) iH2 observed in the presence of large (or small) cations in the electrolyte solution,

correlate to high (or low) selectivity for conversion of CO2 to CO. In the presence of the

chloride, iodide, and hydroxide salts, the selectivity for CO was the highest for the largest cations

(e.g., Cs+, Rb

+).

Effects of cation size on product selectivity, similar to what is reported here, for the

electrochemical reduction of CO2 to methane, ethylene, methanol, and formic acid, has been

observed previously. For example, Hori et al. showed that large cations favor formic acid

production on a mercury electrode [44]. The role of ion selection for formic acid formation is

particularly interesting as this reaction has the same rate limiting step as the CO reaction, the

production of an unstable anion radical intermediate, “CO2∙ However, in this case, little .”־

explanation was given regarding the influence of cation size on selectivity of formic acid

formation over hydrogen evolution.

One would expect that the influence of cations on CO2 reduction is primarily controlled

by ion adsorption at the electrode surface. In general, smaller cations have larger hydration

Page 42: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

33

powers, and consequently, have less adsorption on electrode surfaces. Ion adsorption can

stabilize some reactions while suppressing others on an electrode surface. This, in turn,

influences product selectivity, current density, and energetic efficiency.

In our work, the silver cathode used in the reaction is operated well below its point of

zero charge (pzc), and, as a result, cation adsorption is favored on the negatively charged

electrode surface [46]. Smaller cations have larger hydration powers (i.e.,

Li+>Na

+>K

+>Rb

+>Cs

+) [47]. Consequently, the smaller a cation is, the smaller is its propensity

to adsorb on an electrode surface.

Ions adsorbed on an electrode surface influence the reaction in two primary ways. First,

ion adsorption blocks surface sites and thereby decreases the current density [48]. Second, ions

on an electrode surface modify the potential of the outer Helmholtz layer (ϕ2) and the electrode

surface [48]. Because larger cations exhibit a higher propensity for adsorption on an electrode

surface, the electrode potential will be expected to become more positive for larger cations (i.e.,

Li+<Na

+<K

+<Rb

+<Cs

+) [48]. Maznichenko et al. validated this theory by showing that the

measured overpotentials associated with the H2 reduction reaction increased with increasing

cation size (i.e., Li+<Na

+<K

+<Rb

+<Cs

+) [49]. The resulting less negative ϕ2 causes a decrease in

[H+] at the electrode surface as described by the following equation: [49]

(2.1)

where F is Faraday’s constant, R is the gas constant, and T is the absolute temperature.

Since the rate limiting step in this reaction is: , increasing [H+] at the electrode

surface (for example in the presence of smaller and thus more hydrated cations) increases the

reaction rate and decreases the overpotentials for H2 production [44]. As a result, H2 production

Page 43: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

34

increases with decreasing cation size (Li+>Na

+>K

+>Rb

+>Cs

+), as we also observed here

(Figure 2.2a).

The second way cations influence the surface reaction is by stabilizing anions on the

cathode surface [50]. For example, for the reaction of S2O8-2

to SO4-2

, large cations within the

double layer of the cathode stabilize the reactive anion and enable much higher current densities,

whereas in the discharge of H+ ions, cations within the double layer repel the reacting particle,

thereby hindering the reaction, resulting in much lower current densities [51]. Further

confirming the mechanism by which large cations stabilize anions on the surface, Frumkin et al.

observed an increase in the adsorption of anions on a cathode in the presence of large cations

(i.e., Cs+), thereby showing that ion-pairing is more pronounced with larger cations, and as a

result, drastically increases the interaction between cations and anions in the double layer [50].

While CO2 and CO are neutral products, the rate limiting step in the reaction is:

, which forms the negatively charged “CO2∙ on the cathode [7,52,53]. Cation adsorption ”־

on the cathode may stabilize “CO2∙ on the electrode surface through ion-pairing, thereby ”־

driving the reaction, improving the current density for a given overpotential. A similar trend was

observed by Ashworth et al. Similar to the situation here, they observed that cation adsorption

improved the reaction kinetics of converting a neutral species with an unstable anion

intermediate on the cathode [54]. Frumkin later proposed that the anion intermediate was

stabilized by the cations on the electrode surface, thereby enabling higher current densities from

electrolyte solutions with large cations [50].

Our systematic investigation of alkali cations on the electrochemical chemical reduction

of CO2 while using a Ag electrode indicates that the cation size plays a significant role in product

selectivity. Furthermore, the similar trends between this work and the aforementioned literature

Page 44: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

35

survey, suggest that the influence of cation size on performance stems from cation adsorption on

a cathode surface. Larger cations have a higher propensity for electrode adsorption due to their

low hydration powers, whereas smaller cations are more likely to be hydrated in solution and

consequently, less likely to adsorb on an electrode surface. Cations absorbed on a cathode repel

H+ ions from the cathode. This explains the experimentally observed trend of decreased H2

evolution when larger cations are present in the electrolyte. Additionally, cation adsorption may

stabilize “CO2∙ on the cathode surface, similar to that reported by Ashworth et al., thereby ”־

driving the reaction of CO2 to CO. This trend was experimentally confirmed in Figure 2.2b.

2.3.2 Effect of Anion Size

In this study, we also studied the effect of different anions on CO2 reduction. Figure 2.3

shows the partial current density for CO and the individual polarization curves for the

electrochemical reduction of CO2 to CO in the presence of three anions (chloride, iodide and

hydroxide) paired with a potassium cation. These anions strongly influence the polarization of

both the anode and the cathode. On the anode, we observe a drastic improvement in the oxygen

evolution reaction in the presence of a hydroxide anion (Figure 2.3b). The trend of improved

anode kinetics in the presence of hydroxide solutions also was observed in the presence of all of

Figure 2.3: (a) Partial current densities of CO, and (b) Individual electrode polarization curves for KOH, KCl, and

KI for cell potentials between 2 and 3 V.

-2.5

-2

-1.5

-1

-0.5

0

0.5

1

1.5

-150-100-500

Ele

ctro

de

Po

tent

ial

(vs.

Ag/

AgC

l)

Current Density (mA/cm2)

KOH

KCl

KI

-120

-100

-80

-60

-40

-20

0

-3.00-2.75-2.50-2.25-2.00

Cu

rre

nt

De

nsi

ty (

mA

/cm

2)

Cell Potential (V)

KOH

KCl

KI

(a) (b)

Anode

Cathode

Page 45: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

36

the other cations tested. While many factors

are at play, the predominant effect from the

addition of a hydroxide anion probably stems

from an increase in the solution pH and,

consequently, improved oxygen evolution

kinetics. On the cathode, we observed a less

pronounced, yet consistent trend.

Specifically, we see in Figure 2.3b that small

anions seemed to reduce the cathode

overpotentials at a given current density. Furthermore, as seen in Figure 2.3a, the reduced

overpotentials result in higher partial current densities for CO across the range of conditions

tested for small anions (OH-<Cl

-<I

-). This observation is in agreement with simulations which

suggest that smaller anions improve the extent of cation adsorption on a negative electrode [55].

Further confirming this explanation, Frumkin, et al. showed that a decrease in cation size

improves anion absorption on the anode. In a similar fashion, our results suggest that decreasing

the anion size improves cation absorption on a cathode, and consequently reduces the

overpotentials (Figure 2.3b) and improves the current density for CO (Figure 2.3a) [50].

2.3.3 Effect of pH

Figure 2.4 summarizes the influence of the electrolyte solution pH on the partial current

density of CO. Small changes in the partial current density for CO are seen between pH 13 and

pH 2. However, at pH 1, the partial current density drastically decreases.

Because pH effects more than just the reduction of CO2 to CO on the cathode (e.g., H2

evolution on the cathode and O2 evolution on the anode), changing the pH can significantly

Figure 2.4: Effect of pH on the partial current density of

CO production.

-60

-50

-40

-30

-20

-10

0

-2-1.5-1-0.5

Cu

rre

nt D

en

sity

(m

A/c

m2 )

Cathode Potential (vs. Ag/AgCl)

pH 0

pH 1

pH 4

pH 10

pH 13

pH 14

0

20

40

60

80

100

-3.5-3-2.5-2

Fara

da

ic E

ffe

cie

ncy

(%

)

Cell Potential (V)

pH 0 pH 1 pH 2 pH 4 pH 10 pH 13 pH 14

0

20

40

60

80

100

-3.5-3-2.5-2

Fara

dai

c Ef

feci

ency

(%

)

Cell Potential (V)

pH 0 pH 1 pH 2 pH 4 pH 10 pH 13 pH 14

Page 46: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

37

affect product selectivity. Figure 2.5 shows

the influence of the electrolyte solution pH

on the Faradaic efficiency of the reactor for

CO. Very high pH levels (i.e., pH 14)

improve the selectivity for CO production

even at low cell potentials. Also, low pH

levels decrease the selectivity for CO as

opposed to H2. Specifically, at a pH of 2, a

decrease in the Faradaic efficiency is

observed, at a pH of 1, the a further decrease in the Faradaic efficiency is observed, and, at a pH

of 0, the reactor is completely selective for H2.

The transition from high selectivity for CO in basic conditions to high selectivity for H2

in acidic conditions can be explained via individual electrode plots. Figure 2.6 shows the

cathode and anode polarization for a pH range of 0 to 14. In very basic conditions, an

improvement in anode performance is observed with little improvement on the cathode.

In acidic conditions, the cathode

overpotential is drastically reduced

because the cathode is flooded with

excess protons, which drive the

evolution of H2 as opposed to the

reduction of CO2 to CO. Furthermore,

neutral conditions experience “pH

sheltering” as the electrode surface pH

Figure 2.5: Effect of pH on the Faradaic efficiency for CO

production

0

20

40

60

80

100

-3.5-3-2.5-2

Fara

dai

c Ef

feci

en

cy (

%)

Cell Potential (V)

pH 0

pH 1

pH 2

pH 4

pH 10

pH 13

pH 14

0

20

40

60

80

100

-3.5-3-2.5-2

Fara

da

ic E

ffe

cie

ncy

(%

)

Cell Potential (V)

pH 0

pH 1

pH 2

pH 4

pH 10

pH 13

pH 14

0

20

40

60

80

100

-3.5-3-2.5-2

Fara

dai

c Ef

feci

en

cy (

%)

Cell Potential (V)

pH 0

pH 1

pH 2

pH 4

pH 10

Figure 2.6: Effect of pH on individual electrode polarization plots

for electrolytes with levels of pH ranging from 0 to 14.

-2.5-2

-1.5-1

-0.50

0.51

1.52

2.5

-150-125-100-75-50-250

Ele

ctro

de

Po

t. (

vs. A

g/A

gCl)

Current Density (mA/cm2)

pH 0

pH 1

pH 2

pH 4

pH 10

pH 13

pH 14

Page 47: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

38

will be sheltered from the electrolyte solution pH because the anode and cathode reactions

produce and consume protons [17,56,57].

2.3.4 Comparison to Literature

Figure 2.7 summarizes the data obtained in this study. Specifically the measured current

densities plotted as a function of the Faradaic and energetic efficiencies for both the data

reported here and state-of-the-art data reported earlier [15,29-31,40,58-60]. We observe a drastic

improvement in the achieved resultant energetic (~50 %) and Faradaic (~100 %) efficiencies at

high current densities. In particular, with a 1 M RbOH electrolyte solution, current densities of

175 mA/cm2 and 65 mA/cm2 were achieved at cathode potentials of -2.0 V and -1.73 V,

respectively (Figure 2.2). This data corresponds to Faradaic efficiencies of 92 % and energetic

efficiencies of 43 and 41 %, respectively, as seen in Figure 2.4. This is a drastic improvement

compared to state-of-the-art electrochemical reduction of CO2, which tend to have drastically

reduced faradaic efficiencies at increased current densities.

Figure 2.7: Comparison of the (a) Faradaic and (b) energetic efficiencies and current densities for CO2 reduction to

carbon monoxide from literature and this work.[15,29-31,40,58-60].

2.4 Conclusions

Electrolyte composition influences many factors in an electrochemical reactor, including

conductivity, and pH. Here, cation size within electrolyte solutions was observed to play a

0

20

40

60

80

100

0 50 100 150 200

Fara

dai

c Ef

feci

ency

Current Density (mA/cm2)

This Data

Literature

0

20

40

60

80

100

0 50 100 150 200

Ener

geti

c Ef

feci

en

cy

Current Density (mA/cm2)

This Data

Literature

(a) (b)

0

10

20

30

40

50

60

70

80

90

100

0 200 400 600 800 1000 1200 1400 1600

Ene

rgy

Effi

cie

ncy

Current Density (mA/cm2)

This work

Data from Literature

Hydrocarbons0

20

40

60

80

100

0 50 100 150 200

Fara

dai

c Ef

feci

ency

Current Density (mA/cm2)

This Data

Literature

0

10

20

30

40

50

60

70

80

90

100

0 200 400 600 800 1000 1200 1400 1600

Ene

rgy

Effi

cie

ncy

Current Density (mA/cm2)

This work

Data from Literature

Hydrocarbons 0

20

40

60

80

100

0 50 100 150 200

Ener

geti

c Ef

feci

en

cy

Current Density (mA/cm2)

This Data

Literature

0

10

20

30

40

50

60

70

80

90

100

0 200 400 600 800 1000 1200 1400 1600

Ene

rgy

Effi

cie

ncy

Current Density (mA/cm2)

This work

Data from Literature

Hydrocarbons

0

10

20

30

40

50

60

70

80

90

100

0 200 400 600 800 1000 1200 1400 1600

Ene

rgy

Effi

cie

ncy

Current Density (mA/cm2)

This work

Data from Literature

Hydrocarbons

0%

20%

40%

60%

80%

100%

-3.25-3.00-2.75-2.50-2.25-2.00-1.75

Fara

dai

c Ef

fici

en

cy (%

)

Cell Potential

NaOH KOH RbOH CsOH

0%

20%

40%

60%

80%

100%

-3.25-3.00-2.75-2.50-2.25-2.00-1.75

Fara

dai

c Ef

fici

en

cy (%

)

Cell Potential

NaOH KOH RbOH CsOH

Page 48: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

39

significant role in the electrochemical reduction of CO2 into CO. Specifically, small cations

(e.g., Li+, and Na

+) favor H2 evolution, and, consequently, have a low Faradaic yield for CO

production, whereas large cations (e.g., Rb+, Cs

+) favor CO production and suppress H2

evolution, resulting in a high Faradaic yield for CO, here up to 96 %. We observed a highest

partial current density of 175 mA/cm2 for CO at a cell potential of 3.0 V in the presence of

RbOH. Furthermore, we observed strong trends between anion choice and cell performance.

Anions influence both the anode and cathode polarization and consequently, the partial current

density for CO. Specifically, our data suggests that smaller anions reduce the overpotentials and

that the presence of OH- (high pH) further improves performance.

A major obstacle to the broader application of the electrochemical reduction of CO2 lies

in the challenge of simultaneously achieving high current densities and energetic efficiencies [7].

Here we demonstrate a major step towards improving both the energetic efficiencies and current

densities of the electrochemical reduction of CO2 to CO on a Ag cathode. In particular, we

observe a drastic improvement in the achieved resultant energetic (~50 %) and Faradaic (~90 %)

efficiencies at high current densities.

The work reported here suggest that further exploration of the effect of cations and anions

on the electrochemical conversion of CO2 to CO, or to other products, may lead to further

improvements in performance. In particular, further exploration of anion effects on the

electrochemical reduction of CO2 may be promising, as anions influence many parameters,

including product selectivity, cation adsorption (through ion-pairing), anode kinetics, and

conductivity.

Page 49: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

40

2.5 References

1. National Petroleum Council. Committee on Global Oil and Gas., Hard truths: Facing the

Hard Truths About Energy: A Comprehensive View to 2030 of Global Oil and Natural

Gas, National Petroleum Council, Washington, DC, 2007.

2. G. Centi and S. Perathoner, Opportunities and Prospects in the Chemical Recycling of

Carbon Dioxide to Fuels, Catalysis Today, 2009, 148 (3-4), p. 191-205.

3. K. S. Lackner, Carbonate Chemistry for Sequestering Fossil Carbon, Annual Review of

Energy and the Environment, 2002, 27 p. 193-232.

4. M. Lenzen, Current State of Development of Electricity-Generating Technologies: A

Literature Review, Energies, 2010, 3 (3), p. 462-591.

5. S. I. Plasynski, J. T. Litynski, H. G. McIlvried and R. D. Srivastava, Progress and New

Developments in Carbon Capture and Storage, Critical Reviews in Plant Sciences, 2009,

28 (3), p. 123-138.

6. U.S.EPA, U.S. EPA, 2011, pp. EPA 430-R-411-005.

7. D. T. Whipple and P. J. A. Kenis, Prospects of CO2 Utilization via Direct Heterogeneous

Electrochemical Reduction, Journal of Physical Chemistry Letters, 2010, 1 (24), p. 3451-

3458.

8. S. Bachu, CO2 Storage in Geological Media: Role, Means, Status and Barriers to

Deployment, Progress in Energy and Combustion Science, 2008, 34 (2), p. 254-273.

9. K. van Alphen, P. M. Noothout, M. P. Hekkert and W. C. Turkenburg, Evaluating the

Development of Carbon Capture and Storage Technologies in the United States,

Renewable & Sustainable Energy Reviews, 2010, 14 (3), p. 971-986.

10. K. Hara and T. Sakata, Electrocatalytic Formation of CH4 from CO2 on a Pt Gas

Diffusion Electrode, Journal of the Electrochemical Society, 1997, 144 (2), p. 539-545.

11. J.M. Duthie and H.W. Whittington, Securing Renewable Energy Supplies Through

Carbon Dioxide Storage in Methanol, Power Engineering Society Summer Meeting,

2002 IEEE, 2002, 1 p. 145-150.

12. Paul McArdle, Emissions of Greenhouse Gases in the United States 2008, 2008.

13. G. Arai, T. Harashina and I. Yasumori, Selective Electrocatalytic Reduction of Carbon-

Dioxide to Methanol on Ru-Modified Electrode, Chemistry Letters, 1989, (7), p. 1215-

1218.

14. D. P. Summers, S. Leach and K. W. Frese, The Electrochemical Reduction of Aqueous

Carbon-Dioxide to Methanol at Molybdenum Electrodes with Low Overpotentials,

Journal of Electroanalytical Chemistry, 1986, 205 (1-2), p. 219-232.

15. Y. Hori, H. Wakebe, T. Tsukamoto and O. Koga, Electrocatalytic Process of CO

Selectivity in Electrochemical Reduction of CO2 at Metal-Electrodes in Aqueous-Media,

Electrochimica Acta, 1994, 39 (11-12), p. 1833-1839.

16. M. Azuma, K. Hashimoto, M. Hiramoto, M. Watanabe and T. Sakata, Electrochemical

Reduction of Carbon-Dioxide on Various Metal-Electrodes in Low-Temperature Aqueous

KHCO3 Media, Journal of the Electrochemical Society, 1990, 137 (6), p. 1772-1778.

17. M. N. Mahmood, D. Masheder and C. J. Harty, Use of Gas-Diffusion Electrodes for

High-Rate Electrochemical Reduction of Carbon-Dioxide .1. Reduction at Lead, Indium-

Page 50: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

41

Impregnated and Tin-Impregnated Electrodes, Journal of Applied Electrochemistry,

1987, 17 (6), p. 1159-1170.

18. D. T. Whipple, E. C. Finke and P. J. A. Kenis, Microfluidic Reactor for the

Electrochemical Reduction of Carbon Dioxide: The Effect of pH, Electrochemical and

Solid-State Letters, 2010, 13 (9), p. D109-D111.

19. Y. Akahori, N. Iwanaga, Y. Kato, O. Hamamoto and M. Ishii, New Electrochemical

Process for CO2 Reduction to Formic Acid from Combustion Flue Gases,

Electrochemistry, 2004, 72 (4), p. 266-270.

20. F. Koleli, T. Atilan, N. Palamut, A. M. Gizir, R. Aydin and C. H. Hamann,

Electrochemical Reduction of CO2 at Pb- and Sn-electrodes in a Fixed-bed Reactor in

Aqueous K2CO3 and KHCO3 Media, J. of Appl. Electrochem., 2003, 33 (5), p. 447-450.

21. N. Furuya, T. Yamazaki and M. Shibata, High Performance Ru-Pd Catalysts for CO2

Reduction at Gas-Diffusion Electrodes, Journal of Electroanalytical Chemistry, 1997, 431

(1), p. 39-41.

22. S. Kaneco, R. Iwao, K. Iiba, K. Ohta and T. Mizuno, Electrochemical Conversion of

Carbon Dioxide to Formic Acid on Pb in KOH/Methanol Electrolyte at Ambient

Temperature and Pressure, Energy, 1998, 23 (12), p. 1107-1112.

23. S. Kaneco, H. Katsumata, T. Suzuki and K. Ohta, Electrochemical Reduction of CO2 to

Methane at the Cu Electrode in Methanol with Sodium Supporting Salts and its

Comparison with Other Alkaline Salts, Energy Fuels, 2006, 20 (1), p. 409-414.

24. S. Kaneco, K. Iiba, H. Katsumata, T. Suzuki and K. Ohta, Electrochemical Reduction of

High Pressure CO2 at a Cu Electrode in Cold Methanol, Electrochimica Acta, 2006, 51

(23), p. 4880-4885.

25. R. L. Cook, R. C. Macduff and A. F. Sammells, High-Rate Gas-Phase CO2 Reduction to

Ethylene and Methane Using Gas-Diffusion Electrodes, Journal of the Electrochemical

Society, 1990, 137 (2), p. 607-608.

26. S. Kaneco, H. Katsumata, T. Suzuki and K. Ohta, Electrochemical Reduction of Carbon

Dioxide to Ethylene at a Copper Electrode in Methanol Using Potassium Hydroxide and

Rubidium Hydroxide Supporting Electrolytes, Electrochimica Acta, 2006, 51 (16), p.

3316-3321.

27. T. Mizuno, K. Ohta, M. Kawamoto and A. Saji, Electrochemical Reduction of CO2 on Cu

in 0.1 M KOH-methanol, Energy Sources, 1997, 19 (3), p. 249-257.

28. J. Yano and S. Yamasaki, Pulse-mode Electrochemical Reduction of Carbon Dioxide

Using Copper and Copper Oxide Electrodes for Selective Ethylene Formation, Journal of

Applied Electrochemistry, 2008, 38 (12), p. 1721-1726.

29. N. Furuya and S. Koide, Electroreduction of Carbon-Dioxide by Metal Phthalocyanines,

Electrochimica Acta, 1991, 36 (8), p. 1309-1313.

30. T. Yamamoto, D. A. Tryk, A. Fujishima and H. Ohata, Production of Syngas Plus

Oxygen from CO2 in a Gas-Diffusion Electrode-Based Electrolytic Cell, Electrochimica

Acta, 2002, 47 (20), p. 3327-3334.

31. C. Delacourt, P. L. Ridgway, J. B. Kerr and J. Newman, Design of an Electrochemical

Cell Making Syngas (CO+H2) from CO2 and H2O Reduction at Room Temperature,

Journal of the Electrochemical Society, 2008, 155 (1), p. B42-B49.

Page 51: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

42

32. F. Bidrawn, G. Kim, G. Corre, J. T. S. Irvine, J. M. Vohs and R. J. Gorte, Efficient

Reduction of CO2 in a Solid Oxide Electrolyzer, Electrochemical and Solid-State Letters,

2008, 11 (9), p. B167-B170.

33. R. M. M. Abbaslou, J. S. S. Mohammadzadeh and A. K. Dalai, Review on Fischer-

Tropsch Synthesis in Supercritical Media, Fuel Processing Technology, 2009, 90 (7-8), p.

849-856.

34. M. E. Dry, High Quality Diesel via the Fischer-Tropsch Process - a Review, Journal of

Chemical Technology and Biotechnology, 2002, 77 (1), p. 43-50.

35. Mark E. Dry, Practical and theoretical aspects of the catalytic Fischer-Tropsch process,

Applied Catalysis A: General, 1996, 138 (2), p. 319-344.

36. S. D. Ebbesen and M. Mogensen, Electrolysis of Carbon Dioxide in Solid Oxide

Electrolysis Cells, Journal of Power Sources, 2009, 193 (1), p. 349-358.

37. S. H. Jensen, X. F. Sun, S. D. Ebbesen, R. Knibbe and M. Mogensen, Hydrogen and

Synthetic Fuel Production Using Pressurized Solid Oxide Electrolysis Cells, International

Journal of Hydrogen Energy, 2010, 35 (18), p. 9544-9549.

38. V. Kaplan, E. Wachtel, K. Gartsman, Y. Feldman and I. Lubomirsky, Conversion of CO2

to CO by Electrolysis of Molten Lithium Carbonate, Journal of the Electrochemical

Society, 2010, 157 (4), p. B552-B556.

39. Z. L. Zhan and L. Zhao, Electrochemical Reduction of CO2in Solid Oxide Electrolysis

Cells, Journal of Power Sources, 2010, 195 (21), p. 7250-7254.

40. Brian A Rosen, Amin Salehi-Khojin, Michael R Thorson, Wei Zhu, Devin T Whipple,

Paul J A Kenis and Richard I Masel, Selective Conversion of CO2 to CO at Low

Overpotentials, Science, 2011, p.

41. Emily Barton Cole, Prasad S. Lakkaraju, David M. Rampulla, Amanda J. Morris, Esta

Abelev and Andrew B. Bocarsly, Using a One-Electron Shuttle for the Multielectron

Reduction of CO2 to Methanol: Kinetic, Mechanistic, and Structural Insights, Journal of

the American Chemical Society, 2010, 132 (33), p. 11539-11551.

42. G. Seshadri, C. Lin and A. B. Bocarsly, A New Homogeneous Electrocatalyst for the

Reduction of Carbon-Dioxide to Methanol at Low Overpotential, Journal of

Electroanalytical Chemistry, 1994, 372 (1-2), p. 145-150.

43. A. Murata and Y. Hori, Product Selectivity Affected by Cationic Species in

Electrochemical Reduction of CO2 and CO at a Cu Electrode, Bulletin of the Chemical

Society of Japan, 1991, 64 (1), p. 123-127.

44. Y. Hori and S. Suzuki, Electrolytic Reduction of Carbon-Dioxide at Mercury-Electrode

in Aqueous-Solution, Bulletin of the Chemical Society of Japan, 1982, 55 (3), p. 660-665.

45. R. S. Jayashree, L. Gancs, E. R. Choban, A. Primak, D. Natarajan, L. J. Markoski and P.

J. A. Kenis, Air-breathing laminar flow-based microfluidic fuel cell, Journal of the

American Chemical Society, 2005, 127 (48), p. 16758-16759.

46. G. J. Clark, T. N. Andersen, R. S. Valentine and H. Eyring, A Comparison of the

Immersion and Open-Circuit Scrape Methods for Determining the Potential of Zero

Charge of Metal Electrodes, Journal of the Electrochemical Society, 1974, 121 (5), p.

618-622.

47. J.O.M. Bockris and A.K.N. Reddy, Modern electrochemistry, Plenum Press, 1998.

Page 52: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

43

48. P. Delahay, Double layer and electrode kinetics, Interscience Publishers, 1965.

49. E. A. Maznichenko, B. B. Damaskin and Z. A. Iofa, Effect of alkali metal cations on the

overvoltage of hydrogen on mercury in acid solutions, Dokl. Akad. Nauk SSSR, 1961,

138 p. 1377.

50. A. N. Frumkin, Influence of cation adsorption on the kinetics of electrode processes,

Transactions of the Faraday Society, 1959, 55 p. 156-167.

51. Krjukova, Dokl. Akad. Nauk SSSR, 1949, 65 p. 517.

52. R. P. S. Chaplin and A. A. Wragg, Effects of Process Conditions and Electrode Material

on Reaction Pathways for Carbon Dioxide Electroreduction with Particular Reference to

Formate Formation, Journal of Applied Electrochemistry, 2003, 33 (12), p. 1107-1123.

53. Donald A. Tryk and Akira Fujishima, Global Warming Electrochemists Enlisted in War:

The Carbon Dioxide Reduction Battle, The Electrochemical Society Interface, 2001, p.

32-36.

54. Ashworth, Coll. Czech. Chem. Comm., 1948, 13 p. 229.

55. Paul S. Crozier, Richard L. Rowley and Douglas Henderson, Molecular-dynamics

simulations of ion size effects on the fluid structure of aqueous electrolyte systems

between charged model electrodes, Journal of Chemical Physics, 2001, 114 (17), p.

7513-7517.

56. D. T. Whipple, E. C. Finke and P. J. A. Kenis, Microfluidic Reactor for the

Electrochemical Reduction of Carbon Dioxide: The Effect of pH, Electrochemical and

Solid State Letters, 2010, 13 (9), p. D109-D111.

57. H. Li and C. Oloman, Development of a Continuous Reactor for the Electro-reduction of

Carbon Dioxide to Formate - Part 2: Scale-up, Journal of Applied Electrochemistry,

2007, 37 (10), p. 1107-1117.

58. K. W. Frese and S. Leach, Electrochemical Reduction of Carbon-Dioxide to Methane,

Methanol, and CO on Ru Electrodes, Journal of the Electrochemical Society, 1985, 132

(1), p. 259-260.

59. T. Saeki, K. Hashimoto, A. Fujishima, N. Kimura and K. Omata, Electrochemical

Reduction of CO2 with High-Current Density in a CO2-Methanol Medium, Journal of

Physical Chemistry, 1995, 99 (20), p. 8440-8446.

60. Eric Dufek, Tedd Lister and Michael McIlwain, Bench-scale electrochemical system for

generation of CO and syn-gas, Journal of Applied Electrochemistry, 2011, 41 (6), p. 623-

631.

Page 53: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

44

Chapter 3

Amine Based Electrolytes for the Electrochemical Reduction of CO21

3.1 Introduction

In Chapter 2, we showed that electrolyte solutions can drastically effect the

electrochemical reduction of CO2 both in terms of current density and faradaic efficiency.

Specifically, we identified that cation size, anion size, and pH play key roles in the product

distribution. Similarly, in the case of many other reactions, cation and anion selection has shown

to drastically effect the reaction kinetics and products [1-7]. We are interested in using

electrolyte solutions to further improve the selectivity, and furthermore, reduce the

overpotentials for the reaction.

1 Part of this work has been published: Rosen, B. A., Salehi-Khojin, A., Thorson, M. R., Zhu, W., Whipple, D. T.;

Kenis, P. J. A., Masel, R.I., “Ionic Liquid–Mediated Selective Conversion of CO2 to CO at Low Overpotentials.”

Science 2011, 334, 643-644. http://www.sciencemag.org/content/334/6056/643

Figure 3.1: Schematic of the proposed mechanism by which EMIM BF4 may stabilize the CO2-

intermediate on the electrode surface.

Page 54: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

45

Bockris, etc. showed that the overpotential for CO2 conversion is quite large due to the

high potential to form a CO2 radical intermediate [8-10]. Specifically, the overpotential for the

formation of the “CO2 intermediate” is -1.9 V vs. RHE. If a complex can be formed between a

“co-catalyst” species (as proposed in Figure 3.1) and the intermediate on the catalyst surface, the

potential required to form the intermediate could be reduced (Figure 3.2), thereby reducing the

potential for the overall reaction.

In this study, we explore the use of an ionic liquid, 1-ethyl-3-methylimidazolium

tetrafluoroborate, EMIM BF4, to stabilize the intermediate and thereby reduce the overpotential

necessary for the reaction. Ionic liquids are particularly interesting as they have the advantages

of having relatively high conductivities due to their ionic nature, high solubilities for CO2, and

“tune-ability” due to the ability to change the chemical structure or the cationic/ anionic species

[11,12]. The high solubility for CO2 inherent to many ionic liquids makes them potentially

Figure 3.2: . A schematic of how the free energy of the system changes during the reaction CO2 + 2H

+ +

2e– CO + H2O in water or acetonitrile (solid line) or EMIM-BF4 (dashed line). From Rosen, B. A.,

Salehi-Khojin, A., Thorson, M. R., Zhu, W., Whipple, D. T.; Kenis, P. J. A., Masel, R.I., “Ionic Liquid–

Mediated Selective Conversion of CO2 to CO at Low Overpotentials.” Science 2011, 334, 643-

644.Reprinted with permission from AAAS [2].

Intermediate

stabilized

Page 55: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

46

excellent electrolytes because (a) they could be used for both CO2 capture and CO2 reduction and

(b) they may improve CO2 mass transport by increasing the local concentration of CO2 near the

cathode of an electrochemical [13-15].

This study plans to build on work done by our collaborators at Dioxide Materials, who

have identified that EMIM BF4 reduces the overpotential necessary for CO2 reduction to CO by

using a three electrode reactor. Here, in collaboration with Brian Rosen and Dr. Richard Masel

at Dioxide Materials, I build on their results of an early onset potential for CO2 reduction to CO

in an attempt to show a reduced overall reactor potential with this ionic liquid, EMIM BF4, in a

flowing electrochemical reactor. The electrochemical flow reactor discussed in-detail in

Chapter 2 was chosen for this study because (a) the small band gap between the cathode and

anode for the electrolyte stream makes for small ohmic losses, and (b) the flowing electrolyte

stream enables in-situ electrode characterization of both the anode and cathode electrode

polarization.

3.2 Experimental

3.2.1 Reactor Testing

The reactor operational procedure was identical to that used in Chapter 2. The following

electrolyte solutions were used in these studies: 18 mole% EMIM BF4 in 18.2 MΩ∙cm H2O,

1 M KCl in H2O, and 0.1 M H2SO4 in H2O. All of the experiments in this study used 5 mg/cm2

Ag on Sigracet 35BC cathodes. All of the anodes used in this study had a 2 mg/cm2 Pt loading

on Sigracet 35BC unless otherwise specified.

Page 56: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

47

3.2.2. Reference Electrodes

Cathode potentials for all EMIM BF4 electrolyte solutions were measured relative to a

non-aqueous Ag/Ag+ reference electrode (BASi, MF-2062). The non-aqueous reference

electrode was prepared with an internal solution consisting of: 0.1 M tetrabutlyammonium

perchlorate (TBAP, Sigma Aldrich) and 0.01 M AgNO3 (BASi) in acetonitrile (Sigma Aldrich).

The Ag/Ag+ reference electrode was calibrated relative to the potential of the

ferrocene/ferrocenium couple in a 2.25 mM ferrocene (Sigma Aldrich), 18 mole% EMIM BF4

aqueous solution, and a 2.25 mM ferrocene in acetonitrile solution. Figure 3.3 shows anodic and

cathodic waves for ferrocene using a Ag/Ag+ reference electrode in acetonitrile and

18 mole% EMIM BF4 solutions. The potential difference between the ferrocene/ferrocenium

couple relative and the Ag/Ag+ reference electrode was +100 mV in the acetonitrile solution and

-23 mV in the EMIM BF4 solution. The shift in potential is believed to be due to differences in

the junction between the reference electrode and the acetontile or EMIM BF4 solutions. Based

on this data, we conclude that the Ag/Ag+ acetonitrile reference electrode will read a potential

Figure 3.3: Anodic and cathodic waves for Ferrocene using a Ag/Ag+ reference electrode in acetonitrile

and 18% EMIM BF4 solutions.

-100

-50

0

50

100

150

200

-0.6 -0.4 -0.2 0 0.2 0.4 0.6

Cu

rre

nt

(uA

)

Potential (vs Ag/Ag+)

Acetonitrile

EMIM BF4

Page 57: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

48

which is 123 mV higher in the acetonitrile solution as compared to the EMIM BF4 solution.

Pavlishchuk and Addison report that the Ag/Ag+ reference electrode used in this study is 537 mV

off of SHE [16]. Consequently, the Ag/Ag+ reference electrode is 631 mV (537+123) below

SHE.

For all individual electrode potential measurements relative to an aqueous solution, a

Ag/AgCl reference electrode (BASi, RE-5B) was used in the aqueous stream.

3.2.3 Cyclic Voltammetry

Cyclic voltammetry was carried out in-situ within the electrochemical reactor relative to a

Ag/AgCl or a Ag/Ag+ reference electrode for aqueous or non-aqueous electrolyte solutions,

respectively. A reference electrode was placed upstream in the electrolyte solution and the

cathode potential was controlled relative to reference electrode at a scan rate of 20 mV/s. For all

of the tests, either N2 (an inert gas) or CO2, a reactant gas was flowed over the cathode chamber

at a flow rate of 5 SCCM. The electrolytes were pumped at 0.5 mL/min using a syringe pump

(Harvard Apparatus, PHD 2200).

3.2.4 Electrochemical impedance spectroscopy

Electrochemical impedance spectroscopy (EIS) was used to determine the internal reactor

resistance of the electrochemical reactor used in these studies with the various combinations of

electrolytes and membranes. The impedance spectra were obtained using a Frequency Response

Analyzer (FRA) module controlled by the potentiostat (Autolab PGSTAT-30, EcoChemie). EIS

was done at a potential of 3V to ensure that the reaction was selective for the desired product,

carbon monoxide.

Page 58: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

49

3.3 Results and Discussion

3.3.1 Cyclical Voltammetry

Figure 3.4 shows in-situ cyclic

voltammograms (CV) using the single

channel electrochemical reactor described in

Chapter 2 [17] with a Ag cathode with either

a 1 M KCl aqueous electrolyte or a

18 mole% EMIM BF4 in H2O electrolyte

flowed at 0.5 mL/min and with either N2 or

CO2 flowed behind the cathode at 5 SCCM. Because N2 is an inert gas at the potentials applied,

when flowing N2 over the cathode, the only potential product is H2 from water electrolysis [18].

The results obtained when using an ionic liquid electrolyte, EMIM BF4, were compared to results

obtained when using a more traditional electrolyte stream, 1 M aqueous KCl [17,19-21]. In

Figure 3.4, as the cathode potential was scanned in the more negative direction, an increase in

the cathode current density was observed for all four conditions tested, corresponding to an

increase in reduction current. When N2 is flowed over the cathode, for both the EMIM BF4 and

KCl solutions, the observed current is assumed to go towards the evolution of H2 as no CO2 is

available for reduction, and N2 is inert at the potentials applied here [18]. An increase in current

with the substitution of CO2 for N2 suggests that the addition of CO2 results in the reduction of

CO2 to carbon fuels (i.e., CO, CH4, CH3OH, CO2H2, or C2H4). Specifically, the substitution of

CO2 for N2 with both the aqueous and ionic liquid solutions resulted in a considerable increase in

current density, thereby suggesting that the main products in the presence of CO2 are from the

reduction of CO2. Furthermore, the onset potential (as defined as the potential at which a current

Figure 3.4: Cyclic voltammograms at a scan rate of

20 mV/s over a Ag cathode with EMIM BF4 or

aqueous catholytes and a N2 or CO2 gas stream

flowing over the cathode.

-50

-45

-40

-35

-30

-25

-20

-15

-10

-5

0

-1.75 -1.5 -1.25 -1 -0.75 -0.5

Cu

rre

nt d

en

sity

(m

A/c

m2 )

Potential (vs. Ag/AgCl)

1M KCl with N2

1M KCl with CO2

18% EMIM BF4 with N2

18% EMIM BF4 with CO2

Page 59: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

50

density of 5 mA/cm2) with the CO2

gas stream is decreased in the case of

the ionic liquid electrolyte solution (-

1.08 V) as compared to the aqueous

solution (-1.246 V). However, the

EMIM BF4 solution exhibited

hysteresis on the return scan, thereby

suggesting that the system may be

mass transfer limited. The decreasing

slope beyond -1.35 for the EMIM BF4

solution also suggests that the system may become mass transport limited at higher current

densities, which is expected due to the poor viscosity of the EMIM BF4 solution. In the presence

of the EMIM BF4 solution, the rate of hydrogen evolution compared to that in the aqueous

solution also increases, but to a lesser extent than the CO2 reduction. Because different reference

electrodes were used for the ionic liquid solution and the aqueous solution, the early onset

potential needed to be confirmed via the total applied potential in a flow cell.

3.3.2 Individal Electrode Potentials

Using the single electrolyte channel reactor with a Ag cathode and a Pt anode and a CO2

gas stream (5 SCCM), the individual electrode performance of the reactor with a 18 mole%

EMIM BF4 electrolyte solution and a 1M KCl aqueous solutions, both flowed at 0.5 ml/min, are

compared in Figure 3.5. Relative to the aqueous stream, the ionic liquid solution exhibits poor

anode performance, which is hypothesized to be due to the ionic liquid, EMIM BF4, poisoning

the anode. Regarding the cathode performance, the overpotentials at current densities appear to

Figure 3.5: Individual electrode potentials of 18% EMIM

BF4 and 1M KCl for a Ag cathode and a Pt Anode in a single

electrolyte channel reactor

Page 60: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

51

be lower when using the EMIM BF4 electrolyte as compared to when using the aqueous KCl

electrolyte, however, a greater overpotential is required at higher current densities. The

increased cathode overpotentials confirms that the reactor may be mass transport limitations. In

further support of this theory, the diffusivity of CO2 in EMIM BF4 is ~10 times worse than the

diffusivity of CO2 in an aqueous liquid [2]. To independently address and study the influence of

the electrolyte on the cathode and anode, we developed a dual electrolyte reactor.

3.3.3 Dual Electrolyte Reactor

The dual electrolyte reactor, as shown in Figure 3.6 and originally discussed in Chapter 2,

was originally developed to enable the study of the influence of EMIM BF4 on the anode and

cathode performance. This design is similar to the single electrolyte channel electrochemical

reactor used previously [17,19]. The reactor has a separate anode and cathode GDEs backed by

aluminum current collectors. The anode is exposed to the atmosphere to enable the O2 to escape

the reactor. A gas channel above the cathode enables both the feed of CO2 and the escape of CO

through the vapor phase. Two electrolytes, an anolyte and a catholyte, flow over the anode and

cathode, respectively. The electrolyte streams are separated via a membrane. The dual

Figure 3.6: Schematic diagram of the dual electrolyte microfluidic reactor for CO2 conversion.

Page 61: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

52

electrolyte streams enable separate

control of conditions at each electrode

and can thereby prevent the ionic liquid

from poisoning the anode via physical

separation. Reference electrodes are

placed in each of the outlet streams

(Ag/Ag+ in the ionic liquid and

Ag/AgCl in the aqueous stream). The

reference electrodes enable

measurement of the potentials of the

individual electrodes as well as the

junction potential between the two electrolytes (the difference between the reference electrodes

at open circuit).

3.3.4 Membranes

Membrane selection is critical for maintaining ionic contact between the two electrolyte

streams (thereby reducing reactor resistance), preventing membrane poisoning, limiting the

junction potential and preventing mixing of the electrolyte streams.

Here, several potential membranes were explored, a cation exchange membrane, an anion

exchange membrane, and a porous separator. The cation exchange membrane, Nafion (Dupont),

has the advantage of drastically reducing fuel crossover while allowing for the transfer of H+

cations in a low resistance environment, and is consequently the most commonly used fuel cell

membrane. Anionic membranes (i.e., Acta Membrane) allow the transfer of an anion, while

limiting cation crossover. Lastly, Hollinger et al. identified that filter paper (Track-Etch

Figure 3.7: Individual electrode potentials of EMIM BF4

and KCl for in a single electrode reactor and a combination

of EMIM BF4 and KCl in a dual electrolyte reactor with

Nafion or tracketch membranes.

Page 62: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

53

membranes) can be used in-lieu of a

more traditional membrane to reduce

crossover of unwanted species through

the reduction of the effective cross-

sectional area [22]. Figure 3.7 shows

overlays the individual cathode (as

measured via Ag/Ag+ reference

electrode in catholyte) and anode (as

measured via Ag/AgCl reference

electrode in anolyte) polarization

curves for the aforementioned three

membranes (Nafion, Tracketck, and

Acta membrane) in the dual electrolyte reactor (18 % EMIM BF4 at the Ag cathode and 1M

H2SO4 at the Pt Anode) as compared to a single channel reactor with a 18 mole% EMIM BF4

electrolyte and a 1 M KCl aqueous electrolyte. The anode polarization curves show that (1) a

Nafion membrane prevents an EMIM BF4 solution from poisoning the anode, (2) a Track-Etched

membrane reduced, but didn’t eliminate the poisoning on the anode, and (3) the anion exchange

membrane (Acta), reduces the poisoning but has less of an effect at high current densities, where

the anion crossover rate is high. The effectiveness of the cation exchange membrane (which

prevents anion diffusion across the membrane) and the ineffectiveness of the anion exchange

membrane (which allows anion diffusion across the membrane) at high current densities, where

the anion crossover rate is high, suggests that the anion, BF4-, is responsible for the anode

poisoning.

Figure 3.8: Nyquist plots from EIS of 3 different reactor

electrolyte arrangements. Aq. is a 1 M KCl aqueous

solution, IL is an 18 mole% EMIM BF4 solution, and

IL_N_Aq is an 18 mole% EMIM BF4 catholyte and a 1 M

KCl anolyte separated by a Nafion membrane

0.00

2.00

4.00

6.00

8.00

10.00

0 5 10 15-l

m Z

-cm

2)

Re Z (Ω-cm2)

Aq.

IL

IL_N_Aq.

Page 63: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

54

While a Nafion membrane exhibits superior behavior in terms of preventing anode

poisoning, issues with membrane poisoning from ionic liquids have previously been reported

[23]. To study the EMIM BF4 / Nafion interaction, impedance was carried out with and without

the presence of the membrane. Nyquist plots of the EIS data from the reactor with various

electrolytes and membranes are shown in Figure 3.8. The distance between the vertical axis and

the high frequency points (the points closest to the origin) represents the ohmic resistance (RΩ)

or reactor resistance (Rreactor) of the system and is composed of the electrolyte resistance and any

resistance in the components of the reactor.

The membrane resistance can be estimated from the Nyquist plots by measuring the

resistance with the membrane and subtracting the resistance without the membrane. From

Figure 3.8, we identify the resistance of the reactor with and without the membrane.

Specifically, with a single channel reactor using an 18 mole% EMIM BF4 electrolyte or a 1 M

KCl aqueous electrolyte, the reactor resistance is 3.41 or 2.65 Ω cm2, respectably. Based on

these resistances, the combined resistance of an arrangement with a 212 Nafion membrane

separating an 18 mole% EMIM BF4 electrolyte and a 1 M KCl electrolyte (with no poisoning or

no junction potential) could be approximated by the mathematical addition of these resistances,

6.06 Ω cm2. However, the Nyquist plots shows that for the aforementioned

electrolyte/membrane arrangement, the total reactor resistance is 12.73 Ω cm2. In conclusion,

the ionic liquid, EMIM BF4, appears to be swelling or blocking sites in the Nafion membrane

and thereby increasing the reactor resistance.

Additionally, we measured the reactor resistance in the single channel reactor with only

18 % EMIM BF4 and with only 1 M KCl. The reactor resistances in these two cases were

expected to be quite similar as both solutions have very similar conductivities. However, with

Page 64: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

55

the EMIM BF4 electrolyte solution in the reactor, an additional resistance of 0.85 Ω cm2 was

recorded, which suggests that the EMIM BF4 electrolyte solution is interfering with one of the

electrodes, and thereby increasing the reactor resistance. We hypothesize that the additional

reactor resistance originates from EMIM BF4 swelling the Nafion binder used to hold the catalyst

metal (Ag or Pt) to the GDE (Sigracet 35BC) because a similar phenomena was observed by

Schmidt, et al. Specifically, he observed that ionic liquids can swell Nafion [23]. With the

EMIM electrolyte, the reactor is resistance limited.

3.3.5 Early Onset Potenial in Operational Reactor

Figure 3.9 shows the faradaic efficiency for CO and H2, total energetic efficiency, and

reaction turnover number between reactor potentials of 1.5 and 2.5 V reported by my

collaborator, Brian Rosen, at Dioxide Materials when using a EMIM BF4 – Nafion – H2SO4 dual

electrolyte setup [2]. The onset of CO production is achieved at a total reactor potential of only

Figure 3.9: Performance for EMIM BF4 – Nafion – H2SO4 setup a dual electrolyte setup in terms of

faradaic efficiency for CO, and H2, total energetic efficiency, and reaction turnover number. From

Rosen, B. A., Salehi-Khojin, A., Thorson, M. R., Zhu, W., Whipple, D. T.; Kenis, P. J. A., Masel, R.I.,

“Ionic Liquid–Mediated Selective Conversion of CO2 to CO at Low Overpotentials.” Science 2011, 334,

643-644.Reprinted with permission from AAAS [2].

Page 65: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

56

1.5 V in the presence of an

18 % EMIM BF4 catholyte

This low cell potential,

1.5 V, constitutes a drastic

reduction in the onset

potential for CO production

(0.17 V overpotential),

which theoretically occurs at

1.33 V [6,7,19].

Furthermore, the results

show that excellent selectivity for CO can be achieved between reactor potentials of 1.5 and

2.5 V.

3.3.6 Low Current Densities

Figure 3.10 shows the current densities achieved with a single electrolyte channel reactor

with a 1 M KCl aqueous electrolyte solution compared to the current density achieved with a

18 %EMIM BF4 –Nafion – 1M H2SO4 electrolyte setup in the dual electrolyte reactor. While

EMIM BF4 drastically reduces the onset potential, the current densities are too low in the present

setup for immediate application. Specifically, drastic reductions in the cell resistance (or

membrane poisoning) are necessary to capitalize on the low onset potentials observed in the

presence of EMIM BF4.

3.4 Conclusions

EMIM BF4 solutions show tremendous potential for reducing the required overpotentials

for reaction. Specifically, the cyclical voltammetry confirms that earlier onset potential are

Figure 3.10:Current density output for an electrochemical cell with an

18% EMIM BF4 solution as compared to an electrochemical cell with a

1 M KCl electrolyte solution.

-120

-100

-80

-60

-40

-20

0

-3.25-3.00-2.75-2.50-2.25-2.00-1.75

Cu

rre

nt D

en

sity

(mA

/cm

2)

Cell Potential

EMIM BF4

KCl

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

-4.0-3.5-3.0-2.5-2.0-1.5

Cu

rre

nt D

en

sity

(mA

/cm

2)

Cell Potential (V)

Page 66: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

57

possible in the flowing reactors. However, both the CV and single electrode plots suggest that

mass transport may hinder the reaction rate on the electrode surface.

The individual electrode polarization plots clearly show that EMIM BF4 poisons the Pt

anode. However, the development of the dual electrolyte reactor, with a Nafion cation-exchange

membrane, enables a drastic reduction in the extent of cathode poisoning observed. Because the

poisoning was only eliminated via the use of a cation exchange membrane eliminates, we believe

the BF4 group is responsible for the anode poisoning. While a Nafion membrane helps to

overcome anode poisoning, the Nafion membrane appears to be poisoned by an EMIM BF4

solution as is evident by an increased reactor resistance with the addition of the membrane.

In spite of the high reactor resistances, the dual electrolyte reactor enabled the

identification of the onset potential for CO. The onset of CO production is achieved at a total

reactor potential of only 1.5 V in presence of 18 % EMIM BF4, which constitutes a total reactor

overpotential of 0.17 V. The low onset potential observed here constitutes a drastic reduction in

the onset potential for CO production in an arrangement that achieves excellent selectivity for

CO. The reactor developed in this study served the purpose of analyzing the overall reactor

onset potential, however, practical applications require much higher current densities, which

result in large overpotentials in the present setup.

3.5 References

1. S. Kaneco, H. Katsumata, T. Suzuki and K. Ohta, Electrochemical Reduction of CO2 to

Methane at the Cu Electrode in Methanol with Sodium Supporting Salts and its

Comparison with Other Alkaline Salts, Energy Fuels, 2006, 20 (1), p. 409-414.

2. Brian A. Rosen, Amin Salehi-Khojin, Michael R. Thorson, Wei Zhu, Devin T. Whipple,

Paul J. A. Kenis and Richard I. Masel, Ionic Liquid–Mediated Selective Conversion of

CO2 to CO at Low Overpotentials, Science, 2011, 334 (6056), p. 643-644.

3. E. B. Cole, P. S. Lakkaraju, D. M. Rampulla, A. J. Morris, E. Abelev and A. B. Bocarsly,

Using a One-Electron Shuttle for the Multielectron Reduction of CO2 to Methanol:

Kinetic, Mechanistic, and Structural Insights, Journal of the American Chemical Society,

2010, 132 (33), p. 11539-11551.

Page 67: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

58

4. A. J. Morris, R. T. McGibbon and A. B. Bocarsly, Electrocatalytic Carbon Dioxide

Activation: The Rate-Determining Step of Pyridinium-Catalyzed CO2 Reduction,

Chemsuschem, 2011, 4 (2), p. 191-196.

5. G. Seshadri, C. Lin and A. B. Bocarsly, A New Homogeneous Electrocatalyst for the

Reduction of Carbon-Dioxide to Methanol at Low Overpotential, Journal of

Electroanalytical Chemistry, 1994, 372 (1-2), p. 145-150.

6. Y. Hori, Electrochemical CO2Reduction on Metal Electrodes, Modern Aspects of

Electrochemistry, 2008, 42 p. 89-189.

7. Yoshio Hori, Hidetoshi Wakebe, Toshio Tsukamoto and Osamu Koga, Electrocatalytic

process of CO selectivity in electrochemical reduction of CO2at metal electrodes in

aqueous media, Electrochimica Acta, 1994, 39 (11–12), p. 1833-1839.

8. J. O. Bockris and J. C. Wass, The Photocatalytic Reduction of Carbon-Dioxide, Journal

of the Electrochemical Society, 1989, 136 (9), p. 2521-2528.

9. J.O.M. Bockris and A.K.N. Reddy, Modern electrochemistry, Plenum Press, 1998.

10. K. Chandrasekaran and L. O'M Bockris, In-situ spectroscopic investigation of adsorbed

intermediate radicals in electrochemical reactions: CO2− on platinum, Surface Science,

1987, 185 (3), p. 495-514.

11. J. E. Bara, T. K. Carlisle, C. J. Gabriel, D. Camper, A. Finotello, D. L. Gin and R. D.

Noble, Guide to CO2Separations in Imidazolium-Based Room-Temperature Ionic

Liquids, Industrial & Engineering Chemistry Research, 2009, 48 (6), p. 2739-2751.

12. S. J. Zhang, Y. H. Chen, F. W. Li, X. M. Lu, W. B. Dai and R. Mori, Fixation and

Conversion of CO2Using Ionic Liquids, Catalysis Today, 2006, 115 (1-4), p. 61-69.

13. A. Yokozeki, M. B. Shiflett, C. P. Junk, L. M. Grieco and T. Foo, Physical and Chemical

Absorptions of Carbon Dioxide in Room-Temperature Ionic Liquids, Journal of Physical

Chemistry B, 2008, 112 (51), p. 16654-16663.

14. L. E. Barrosse-Antle and R. G. Compton, Reduction of Carbon Dioxide in 1-butyl-3-

methylimidazolium Acetate, Chemical Communications, 2009, (25), p. 3744-3746.

15. S. S. Moganty and R. E. Baltus, Diffusivity of Carbon Dioxide in Room-Temperature

Ionic Liquids, Industrial & Engineering Chemistry Research, 2010, 49 (19), p. 9370-

9376.

16. Vitaly V. Pavlishchuk and Anthony W. Addison, Conversion constants for redox

potentials measured versus different reference electrodes in acetonitrile solutions at

25°C, Inorganica Chimica Acta, 2000, 298 (1), p. 97-102.

17. D. T. Whipple, E. C. Finke and P. J. A. Kenis, Microfluidic Reactor for the

Electrochemical Reduction of Carbon Dioxide: The Effect of pH, Electrochemical and

Solid State Letters, 2010, 13 (9), p. D109-D111.

18. H. De Jesus-Cardona, C. del Moral and C. R. Cabrera, Voltammetric Study of

CO2Reduction at Cu electrodes Under Different KHCO3 Concentrations, Temperatures

and CO2 Pressures, Journal of Electroanalytical Chemistry, 2001, 513 (1), p. 45-51.

19. D. T. Whipple and P. J. A. Kenis, Prospects of CO2 Utilization via Direct Heterogeneous

Electrochemical Reduction, Journal of Physical Chemistry Letters, 2010, 1 (24), p. 3451-

3458.

Page 68: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

59

20. S. Kaneco, K. Iiba, H. Katsumata, T. Suzuki and K. Ohta, Electrochemical Reduction of

High Pressure CO2at a Cu Electrode in Cold Methanol, Electrochimica Acta, 2006, 51

(23), p. 4880-4885.

21. Y. Hori, A. Murata and R. Takahashi, Formation of Hydrocarbons in the

Electrochemical Reduction of Carbon-Dioxide at a Copper Electrode in Aqueous-

Solution, Journal of the Chemical Society, Faraday Transactions 1, 1989, 85 p. 2309-

2326.

22. A. S. Hollinger, R. J. Maloney, R. S. Jayashree, D. Natarajan, L. J. Markoski and P. J. A.

Kenis, Nanoporous separator and low fuel concentration to minimize crossover in direct

methanol laminar flow fuel cells, Journal of Power Sources, 2010, 195 (11), p. 3523-

3528.

23. C. Schmidt, T. Glück and G. Schmidt-Naake, Modification of Nafion Membranes by

Impregnation with Ionic Liquids, Chemical Engineering & Technology, 2008, 31 (1), p.

13-22.

Page 69: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

60

Chapter 4

Increased Current Densities for Amine-Based EMIM+ Electrolytes

4.1 Introduction

As discussed in Chapter 3, one of the biggest challenges for the electrochemical reduction

of CO2 to CO lies in achieving high energetic efficiencies [1-4] at reasonable current densities

[5,6]. Towards achieving this goal, several electrolyte solutions have shown much promise [4,7-

13]. Specifically, pyridine has been shown to change product selectivity [7-9], other electrolyte

solutions have been shown to improve partial current densities for the desired products

[10,14,15] and EMIM BF4 has been shown to reduce the onset potential for CO production [4].

While EMIM BF4 drastically reduced the onset potential for CO production, the current

densities achieved with EMIM BF4 are quite low in comparison to more traditional electrolyte

solutions due to the large reactor resistance with EMIM BF4 in the dual electrolyte reactor [4].

As discussed in Chapter 3, the primary reason for the low current densities lies in the high

reactor resistances stemming from membrane poisoning.

This study explores substituting smaller anions (i.e., OH- and Cl

-) in place of BF4

- on an

EMIM+ cation. We expect this substitution to (a) address anode poisoning without the use of a

dual electrolyte reactor and (2) improve both reaction selectivity (for CO) and current density

based on correlations identified in Chapter 2 between decreased anion size and improved

performance. Furthermore, we study the effect of bolstering the electrolyte conductivity with

additional salts to improve reaction rates, as well as the effect of decreasing the pH via the

addition of hydroxide to improve anode kinetics [16].

Page 70: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

61

4.2 Experimental

4.2.1 Reactor Testing

This study used the same equipment and procedures discussed in Chapter 2. For all of

the experiments carried out in this study, a 1 mg/cm2 Ag cathode loading and a 6.67 mg/cm

2 Pt

black anode loading was used. All gas flow rates were 5 SCCM and all liquid flow rates were

0.5 mL/min. All product analysis was carried out with the Thermo GC discussed in Chapter 2.

4.2.2 Ionic Liquid Solution Preparation

EMIM BF4 (Product Number:711721; Provider: Sigma Aldrich), EMIM Cl (30764;

Sigma Aldrich), KCl (P1597; Sigma Aldrich) and KOH (44016; Sigma Aldrich) were obtained

and used as provided unless otherwise specified. EMIM OH was produced via ion exchange.

Specifically, 26 g of EMIM BF4 was mixed with 8 g of KOH and 20 mL of H2O. KBF4 formed

a gel immediately upon mixing. The solution was mixed for 30 minutes. The mother liquor was

separated from the gel via vacuum filtration (2x) using 10 µm pore-size filter paper.

4.3 Results and Discussion

4.3.1 Propensity for Precipitation

Table 4.1 shows the results from the addition of 1 mL of various ionic liquid solutions

with various 1 mL salt solutions. Both the BMIM+ BF4

- and EMIM

+ BF4

- ionic liquid solutions

precipitated with most additive solutions. The precipitant solutions were analyzed with

elemental analysis for both the KCl / EMIM BF4 and NaCl / EMIM BF4 solutions and the

precipitant was found to contain only trace amounts of C, H or N, thereby confirming that the

precipitates were salts of the BF4 anion. While many BMIM+ PF4

- solutions did not precipitate

with many solutions, the majority of these solutions phase separated. However, the EMIM+ Cl

-

Page 71: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

62

ionic liquid solution was stable with the addition of most salts. This suggests that the EMIM+ Cl

-

ionic liquid solution’s conductivity could be bolstered with the addition of other electrolytes.

Table 4.1: Solid form data from the addition of salt “additive”

solution to various ionic liquid solutions

Additives 18 mole% Ionic Liquid Solutions

EMIM Cl EMIM BF4 BMIM BF4 BMIM PF4

KOH -a Gel Crystal -

NaOH - Crystal Crystal -

Choline

Ace. - - - -

Choline Cl - - - -

Choline OH - - - Gel

NaCl - Crystal Crystal -

CsCl - Gel Gel -

CaCl2 Crystal Gel Gel -

MgSO4 - - - -

NaHCO3 - Crystal Crystal -

KHCO3 - Crystal Crystal -

K2CO3 - Crystal Crystal Gel

NH4HCO3 - - - -

CsOH - Gel Gel - a “-“ indicates that no precipitate was observed upon the

addition of an additive to an ionic liquid solution

4.3.2 Individual Electrode Characterization of EMIM+ Cl

- Solution

A dual electrolyte reactor developed

for the EMIM BF4 electrolyte solution

overcame anode poisoning from the addition

of the ionic liquid solution as discussed in

Chapter 3, however, at the expense of low

current densities due to high membrane

resistances. Because the source of the anode

poisoning (from the BF4- or the EMIM

+ ion)

was not previously confirmed, we analyzed

Figure 4.1: Individual electrode polarization plots for a

single channel reactor with either an EMIM Cl solution or

an aqueous KCl solution and for a dual electrolyte reactor

with an EMIM BF4 solution.

-2

-1.5

-1

-0.5

0

0.5

1

1.5

2

-150-100-500

Ele

ctro

de

Po

ten

tial

(V v

s A

g/A

gCl)

Current Density (mA/cm2)

EMIM BF4

EMIM Cl

KCl

-2

-1.5

-1

-0.5

0

0.5

1

1.5

2

-150-100-500

Ele

ctro

de

Po

ten

tial

(V v

s A

g/A

gCl)

Current Density (mA/cm2)

EMIM BF4

EMIM Cl

KCl

-2

-1.5

-1

-0.5

0

0.5

1

1.5

2

-150-100-500

Ele

ctro

de

Po

ten

tial

(V v

s A

g/A

gCl)

Current Density (mA/cm2)

EMIM BF4

EMIM Cl

KCl

-2

-1.5

-1

-0.5

0

0.5

1

1.5

2

-150-100-500

Ele

ctro

de

Po

ten

tial

(V v

s A

g/A

gCl)

Current Density (mA/cm2)

EMIM BF4

EMIM Cl

KCl

-2

-1.5

-1

-0.5

0

0.5

1

1.5

2

-150-100-500

Ele

ctro

de

Po

ten

tial

(V v

s A

g/A

gCl)

Current Density (mA/cm2)

EMIM BF4

EMIM Cl

KCl

-2

-1.5

-1

-0.5

0

0.5

1

1.5

2

-150-100-500

Ele

ctro

de

Po

ten

tial

(V v

s A

g/A

gCl)

Current Density (mA/cm2)

EMIM BF4

EMIM Cl

KCl

Page 72: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

63

the individual electrode polarization of an EMIM Cl solution in a single electrolyte reactor.

Figure 4.1 shows the cathode and anode electrode polarization in the single channel flow reactor

with a 4.7 mole% EMIM Cl electrolyte solution compared to that of a 1 M aqueous KCl

electrolyte solution in the single channel flow reactor and an 18 mole% EMIM BF4 electrolyte

solution in the dual channel flow reactor. With the EMIM Cl solution, no anode poisoning was

observed, confirming that the BF4 anion is responsible for the anode poisoning. Consequently, a

second electrolyte stream, the anolyte, is not necessary when using an EMIM Cl electrolyte

solution. Furthermore, the current densities achieved with the EMIM Cl solution are comparable

to that of an aqueous solution.

4.3.3 EMIM+ Cl

- Performance in the Single Channel Electrolyte Reactor

Figure 4.2a shows the partial current densities for CO and H2 when using a 4.7 mole%

EMIM Cl electrolyte pumped through the reactor at 0.5 mL/min with a Ag cathode and a Pt

anode. High selectivity for CO is maintained with the substitution of the BF4 anion with the Cl

anion on the EMIM complex. Figure 4.2b shows a comparison of the current densities in a

single channel flow reactor with an aqueous 1 M KCl electrolyte, an 18 mole% EMIM BF4

Figure 4.2: (a) Partial current densities for CO and H2 for a 4.7 mole% EMIM Cl solution flowed at 0.5 mL/min with

a Ag cathode and a Pt anode. (b) comparison of comparison of the current densities in a single channel flow

reactor with an aqueous KCl electrolyte, an EMIM BF4 electrolyte and an EMIM Cl electrolyte.

0

20

40

60

80

-3.25-2.75-2.25-1.75-1.25

Cu

rre

nt D

en

sity

(mA

/cm

2 )

Cell Potential (V)

CO

H2

-120

-100

-80

-60

-40

-20

0

-3.3-2.8-2.3-1.8-1.3

Cu

rre

nt D

en

sity

(mA

/cm

2 )

Cell Potential

EMIM BF4

EMIM Cl

KCl

(a) (b)

0

20

40

60

80

-3.25-2.75-2.25-1.75-1.25

Cu

rre

nt D

en

sity

(mA

/cm

2)

Cell Potential (V)

CO

H2

-120

-100

-80

-60

-40

-20

0

-3.3-2.8-2.3-1.8-1.3

Cu

rre

nt D

en

sity

(mA

/cm

2)

Cell Potential

EMIM BF4

EMIM Cl

KCl

-2

-1.5

-1

-0.5

0

0.5

1

1.5

2

-150-100-500

Ele

ctro

de

Po

ten

tial

(V v

s A

g/A

gCl)

Current Density (mA/cm2)

EMIM BF4

EMIM Cl

KCl

-2

-1.5

-1

-0.5

0

0.5

1

1.5

2

-150-100-500

Ele

ctro

de

Po

ten

tial

(V v

s A

g/A

gCl)

Current Density (mA/cm2)

EMIM BF4

EMIM Cl

KCl

-2

-1.5

-1

-0.5

0

0.5

1

1.5

2

-150-100-500

Ele

ctro

de

Po

ten

tial

(V v

s A

g/A

gCl)

Current Density (mA/cm2)

EMIM BF4

EMIM Cl

KCl

Page 73: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

64

electrolyte and a 4.7 mole% EMIM Cl electrolyte. When using an EMIM Cl solution, the partial

current density for CO drastically increases such that its performance is comparable to that with

an aqueous electrolyte stream with the BF4 for Cl substitution.

Figure 4.3a shows the conductivity of the EMIM Cl solution with various concentrations

of EMIM Cl. The conductivity can only be improved roughly 7% by changing the concentration

of EMIM Cl because the conductivity of the earlier trial (at a concentration of 4.7 mole%) was

close to the maximum conductivity achievable with an EMIM Cl / aqueous mixture.

4.3.4 EMIM+ Cl

- Conductivity Studies

While EMIM Cl performs well compared to EMIM BF4, the performance was further

improved via bolstering the conductivity with a secondary salt, KOH, as was shown to be

possible in section 4.3.1. Figure 4.4 shows the performance of the reactor with (a) 4.7 mole%

EMIM OH and (b) 4.7 mole% EMIM Cl with a secondary ion, KOH at 1 M. When using the

EMIM OH electrolyte, the current density was expected to be greater than when using the

EMIM Cl electrolyte because of the effect of pH shown in Chapter 2. However, while the

current density with the EMIM OH electrolyte was better than with the EMIM BF4 electrolyte,

with the EMIM OH electrolyte, the current density was worse than with the EMIM Cl

Figure 4.3: (a) Conductivity for various concentration EMIM Cl solutions and (b) individual polarization plots for

EMIM Cl electrolyte solutions with concentrations between 4.7 and 10.9 mole%.

90.0

92.0

94.0

96.0

98.0

4 5 6 7 8 9 10 11 12

Co

nd

uct

ivit

y (m

S/cm

)

mole% EMIM Cl

-2.00

-1.50

-1.00

-0.50

0.00

0.50

1.00

1.50

-40-35-30-25-20-15-10-50

Ele

ctro

de

Po

ten

tial

(V v

s A

g/A

gCl)

Current Density (mA/cm2)

10.9% EMIM Cl

8.9% EMIM Cl

7.5% EMIM Cl

5.8% EMIM Cl

4.7% EMIM Cl

(a) (b)

90.0

92.0

94.0

96.0

98.0

4 5 6 7 8 9 10 11 12

Co

nd

uct

ivit

y (m

S/cm

)

mole% EMIM Cl

-2.00

-1.50

-1.00

-0.50

0.00

0.50

1.00

1.50

-40-35-30-25-20-15-10-50

Ele

ctro

de

Po

ten

tial

(V v

s A

g/A

gCl)

Current Density (mA/cm2)

10.9% EMIM Cl

8.9% EMIM Cl

7.5% EMIM Cl

5.8% EMIM Cl

4.7% EMIM Cl

90.0

92.0

94.0

96.0

98.0

4 5 6 7 8 9 10 11 12

Co

nd

uct

ivit

y (m

S/cm

)

mole% EMIM Cl

-2.00

-1.50

-1.00

-0.50

0.00

0.50

1.00

1.50

-40-35-30-25-20-15-10-50

Ele

ctro

de

Po

ten

tial

(V v

s A

g/A

gCl)

Current Density (mA/cm2)

10.9% EMIM Cl

8.9% EMIM Cl

7.5% EMIM Cl

5.8% EMIM Cl

4.7% EMIM Cl

Page 74: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

65

electrolyte. This is believed to be due to the lower conductivity (70 mS/cm vs. 90 mS/cm) with

the hydroxide salt as opposed the chloride salt. The electrolyte consisting of a mixture of

EMIM Cl and KOH salts outperformed all other electrolyte solutions shown here due to the

higher conductivity (206 mS/cm) of the electrolyte and the improved anode performance at basic

conditions.

4.4 Conclusions

Drastic improvements have been achieved in the partial current density for CO with

EMIM+ based electrolytes. Specifically, substituting the BF4 anion for either a OH anion or a Cl

anion enabled partial current densities for CO greater than 60 mA/cm2 at a reactor potential of

3 V. Furthermore, the substitution of BF4 for OH or Cl eliminated anode poisoning from the BF4

anion, thus enabling the use of a single channel electrolyte stream. The elimination of the

membrane and the second electrolyte stream reduced the reactor resistance (from EMIM+

poisoning the Nafion as discussed in Chapter 3), which led to the higher current densities.

The conductivity of an EMIM Cl / aqueous mixture is limited to <100 mS/cm, however,

the conductivity of the electrolyte steam can be improved with the addition of a secondary salt

(i.e., KOH). Specifically, the addition of KOH to an EMIM Cl solution, drastically increased the

Figure 4.4: Partial current density for (a) CO and (b) H2 with various ionic liquid electrolyte streams.

0

20

40

60

80

100

120

-3.25-2.75-2.25-1.75-1.25

Par

tial

Cu

rre

nt

De

nsi

ty C

O (

mA

/cm

2 )

Cell Potential (V)

EMIM BF4EMIM ClEMIM Cl & KOHEMIM OH

0

5

10

15

20

-3.25-2.75-2.25-1.75-1.25

Par

tial

Cu

rre

nt

De

nsi

ty H

2(m

A/c

m2 )

Cell Potential (V)

EMIM BF4EMIM ClEMIM Cl & KOHEMIM OH

(a) (b)

-2

-1.5

-1

-0.5

0

0.5

1

1.5

2

-150-100-500

Ele

ctro

de

Po

ten

tial

(V v

s A

g/A

gCl)

Current Density (mA/cm2)

EMIM BF4

EMIM Cl

KCl

-2

-1.5

-1

-0.5

0

0.5

1

1.5

2

-150-100-500

Ele

ctro

de

Po

ten

tial

(V v

s A

g/A

gCl)

Current Density (mA/cm2)

EMIM BF4

EMIM Cl

KCl

-2

-1.5

-1

-0.5

0

0.5

1

1.5

2

-150-100-500

Ele

ctro

de

Po

ten

tial

(V v

s A

g/A

gCl)

Current Density (mA/cm2)

EMIM BF4

EMIM Cl

KCl

-2

-1.5

-1

-0.5

0

0.5

1

1.5

2

-150-100-500

Ele

ctro

de

Po

ten

tial

(V v

s A

g/A

gCl)

Current Density (mA/cm2)

EMIM BF4

EMIM Cl

KCl

-2

-1.5

-1

-0.5

0

0.5

1

1.5

2

-150-100-500

Ele

ctro

de

Po

ten

tial

(V v

s A

g/A

gCl)

Current Density (mA/cm2)

EMIM BF4

EMIM Cl

KCl

-2

-1.5

-1

-0.5

0

0.5

1

1.5

2

-150-100-500

Ele

ctro

de

Po

ten

tial

(V v

s A

g/A

gCl)

Current Density (mA/cm2)

EMIM BF4

EMIM Cl

KCl

Page 75: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

66

electrolyte conductivity (206 mS/cm), and consequently, enabled partial current densities for CO

greater than 100 mA/cm2 at a reactor potential of 3V.

4.5 References

1. Y. Hori, Electrochemical CO2 Reduction on Metal Electrodes, Modern Aspects of

Electrochemistry, 2008, 42 p. 89-189.

2. A. T. Bell, Basic Energy Needs: Catalysis for Energy. Department of Energy, 2008.

PNNL17214

3. M. R. Dubois and D. L. Dubois, Development of Molecular Electrocatalysts for CO2

Reduction and H(2) Production/Oxidation, Accounts of Chemical Research, 2009, 42

(12), p. 1974-1982.

4. Brian A. Rosen, Amin Salehi-Khojin, Michael R. Thorson, Wei Zhu, Devin T. Whipple,

Paul J. A. Kenis and Richard I. Masel, Ionic Liquid–Mediated Selective Conversion of

CO2 to CO at Low Overpotentials, Science, 2011, 334 (6056), p. 643-644.

5. D. T. Whipple and P. J. A. Kenis, Prospects of CO2 Utilization via Direct Heterogeneous

Electrochemical Reduction, Journal of Physical Chemistry Letters, 2010, 1 (24), p. 3451-

3458.

6. Narasi Sridhar, Carbon Dioxide Utilization: Electrochemical Conversion of CO2 -

Opportunities and Challenges. Det Norske Veritas, 2011. 07 1-20.

7. E. B. Cole, P. S. Lakkaraju, D. M. Rampulla, A. J. Morris, E. Abelev and A. B. Bocarsly,

Using a One-Electron Shuttle for the Multielectron Reduction of CO2 to Methanol:

Kinetic, Mechanistic, and Structural Insights, Journal of the American Chemical Society,

2010, 132 (33), p. 11539-11551.

8. A. J. Morris, R. T. McGibbon and A. B. Bocarsly, Electrocatalytic Carbon Dioxide

Activation: The Rate-Determining Step of Pyridinium-Catalyzed CO2 Reduction,

Chemsuschem, 2011, 4 (2), p. 191-196.

9. G. Seshadri, C. Lin and A. B. Bocarsly, A New Homogeneous Electrocatalyst for the

Reduction of Carbon-Dioxide to Methanol at Low Overpotential, Journal of

Electroanalytical Chemistry, 1994, 372 (1-2), p. 145-150.

10. S. Kaneco, K. Iiba, H. Katsumata, T. Suzuki and K. Ohta, Electrochemical Reduction of

High Pressure CO2 at a Cu Electrode in Cold Methanol, Electrochimica Acta, 2006, 51

(23), p. 4880-4885.

11. S. Kaneco, K. Iiba, H. Katsumata, T. Suzuki and K. Ohta, Electrochemical Reduction of

High Pressure Carbon Dioxide at a Cu Electrode in Cold Methanol with CsOH

Supporting Salt, Chemical Engineering Journal, 2007, 128 (1), p. 47-50.

12. S. Kaneco, R. Iwao, K. Iiba, K. Ohta and T. Mizuno, Electrochemical Conversion of

Carbon Dioxide to Formic Acid on Pb in KOH/Methanol Electrolyte at Ambient

Temperature and Pressure, Energy, 1998, 23 (12), p. 1107-1112.

13. S. Kaneco, H. Katsumata, T. Suzuki and K. Ohta, Electrochemical Reduction of CO2 to

Methane at the Cu Electrode in Methanol with Sodium Supporting Salts and its

Comparison with Other Alkaline Salts, Energy Fuels, 2006, 20 (1), p. 409-414.

Page 76: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

67

14. A. Murata and Y. Hori, Product Selectivity Affected by Cationic Species in

Electrochemical Reduction of CO2 and CO at a Cu Electrode, Bulletin of the Chemical

Society of Japan, 1991, 64 (1), p. 123-127.

15. Y. Hori and S. Suzuki, Electrolytic Reduction of Carbon-Dioxide at Mercury-Electrode

in Aqueous-Solution, Bulletin of the Chemical Society of Japan, 1982, 55 (3), p. 660-665.

16. D. T. Whipple, E. C. Finke and P. J. A. Kenis, Microfluidic Reactor for the

Electrochemical Reduction of Carbon Dioxide: The Effect of pH, Electrochemical and

Solid State Letters, 2010, 13 (9), p. D109-D111.

Page 77: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

68

Chapter 5

Organometallic Catalysts for the Electrochemical Reduction of CO2

5.1 Introduction

Storage of electrical energy in chemical bonds (i.e., chemical fuels) provides an excellent

storage medium for intermittent renewable energy sources, thereby enabling “artificial

photosynthesis”. However, artificial photosynthesis, using the electrochemical reduction of CO2,

is obstructed by the lack of catalysts for CO2 reduction with (a) sufficiently low overpotentials to

be energetically efficient [1-4] and (b) high enough current densities to be profitable [2,5].

According to a recent DOE report, "electron conversion efficiencies of greater than 50 percent

can be obtained, but at the expense of very high overpotentials (~1.5 V)", consequently more

efficient and higher throughput catalysts need development [1,6-8].

The catalytic activity of all commonly used catalysts’ comprised of a single metal have

previously been studied [3,9-11]. From this, four distinct classes of catalysts have been

identified: catalysts selective for carbon monoxide (e.g., Ag, Au, Zn), catalysts selective for

formic acid (e.g., Sn, Cd, Ti), catalysts selective for hydrogen (e.g., Pt, Ni, Fe) and copper, which

produces a wide range of hydrocarbons [3,9].

A primary challenge lies in the formation of the “CO2- intermediate” as discussed in

Chapter 1 [14-16]. Using one of the catalysts selective for CO, Ag, Rosen et al. (see Chapter 2)

demonstrated the use of an amine in the electrolyte solution, which can serve the role of co-

catalyst and reduce the onset potential for CO2 production [1]. Specifically, EMIM BF4 was

shown to reduce the onset potential for CO2 reduction such that the overpotential is < 0.17 V;

however, at the expense of low current densities [1]. Similarly, another amine, Pyridine, has

Page 78: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

69

been shown to make the electrochemical reduction of

CO2 to CO on a copper electrode more selective for

methanol at a low overpotential [17-19]. Clearly,

amines can play a significant role in assisting the

electrochemical reduction of CO2 in terms of reducing

the overpotentials and increasing the product

selectivity.

Here, we are interested in developing an

organometallic complex to incorporate amines on the

electrode surface and thereby, develop a catalyst with low overpotentials, high selectivity, and

high current densities. Thorum et al. previously developed a copper based amine, copper

complex with 3,5-diamino-1,2,4-triazole

(Figure 5.1), for oxygen reduction

[12,13]. The copper complex was found

to have similar performance to that of

Pt/C (Figure 5.2), which constitutes a

huge breakthrough in the development

of organometallic complexes.

Here, we explored the use of this

copper complex, CuDAT, and

modifications of this complex for CO2

reduction. Specifically, we explored

several Ag and Cu based amine

Figure 5.1: A novel carbon-supported copper

complex of 3,5-diamino-1,2,4-triazole

(CuDAT) developed by the Gewirth lab.

Reprinted with permission from: Matthew S

Thorum, Jessica Yadav and Andrew A Gewirth,

“Oxygen Reduction Activity of a Copper

Complex of 3,5-Diamino-1,2,4-triazole

Supported on Carbon Black.” Angewandte

Chemie International Edition, 2009. 48 (1): p.

165-167. Copyright (2008) John Wiley and

Sons. [12]

Figure 5.2: Individual electrode polarization curves for a

microfluidic H2/O2 fuel cell with a CuDAT catalyst compared t

omore traditional catalysts. Reprinted with permission from:

Fikile R. Brushett, Matthew S. Thorum, Nicholas S. Lioutas,

Matthew S. Naughton, Claire Tornow, Huei-Ru “Molly” Jhong,

Andrew A. Gewirth and Paul J. A. Kenis, “A Carbon-Supported

Copper Complex of 3,5-Diamino-1,2,4-triazole as a Cathode

Catalyst for Alkaline Fuel Cell Applications.” Journal of the

American Chemical Society, 2010. 132 (35): p. 12185-12187.

Copyright (2010) American Chemical Society.[13]

0

0.2

0.4

0.6

0.8

1

0 100 200 300 400

Current Density (mA/cm2)

Po

ten

tial

(V)

vs R

HE

Pt/C cathode

Cu-tri/C cathode

Ag/C cathode

Pt/C anode

1 M KOH

Page 79: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

70

complexes synthesized in the Gewirth lab. Ag was chosen because, as a bare metal, Ag has

excellent selectivity for CO at high current densities and because of the Kenis lab’s extensive

experience with Ag [20-24]. Cu was chosen based on the experience Dr. Gewirth’s lab has with

Cu-based organometallic catalyst development. Furthermore, Cu is the only bare-metal which

has been demonstrated to be capable of producing large hydrocarbons from CO2; however, the

potentials necessary for direct conversion of CO2 to hydrocarbons are energetically inefficient

due to their high overpotentials [21,25-28].

At present, one challenge with using copper catalysts is improving the reaction

selectivity. Copper tends to produces a wide range of products and is unselective for any one

product due to its complex reaction pathway [29-32]. Specifically, methane requires an 8-

electron pathway with many potential alternative reaction pathways, which result in the

formation of different chemicals. For example, ethylene, methanol, carbon monoxide, and

formic acid deviate from the methane reaction pathway after the 4th

, 4th

or 5th

, 2nd

, and 1st

electron transfer, respectively, en route to their own 12, 6, 2 and 2 electron-transfer reaction

pathways. Here we explore the use of Cu-based organometallic complexes to influence the

reaction and change the product selectivity.

5.2 Experimental

5.2.1 Catalyst Synthesis

All catalysts were synthesized by Claire Tornow in the Gewirth Lab. Below are the

synthesis procedures for the catalysts used in this study.

Silver phthalocyanine (75 % dye content, Sigma-Aldrich) (AgPc) was used as received.

Silver phthalocyanine (61.7 mg) was readily dissolved in H2O. Vulcan carbon (162.7 mg) was

Page 80: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

71

added to the solution and the mixture was sonicated for 30 minutes and allowed to sit overnight.

The sample was centrifuged once and placed in oven at 120ºC for 24 hours.

Silver pyrazole (AgPz) was synthesized according to Masciocchi, et al. in the presence of

Vulcan [33]. AgNO3 (35.5 mg) was dissolved in H2O. To this, ~0.5 mL NH3 was added. Vulcan

(160.8 mg) was added and the mixture was sonicated for 10 minutes. H20 solvated pyrazole was

added dropwise and the resulting mixture was sonicated for 30 minutes. This was allowed to sit

overnight. The mixture was centrifuged and placed in a vacuum oven at 90ºC for 4 hours.

Silver complex of tris(2-pyridylmethyl)amine (AgTPA) was synthesized using a 1:1

molar ratio of TPA to Ag. The TPA ligand (44.4 mg) was dissolved in EtOH. Ag2SO4 (30.4 mg)

was dissolved in H2O. Vulcan (156.3 mg) was added to the solution of Ag2SO4 and the solution

was sonicated for 10 minutes. TPA solution was added dropwise to the carbon mixture. The

resulting mixture was sonicated for an additional 30 minutes and allowed to sit overnight. The

mixture was centrifuged once and placed in a vacuum oven at 90ºC for 4 hours.

The silver complex of 3,5-diamino-1,2,4-triazole (AgDAT) was synthesized using a 2:1

molar ratio of DAT to Ag. The DAT ligand (502.4 mg) was dissolved in H2O and Ag2SO4

(393.3 mg) was also dissolved in H2O. Vulcan (2.0325 mg) was added to the solution of Ag2SO4

and sonicated for 10 minutes. Solution of DAT was added dropwise to the carbon suspension

and the resulting mixture was sonicated for 30 minutes and allowed to sit overnight. The mixture

was centrifuged once and placed in a vacuum oven at 90ºC for 4 hours.

The following procedures are for the copper organometallic catalysts synthesized for this

study: The ligands 3-amino-4-methyl-1,2,4-triazole (Sigma-Aldrich), 4-amino-3,5-dipyridyl-

1,2,4-triazole (Sigma-Aldrich), 3,4,5-triamino-1,2,4-triazole (BIONET Research Intermediates,

Key Organics), and 4-amino-3,5-dimethyl-1,2,4-triazole (Alfa Aesar) were used as received. Cu

Page 81: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

72

complexes were synthesized by addition of the respective triazole ligand to an aqueous solution

containing Cu(SO4) · 5 H2O (Sigma-Aldrich, 99.999 %) in a 2:1 molar ratio. The resulting

precipitates were collected via vacuum filtration. The appropriate amounts of the Cu complexes

(to generate loadings of a 30 % complex on carbon) were placed in an aqueous solution

containing Vulcan XC-72 (Cabot Corporation) to create a suspension which was further

sonicated for 1 hour. The material was collected by first centrifuging, then drying the sample

under vacuum at 90ºC to produce the carbon-supported Cu catalysts.

5.2.2 Composition

The organometallic silver catalysts used in this study were analyzed via elemental

analysis by Claire Tornow for their silver content and the AgDAT, AgPc, AgTPA, and AgPz

catalysts were found to have silver mass% loadings of 8.62, 6.25, 4.32, and 2.51 %, respectively.

Furthermore, the copper catalysts, CT01-301, and CT01-289 have metal mass% loadings of 1.21,

and 1.26 %.

5.2.3 UV-Vis Analysis

The analysis of the rate of formic acid production was carried out based on the

concentrations of formic acid in the effluent electrolyte stream. The formic acid concentration

was measured using a modified version of a procedure developed by Sleat and Mah [34].

0.4 mL of a fresh reactant solution (0.5 g citric acid, 10 g acetamide, and 100 mL isopropyl

alcohol) was mixed with 0.2 mL of the effluent sample containing formic acid, 0.02 mL of 30 %

sodium acetate, and 1.4 mL acetic anhydride. Following mixing, these solutions were allowed to

react for 24 hrs before being placed in cuvettes (BrandTech Scientific, Brand-UV Cuvettes 7591)

for UV-Vis (Perkin Elmer, Lambda 650) analysis at a wavelength of 515 nm. For each

experiment, a new calibration curve was prepared using known standard concentrations (1, 5, 10,

Page 82: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

73

20, 40, and 60 mM formic acid in 1 M KOH). A line was fit to the intensity of the UV peak at

515 nm relative to the formic acid concentration using a method of least squares of the error to

make a linear calibration curve. The UV peak intensity of the samples prepared from the

effluent stream was compared to the linear calibration to determine the formic acid concentration

in the effluence electrolyte sample. Figure 5.3 shows vials filled with a mixture of the formic

acid stream and the reactant stream for the analysis of the formic acid concentration.

5.2.4 Three Electrode Cell Operation

The three electrode cell experiments were carried out using a CH Instruments

bipotentiostat. The electrochemical cell consisted of a Pt gauze counter electrode and a “no-

leak” Ag/AgCl reference electrode (Cypress), separated from the working electrode by means of

a Luggin capillary. The catalysts for the three electrode cell experiments were prepared as

follows: catalyst inks containing the carbon supported Ag complex (1.0 mg mL-1) and Nafion

Figure 5.3: (top) Formic acid control solutions mixed with reactant solutions for UV-Vis calibration and (bottom)

effluent electrolyte mixed with reactant solution to analyze in UV-Vis for formic acid concentrations.

1 mM 5 mM 10 mM 20 mM 40 mM 60 mM

2.0 V 2.25 V 2.5 V 2.75 V 3.0 V

Page 83: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

74

solution (4 μL mL-1, 5 wt%, Aldrich) were prepared in a 5% butanol-water mixture and

sonicated prior to electrode preparation. A 20 μL drop of the catalyst ink was deposited on a

rotating ring-disk electrode (Pine Instruments) comprised of a polished (0.05 micron alumina)

glassy carbon disk electrode (0.196 cm2) with a Pt ring.

5.2.5 Reactor Operation

The reactor operational procedure was identical to that used in Chapter 2. Experimental

data, in which current densities were above 200 mA/cm2, was not reported here because, at these

high current densities, CO, and H2 are produced at sufficiently high rates to quickly form bubbles

in the electrolyte solution, partially blocking the electrode and resulting in inconsistent reactor

performance. In all of the studies outlined here, unless otherwise stated, a 1 M KOH electrolyte

solution was flowed through the electrolyte chamber at 0.5 mL/min. All of the experiments in

this study used a 1 mg/cm2 cathode loading on Sigracet 35BC. All of the anodes used in this

study had a 2 mg/cm2 Pt loading on Sigracet 35BC. For the analysis of the copper and

organometallic copper-complexes, a single GC sample (rather than the normal three) was used

for the analysis at each potential applied because the elution times for ethylene requires the GC

to run for 12 minutes for each sample.

5.3 Results and Discussion

5.3.1 Cyclical Voltammetry in a flow reactor

One of the primary advantages of the electrochemical reactor used in this study is the

ability to carry out cyclical voltammetry in-situ for the investigation of catalyst performance.

Figure 5.4 shows cyclical voltammetry scans taken within the flow reactor of the cathode relative

to a Ag/AgCl reference electrode between -0.4 and -1.4 V vs. Ag/AgCl for a silver electrode

(Ag), a carbon supported silver triazole electrode (AgDAT), a carbon supported silver pyrazole

Page 84: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

75

electrode (AgPz), a carbon supported silver tripyridylamine electrode (AgTPA), and a carbon

supported silver phthalocyanine electrode (AgPc). A 1 M KOH electrolyte was flowed through

the electrolyte chamber at 0.5 mL/min and either CO2 or N2 was flowed behind the cathode in

Figure 5.4: Cyclic voltammetry reduction activity measurements for (a) AgPc, (b) AgDAT, (c) AgPz, (d) Ag, and

(e) AgTPA with a CO2 feed compared to their respective activity under Ar. Data was recorded in a flow reactor with

a 1 M KOH electrolyte solution flowed at 0.5 mL/min scanned at 25 mV/s

0

5

10

15

20

25

30

-1.5 -1.3 -1.1 -0.9 -0.7 -0.5

Cu

rre

nt D

en

sity

(mA

/cm

2)

Potential (V vs Ag/AgCl)

N2

CO2

0

10

20

30

40

50

60

70

-1.5 -1.3 -1.1 -0.9 -0.7 -0.5

Cu

rre

nt D

en

sity

(mA

/cm

2 )

Potential (V vs Ag/AgCl)

N2

CO2

0

5

10

15

20

25

30

35

-1.5 -1.3 -1.1 -0.9 -0.7 -0.5

Cu

rre

nt D

en

sity

(mA

/cm

2 )

Potential (V vs Ag/AgCl)

CO2

N2

0

5

10

15

20

25

-1.5 -1.3 -1.1 -0.9 -0.7 -0.5

Cu

rre

nt D

en

sity

(mA

/cm

2)

Potential (V vs Ag/AgCl)

CO2

N2

-4

-2

0

2

4

6

8

10

-1.5 -1.3 -1.1 -0.9 -0.7 -0.5

Cu

rre

nt D

en

sity

(mA

/cm

2)

Potential (V vs Ag/AgCl)

CO2

N2

(a) (b)

(c) (d)

(e)

0

5

10

15

20

25

-1.5 -1.3 -1.1 -0.9 -0.7 -0.5

Cu

rre

nt D

en

sity

(mA

/cm

2)

Potential (V vs Ag/AgCl)

CO2

N2

0

5

10

15

20

25

-1.5 -1.3 -1.1 -0.9 -0.7 -0.5

Cu

rre

nt D

en

sity

(mA

/cm

2)

Potential (V vs Ag/AgCl)

CO2

N2

Page 85: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

76

the gas chamber at a flow rate of 5

SCCM. These graphs provide

information regarding (a) the onset

potentials for various products and (b) the

product selectivity. An increase in current

density resulting from a change in

reactant gas (N2 to CO2) suggests that the

catalyst is active for CO2 reduction and

will produce a carbon product (e.g., CO,

CH4, C2H4, CHOH, or HCO2H). The

more drastic the change in current density, the more selective the catalyst is.

As seen in Figure 5.4, when CO2 is substituted for Ar as the reactant gas, with Ag,

AgDAT, AgPz, or AgPc as the cathode catalysts (a) an earlier onset potential and (b) an increase

in the current density are observed. The improved performance in the presence of CO2 suggests

that Ag, AgDAT, AgPz, and AgPc are active for CO2 reduction and aid the production of a

carbon product (e.g., CO, CH4, C2H4, CHOH, or HCO2H). However, with AgTPA as the

cathode catalysts, when CO2 is substituted for Ar as the reactant gas, a small decrease change in

the current density is observed. This change in catalyst activity suggests CO2 reduction may be

active on this catalyst. However, the decrease in current density also suggests that the kinetics

for CO2 reduction for the AgTPA catalyst may be quite low. Compared to the other catalysts,

AgTPA appears to suppress CO2 reduction kinetics.

The current densities achieved with three of the electrodes with the Ag-amine complex

catalyst (DAT, Pz, Pc) were comparable to that of the Ag catalyst when compared relative to the

Figure 5.5: Cyclic voltammetry measurements comparing the

onset potentials of CO production for Ag particles, AgDAT/C ,

AgPz/C, and AgPc/C. Data was recorded in 1 M KOH, under

CO2, at 25 mV/s. Figure courtesy of Claire Tornow.

0.00

10.00

20.00

30.00

40.00

50.00

60.00

-1.5 -1.3 -1.1 -0.9 -0.7 -0.5

Cu

rre

nt d

en

sity

(mA

/cm

2 )

Potential (V vs Ag/AgCl)

TPA

Ag

Pc

DAT

Pz

0.00

10.00

20.00

30.00

40.00

50.00

60.00

-1.5 -1.3 -1.1 -0.9 -0.7 -0.5

Cu

rre

nt d

en

sity

(mA

/cm

2)

Potential (V vs Ag/AgCl)

TPA

Ag

Pc

DAT

Pz

Page 86: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

77

total mass loading of catalyst. The high performance with these organometallic catalysts is very

promising as Ag is the best catalyst in the literature for CO production.[3]

Figure 5.5 compares the cyclical voltammetry scans for the catalysts determined from the

initial CV scans to be selective for CO (AgTPA, Ag, AgPC, AgDAT, and AgPz). Earlier onsets

are seen with AgPz, AgPc, and AgDAT in the presence of CO as compared to Ag

(Pz<DAT<Pc<Ag). The decrease in current density with AgTPA further suggests that while the

complex may be active for CO2 reduction, the kinetics for CO2 reduction are reduced.

5.3.2 Cyclical Voltammetry in a 3-electrode cell

Figure 5.6 shows cyclical voltammetry taken by Tornow from Dr. Gewirth’s lab in a 3-

electrode cell with a 1 M KOH electrolyte and a scan rate of 25 mV/s. Six catalysts were tested

including: silver (Ag), carbon supported silver (Ag/C), carbon supported silver triazole

(AgDAT), carbon supported silver pyrazile (AgPz), carbon supported silver tripiradine

(AgTPA), and carbon supported silver pthyocyline (AgPc) with either CO2 or Ar bubbled

through the electrolyte prior to the cyclical voltammetry. Similar to the CVs in-situ for the

reactor, these results provide information regarding the onset potentials and the relative current

densities for various catalysts. Furthermore, information can be obtained regarding the activity

for CO2 reduction based any changes in performance from when CO2 (rather than Ar) is bubbled

through the electrolyte.

Page 87: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

78

Figure 5.6: Cyclic voltammetry measurements of reduction activity of Ag catalysts with a CO2 feed compared to

their respective activity under Ar. Data was recorded in 1 M KOH at 25 mV/s. Figure courtesy of Claire Tornow.

(a) (b)

(c) (d)

(e) (f)

Page 88: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

79

The Ag, AgDAT, AgPz, and AgPc

catalysts show a decrease in the onset

potential and an increase in the current

density in presence of CO2 as opposed to

Ar, confirming what was observed earlier in

the flow reactor that these catalysts are

active for CO2 reduction and will produce a

carbon product (e.g., CO, CH4, C2H4,

CHOH, or HCO2H). However, the Ag/C

and AgTPA catalysts show no change in

both the onset potential and the current density in presence of CO2 as opposed to Ar, suggesting

that these catalysts may not have activity for CO2 reduction. Of note, in the flow reactor (section

5.3.1), the addition of CO2 decreased the observed current density for the AgTPA catalyst,

thereby suggesting CO2 reduction activity, but also poor kinetics. However, here, with the three

electrode cell, we do not observe any change in activity from bubbling CO2 into the liquid

electrolyte of the three electrode cell.

The current densities achieved with the Ag-amine complex (DAT, Pz, Pc) catalysts are

comparable (at the same total mass loading) to that observed with the Ag catalyst, which is the

best catalyst in the literature for CO production. While higher current densities are achieved

with Ag/C as opposed to Ag, they have poor selectivity for CO2 reduction.

Similarly to Figure 5.5, we compare the cyclical voltammetry scans for the catalysts

determined from the initial CV scans to be selective for CO (Ag, AgDAT, AgPz, and AgPC) in

Figure 5.7. Table 5.1 compares the onset potentials as determined by the potential at which a

Figure 5.7: Cyclic voltammetry measurements comparing the

onset potentials of CO production for Ag particles, AgDAT/C ,

AgPz/C, and AgPc/C. Data was recorded in 1 M KOH, under

CO2, at 25 mV/s. Figure courtesy of Claire Tornow.

Page 89: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

80

specific current density is achieved. Earlier onsets are seen with the Pz, Pc, and DAT in the

presence of CO as compared to Ag (Pz<DAT<Pc<Ag).

Table 5.1. Comparison of onset potentials for catalysts under CO2 and Ar, reported as the potentials at

which the current density reaches -0.05, -0.1, and -0.2 mA/cm2.

Catalyst

Onset Potentials under CO2 (V) Onset Potentials under Ar (V)

-0.05 mA/cm2 -0.1 mA/cm

2 -0.2 mA/cm

2 -0.05 mA/cm

2 -0.1 mA/cm

2 -0.2 mA/cm

2

Ag -1.20 -1.22 -1.25 -1.38 NA NA

Ag/C -0.86 -0.92 -1.00 -0.93 -1.00 -1.09

AgDAT/C -1.03 -1.14 -1.21 -1.08 -1.20 -1.29

AgTPA/C -1.09 -1.16 -1.24 -1.14 -1.20 -1.26

AgPz/C -1.01 -1.13 -1.21 -1.06 -1.19 -1.30

AgPc/C -1.05 -1.15 -1.24 -1.18 -1.26 -1.34

The onset for H2 is as follows: Ag/C<Pz<DAT<TPA<Pc<Ag. Ag/C has a very early

onset potential in the presence of Ar, thereby suggesting a reduction in the onset potential for H2

production. Because no improvement is observed for the Ag/C catalyst from the bubbling of

CO2 in the electrolyte solution carbon products (e.g., CO, CH4, C2H4, CHOH, or HCO2H) are not

expected in the flow reactor tests.

5.3.3 Operation in a Flow Reactor

Figure 5.8 shows the partial current densities for CO and H2, the cathode overpotentials

and the partial current normalized to the silver loading for the AgDAT, AgPz, AgPc, and AgTPA

organometallic catalysts compared to both Ag and Ag/C in the flow reactor with a 1 M KOH

electrolyte solution (0.5 mL/min) and a CO2 gas feed (5 SCCM). Regarding the partial current

density for CO relative to the absolute loading of catalyst, we observe, a higher current density

for AgDAT and AgPz as compared to Ag. Additionally, we observe a slightly lower current

density for AgPc as compared to Ag and we observe CO suppression with AgTPA. In terms of

Page 90: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

81

the partial current density for CO normalized to the mass of the Ag loading, we observe that the

current density per Ag loading is improved by more than a factor of 10 with the organometallic

catalysts as compared to Ag. This is very promising as the silver content of these organometallic

complexes are quite low in comparison to the pure silver catalyst.

Figure 5.8b shows that the Ag/C catalyst has a large partial current density for H2, which

is the undesired reaction because other means are more efficient means of making H2,. However,

with the amine-based catalysts, low H2 partial current densities are observed. The partial current

densities for H2 increase in the following order: Ag<DAT<Pc<TPA<Pz<Ag/C.

In the cathode electrode polarization plot, Figure 5.8d, the overpotentials for the catalysts

Figure 5.8: Partial current densities for (a) CO and (b) H2 relative to electrode surface area as well as (c) partial

current density for CO relative to the Ag loading and (d) cathode polarization curves for electrodes with 1 mg.cm2

Ag and Ag-ligand catalyst loadings in a flow reactor with a 1 M KOH electrolyte flowed at 0.5 mL/min and CO2

flowed over the cathode at 5 SCCM..

0

20

40

60

80

100

120

-2.0-1.8-1.6-1.4-1.2-1.0

Par

tial

Cu

rre

nt

De

nsi

ty C

O (

mA

/cm

2)

Cathode vs Ag/AgCl (V)

AgDAT/CAgPz/CAgPc/CAgTPA/CAgAg/C

0

500

1,000

1,500

2,000

-2.0-1.8-1.6-1.4-1.2-1.0

Par

tial

Cu

rre

nt

De

nsi

ty C

O (

mA

/mg

Ag)

Cathode vs Ag/AgCl (V)

AgDAT/CAgPz/CAgPc/CAgTPA/CAgAg/C

0

5

10

15

20

25

30

35

40

-2.0-1.8-1.6-1.4-1.2-1.0

Par

tial

Cu

rre

nt

De

nsi

ty H

2(m

A/c

m2)

Cathode vs Ag/AgCl (V)

-2.5

-2.3

-2.1

-1.9

-1.7

-1.5

-1.3

-1.1

0 25 50 75 100 125

Cat

ho

de

Po

ten

tial

(V v

s A

g/A

gCl)

Current Density (mA/cm2)

0

20

40

60

80

100

120

-2.0-1.8-1.6-1.4-1.2-1.0

Par

tial

Cu

rre

nt

De

nsi

ty C

O (

mA

/cm

2)

Cathode vs Ag/AgCl (V)

AgDAT/CAgPz/CAgPc/CAgTPA/CAgAg/C

0

500

1,000

1,500

2,000

-2.0-1.8-1.6-1.4-1.2-1.0

Par

tial

Cu

rre

nt

De

nsi

ty C

O (

mA

/mg

Ag)

Cathode vs Ag/AgCl (V)

AgDAT/CAgPz/CAgPc/CAgTPA/CAgAg/C

0

5

10

15

20

25

30

35

40

-2.0-1.8-1.6-1.4-1.2-1.0

Par

tial

Cu

rre

nt

De

nsi

ty H

2(m

A/c

m2)

Cathode vs Ag/AgCl (V)

-2.5

-2.3

-2.1

-1.9

-1.7

-1.5

-1.3

-1.1

0 25 50 75 100 125

Cat

ho

de

Po

ten

tial

(V v

s A

g/A

gCl)

Current Density (mA/cm2)

(a)

(d)(c)

(b)

Page 91: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

82

increase in the following order: DAT<Pz<Ag<Pc<Ag/C<TPA. The large overpotential for

Figure 5.9: (a), (c), and (e) SEM images of the electrode surface before application in the flow reactor. (b) and (d)

SEM images of the electrode surface after application in the flow reactor with a 1 M KOH electrolyte flowed over

the electrode at 0.5 ml/min and a potential of 3 V applied across the reactor. The following magnifications were

used (a) 50,000×, (b) 100,000×, (c) 10,000×, (e) 10,000×, and (e) 10,000×.

(a) (b)

(c) (d)

(e)

Page 92: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

83

AgTPA explains why both the hydrogen evolution and the CO2 reduction reactions are slow: the

overall kinetics for AgTPA are quite slow relative to the other catalysts. The overall kinetics are

best for AgDAT and AgPz as is evidenced by their low cathode overpotentials.

5.3.4 Imaging of the Catalyst Surface

XPS and SEM analysis enables us to better understand the charge and surface structure of

silver-based organometallic catalysts. Figure 5.9 shows SEM images of a Sigracet 35BC

electrode with AgDAT/C painted on the electrode at a loading of 1 mg/cm2 both before operation

in a flow cell and after operation in a flow cell. The SEM images shown in Figure 5.9 provide

several very interesting findings. First, as seen in Figure 5.9a and b, the particle size of the

catalyst on the surface is roughly 50 microns, which is quite similar to what is observed for the

Ag nanoparticles used earlier in this study. The similar size of the nanoparticles suggests that the

difference between the Ag and AgDAT samples is not a result of a change in particle size.

Second, at a magnification of 10,000×, 50,000× or 100,000×, the catalyst particles on the surface

do not exhibit significant agglomeration after a potential is applied (comparing Figure 5.9 insets

(a) and (c) to (b) and (d)), which suggests the silver particles are not free to agglomerate. Third,

before using the AgDAT electrode in the reactor, crystals can be observed on the electrode

surface. This can be seen in Figure 5.9c and e. However, after flowing a 1 M KOH electrolyte

over the cathode and applying 3 V across the reactor, the crystals go away. SEM elemental

analysis showed that the composition of the crystals was high in Ag, C, H and N, thereby

suggesting that some of the DAT crystallized on the surface and was washed away as DAT has

high solubility by itself in water. However, from Figure 5.9c, it can be seen that the majority of

the surface is covered by nanoparticles, rather than crystals. The crystals are believed to be rich

in excess triazole components which wash away when the reactor is used.

Page 93: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

84

XPS provides insight regarding the charge of the silver on the surface. By looking at the

electrode surface we can see that the silver goes from a Ag+ to Ag after being used in the flow

cell. This suggests that silver on the surface is being reduced, presumably after a potential of 3 V

is applied across the reactor.

5.3.5 Confirmation of Amine Activity

The XPS data showing the transition of the charge on the silver from Ag+ to Ag suggests

that the silver-amine complex for AgDAAT does not stay intact. To confirm the activity of the

Amine, two experiments were performed.

First, 3,5-diamino-1,2,4-triazole (DAT) was added to a Ag/C catalyst solution, which was

shown to not be active for CO2 reduction in Sections 5.3.1, 5.3.2, and 5.3.3, and sonicated before

being hand-painted on a Sigracet 35BC electrode, and tested for CO2 reduction activity in a flow

cell. Specifically, DAT was added to the Ag/C catalysts solution such that the DAT to Ag/C

ratio was 30 to 100 on a per mass basis. In Figure 5.10, the performance of an electrode with

Ag/C catalyst was compared to the performance of an electrode with Ag/C catalyst combined

Figure 5.10: (a) Faradaic efficiency, and (b) partial current density for CO production with a Ag/C electrode as

compared to a Ag/C electrode with the addition of DAT to the catalyst mixture before hand-painting the catalyst on

the Sigercet 35BC electrode.

0

5

10

15

20

25

30

-3.25-3.00-2.75-2.50-2.25-2.00-1.75

Par

tial

Cu

rre

nt

De

nsi

ty fo

r C

O (

mA

)

Cathode (vs Ag/AgCl)

Ag/C+DAT

Ag/C

0

10

20

30

40

50

60

70

80

90

100

-3.25-3.00-2.75-2.50-2.25-2.00-1.75

Fara

dai

c Ef

fici

en

cy (%

)

Cell Potential (V)

Ag/C + DAT

Ag/C

(a) (b)

Page 94: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

85

with DAT. The addition of DAT to the catalyst mixture prior to painting drastically improved

the selectivity of the catalyst for CO as opposed to H2; specifically, the partial current density for

CO increased by more than 20 fold. This suggests that the amine may act as a co-catalyst on the

electrode surface, thereby driving CO production.

Secondly, the activity of the Amine was explored without any silver. Five milligrams of

3,5-diamino-1,2,4-triazole was mixed with 6.5 µl of Nafion in a 1:1 H2O to 2-propanol solution,

sonicated and hand-painted on a Sigracet 35BC electrode. When the electrolyte flowed over the

reactor, some of the DAT was presumed to have come off into solution because the effluent

electrolyte had a yellow tint. After flowing the electrolyte through the reactor for several

minutes, the yellow tint went away and the reactor was tested for CO2 reduction activity.

Figure 5.11 compares the activity of a bare Sigracet 35BC electrode in the flow reactor as

compared to the activity of a Sigracet 35BC electrode with DAT in the flow reactor. The

addition of DAT enabled a small, yet significant, amount of CO production, thus showing that

the DAT is active for CO2 reduction, however, with very poor overall current densities. The

observation that DAT is active for CO2 reduction confirms the theory that the amine is involved

Figure 5.11: The partial current densities for CO and H2 production in the flow cell with (a) DAT painted on a

Sigercet 35BC electrode, and (b) a bare Sigercet 35BC electrode.

0

10

20

30

40

50

60

-2.0-1.8-1.6-1.4-1.2

Part

ial C

urr

en

t D

en

sit

y (

mA

/cm

2)

Cathode Potential (vs Ag/AgCl)

CO

H2

0

10

20

30

40

50

60

-2.0-1.8-1.6-1.4-1.2

Pari

tal C

urr

en

t D

en

sit

y (

mA

/cm

2)

Cathode Potential (vs AgAgCl)

CO

H2

(a) (b)

Page 95: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

86

the CO2 reduction reaction. The low

current densities with the Amine suggest

that it may be acting as a co-catalyst

because the amine alone is unlikely to

achieve high current densities.

5.3.6 pH Perturbation

In Figure 5.12, the influence of

pH is investigated on individual

electrode polarization and partial current

densities for the electrochemical

reduction of CO2 to CO on a AgDAT

cathode catalyst and a Pt anode catalyst

with a 0.5 mL/min 1 M KOH electrolyte

flow rate. Similar to what was seen in

Chapter 2, basic conditions are the most

favorable for CO production.

As shown in Figure 5.12a, the

CO production is unaffected by basic

conditions. However, with acidic

conditions, the partial current density for

CO decreases. Specifically, starting at

pH 4, the partial current density for CO

begins to decrease, and the trend Figure 5.12: Partial current densities for (a) CO, and (b) H2 as well

as individual electrode plots for electrolyte pH solutions between 0

and 14 flowed at 0.5 mL/min in a flow reactor.

0

20

40

60

80

100

120

-2.0-1.8-1.6-1.4-1.2-1.0

Par

tial

Cu

rre

nt

De

nsi

ty C

O (

mA

/cm

2)

Cathode (vs Ag/AgCl)

pH 14

pH 10

pH 4

pH 2

pH 0

0

1

2

3

4

5

6

-2.0-1.8-1.6-1.4-1.2-1.0

Par

tial

Cu

rre

nt

De

nsi

ty H

2(m

A/c

m2 )

Cathode vs Ag/AgCl (V)

0

20

40

60

80

100

120

140

160

180

-2.0-1.5-1.0-0.50.0

Par

tial

Cu

rre

nt

De

nsi

ty H

2(m

A/c

m2)

Cathode (vs Ag/AgCl)

0

1

2

3

4

5

6

-2.0-1.8-1.6-1.4-1.2-1.0

Par

tial

Cu

rre

nt

De

nsi

ty H

2(m

A/c

m2 )

Cathode (vs Ag/AgCl)

-2.0

-1.5

-1.0

-0.5

0.0

0.5

1.0

1.5

2.0

2.5

0 25 50 75 100 125

Cat

ho

de

Po

ten

tial

(vs

Ag/

AgC

l)

Current Density (mA/cm2)

(a)

(b)

(c)

Page 96: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

87

continues as the electrolyte solution

becomes more acidic (i.e., pH 2 and

pH 0).

In Figure 5.12b, the pH can be

seen to drastically effects H2 evolution

on the cathode, and consequently,

change the selectivity for CO to H2. In

acidic conditions, H2 evolution

drastically increases because the cathode

is flooded with excess protons, which

drive the evolution of H2 as opposed to the reduction of CO2 to CO. This is confirmed by the

decrease in cathode overpotential under acidic conditions as seen in Figure 5.12c. Furthemore,

neutral conditions experience “pH sheltering” as the electrode surface pH will be sheltered from

the electrolyte solution pH because the anode and cathode reactions produce and consume

protons [35-37].

The anode is also influenced by pH as shown in Figure 5.12c. Specifically, the anode

overpotential is decreased in basic conditions. Consequently, basic conditions are the most

beneficial for CO2 reduction because smallest overall cell potential is required.

5.3.7 Copper Organometallic Complexes

Figure 5.13 shows the cathode polarization curves for four copper-based organometallic

catalysts compared to a copper catalyst, all on electrodes with a 1 mg/cm2 catalyst loading in an

electrochemical flow reactor with a 1 M KOH electrolyte flowed at 0.5 mL/min and a CO2

Figure 5.13: Cathode polarization curves for a copper and four

copper-based organometallic catalysts on electrodes with a

1 mg/cm2 catalyst loading in an electrochemical flow reactor

with a 1 M KOH electrolyte flowed at 0.5 mL/min and a CO2

reactant stream flowed behind the cathode at 5 SCCM.

-2.5

-2.3

-2.1

-1.9

-1.7

-1.5

-1.3

-1.1

-0.9

0 25 50 75 100 125

Cat

ho

de

Po

ten

tial

(V v

s A

g/A

gCl)

Current Density (mA/cm2)

301 21

40 289

Cu

-2.5

-2.3

-2.1

-1.9

-1.7

-1.5

-1.3

-1.1

-0.9

0 25 50 75 100 125

Cat

ho

de

Po

ten

tial

(V v

s A

g/A

gCl)

Current Density (mA/cm2)

301 21

40 289

Cu

Page 97: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

88

reactant stream flowed behind the cathode at 5 SCCM. The overpotentials for the

organometallic complexes are much higher than that for the lone copper catalyst.

Figure 5.14 shows the product distribution for the same experiment as Figure 5.13. The

partial current densities for H2, formic acid, CO, and CH4 are much lower with the

organometallic complexes as compared to the copper electrode. Furthermore, the onset

potentials for many of the carbon products are delayed. Specifically, the onset potential for

methane is much later with the organometallic complexes as opposed to the copper electrode.

This same trend is seen with carbon monoxide and formic acid, but to a lesser extent. Because of

the later onset potential for the carbon products, the faradaic efficiency for H2, generally the

Figure 5.14: Product selectivity for (a) H2, (b) formic acid, (c) CO, and (d) CH4 for a copper and four copper-based

organometallic catalysts on electrodes with a 1 mg/cm2 catalyst loading in an electrochemical flow reactor.

0

10

20

30

40

50

60

70

-2.3-2.1-1.9-1.7-1.5-1.3-1.1

Par

tial

Cu

rre

nt

De

nsi

ty H

2(m

A/c

m2 )

Cathode vs Ag/AgCl (V)

301

21

40

289

Cu

0

10

20

30

40

50

60

-2.3-2.1-1.9-1.7-1.5-1.3-1.1

Par

tial

Cu

rre

nt

De

nsi

ty C

O (

mA

/cm

2 )

Cathode vs Ag/AgCl (V)

301

21

40

289

Cu

0

5

10

15

20

25

30

35

40

45

-2.3-2.1-1.9-1.7-1.5-1.3-1.1

Par

tial

Cu

rre

nt

De

nsi

ty F

A (

mA

/cm

2 )

Cathode vs Ag/AgCl (V)

301

21

40

289

Cu

0

1

2

3

4

5

6

7

8

9

10

-2.3-2.1-1.9-1.7-1.5-1.3-1.1

Par

tial

Cu

rre

nt

De

nsi

ty C

H4

(mA

/cm

2 )

Cathode vs Ag/AgCl (V)

301

21

40

289

Cu

(a)

(d)(c)

(b)

0

10

20

30

40

50

60

70

-2.3-2.1-1.9-1.7-1.5-1.3-1.1

Par

tial

Cu

rre

nt

De

nsi

ty H

2(m

A/c

m2 )

Cathode vs Ag/AgCl (V)

301

21

40

289

Cu

0

10

20

30

40

50

60

-2.3-2.1-1.9-1.7-1.5-1.3-1.1

Par

tial

Cu

rre

nt

De

nsi

ty C

O (

mA

/cm

2 )

Cathode vs Ag/AgCl (V)

301

21

40

289

Cu

0

5

10

15

20

25

30

35

40

45

-2.3-2.1-1.9-1.7-1.5-1.3-1.1

Par

tial

Cu

rre

nt

De

nsi

ty F

A (

mA

/cm

2 )

Cathode vs Ag/AgCl (V)

301

21

40

289

Cu

0

1

2

3

4

5

6

7

8

9

10

-2.3-2.1-1.9-1.7-1.5-1.3-1.1

Par

tial

Cu

rre

nt

De

nsi

ty C

H4

(mA

/cm

2 )

Cathode vs Ag/AgCl (V)

301

21

40

289

Cu

(a)

(d)(c)

(b)

0

10

20

30

40

50

60

70

-2.3-2.1-1.9-1.7-1.5-1.3-1.1

Par

tial

Cu

rre

nt

De

nsi

ty H

2(m

A/c

m2 )

Cathode vs Ag/AgCl (V)

301

21

40

289

Cu

0

10

20

30

40

50

60

-2.3-2.1-1.9-1.7-1.5-1.3-1.1

Par

tial

Cu

rre

nt

De

nsi

ty C

O (

mA

/cm

2 )

Cathode vs Ag/AgCl (V)

301

21

40

289

Cu

0

5

10

15

20

25

30

35

40

45

-2.3-2.1-1.9-1.7-1.5-1.3-1.1

Par

tial

Cu

rre

nt

De

nsi

ty F

A (

mA

/cm

2 )

Cathode vs Ag/AgCl (V)

301

21

40

289

Cu

0

1

2

3

4

5

6

7

8

9

10

-2.3-2.1-1.9-1.7-1.5-1.3-1.1

Par

tial

Cu

rre

nt

De

nsi

ty C

H4

(mA

/cm

2 )

Cathode vs Ag/AgCl (V)

301

21

40

289

Cu

(a)

(d)(c)

(b)

0

10

20

30

40

50

60

70

-2.3-2.1-1.9-1.7-1.5-1.3-1.1

Par

tial

Cu

rre

nt

De

nsi

ty H

2(m

A/c

m2 )

Cathode vs Ag/AgCl (V)

301

21

40

289

Cu

0

10

20

30

40

50

60

-2.3-2.1-1.9-1.7-1.5-1.3-1.1

Par

tial

Cu

rre

nt

De

nsi

ty C

O (

mA

/cm

2 )

Cathode vs Ag/AgCl (V)

301

21

40

289

Cu

0

5

10

15

20

25

30

35

40

45

-2.3-2.1-1.9-1.7-1.5-1.3-1.1

Par

tial

Cu

rre

nt

De

nsi

ty F

A (

mA

/cm

2 )

Cathode vs Ag/AgCl (V)

301

21

40

289

Cu

0

1

2

3

4

5

6

7

8

9

10

-2.3-2.1-1.9-1.7-1.5-1.3-1.1

Par

tial

Cu

rre

nt

De

nsi

ty C

H4

(mA

/cm

2 )

Cathode vs Ag/AgCl (V)

301

21

40

289

Cu

(a)

(d)(c)

(b)

0

10

20

30

40

50

60

70

-2.3-2.1-1.9-1.7-1.5-1.3-1.1

Par

tial

Cu

rre

nt

De

nsi

ty H

2(m

A/c

m2 )

Cathode vs Ag/AgCl (V)

301

21

40

289

Cu

0

10

20

30

40

50

60

-2.3-2.1-1.9-1.7-1.5-1.3-1.1

Par

tial

Cu

rre

nt

De

nsi

ty C

O (

mA

/cm

2 )

Cathode vs Ag/AgCl (V)

301

21

40

289

Cu

0

5

10

15

20

25

30

35

40

45

-2.3-2.1-1.9-1.7-1.5-1.3-1.1

Par

tial

Cu

rre

nt

De

nsi

ty F

A (

mA

/cm

2 )

Cathode vs Ag/AgCl (V)

301

21

40

289

Cu

0

1

2

3

4

5

6

7

8

9

10

-2.3-2.1-1.9-1.7-1.5-1.3-1.1

Par

tial

Cu

rre

nt

De

nsi

ty C

H4

(mA

/cm

2 )

Cathode vs Ag/AgCl (V)

301

21

40

289

Cu

(a)

(d)(c)

(b)

0

10

20

30

40

50

60

70

-2.3-2.1-1.9-1.7-1.5-1.3-1.1

Par

tial

Cu

rre

nt

De

nsi

ty H

2(m

A/c

m2 )

Cathode vs Ag/AgCl (V)

301

21

40

289

Cu

0

10

20

30

40

50

60

-2.3-2.1-1.9-1.7-1.5-1.3-1.1

Par

tial

Cu

rre

nt

De

nsi

ty C

O (

mA

/cm

2 )

Cathode vs Ag/AgCl (V)

301

21

40

289

Cu

0

5

10

15

20

25

30

35

40

45

-2.3-2.1-1.9-1.7-1.5-1.3-1.1

Par

tial

Cu

rre

nt

De

nsi

ty F

A (

mA

/cm

2 )

Cathode vs Ag/AgCl (V)

301

21

40

289

Cu

0

1

2

3

4

5

6

7

8

9

10

-2.3-2.1-1.9-1.7-1.5-1.3-1.1

Par

tial

Cu

rre

nt

De

nsi

ty C

H4

(mA

/cm

2 )

Cathode vs Ag/AgCl (V)

301

21

40

289

Cu

(a)

(d)(c)

(b)

0

10

20

30

40

50

60

70

-2.3-2.1-1.9-1.7-1.5-1.3-1.1

Par

tial

Cu

rre

nt

De

nsi

ty H

2(m

A/c

m2 )

Cathode vs Ag/AgCl (V)

301

21

40

289

Cu

0

10

20

30

40

50

60

-2.3-2.1-1.9-1.7-1.5-1.3-1.1

Par

tial

Cu

rre

nt

De

nsi

ty C

O (

mA

/cm

2 )

Cathode vs Ag/AgCl (V)

301

21

40

289

Cu

0

5

10

15

20

25

30

35

40

45

-2.3-2.1-1.9-1.7-1.5-1.3-1.1

Par

tial

Cu

rre

nt

De

nsi

ty F

A (

mA

/cm

2 )

Cathode vs Ag/AgCl (V)

301

21

40

289

Cu

0

1

2

3

4

5

6

7

8

9

10

-2.3-2.1-1.9-1.7-1.5-1.3-1.1

Par

tial

Cu

rre

nt

De

nsi

ty C

H4

(mA

/cm

2 )

Cathode vs Ag/AgCl (V)

301

21

40

289

Cu

(a)

(d)(c)

(b)

Page 98: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

89

undesired product, is higher with the organometallic complexes. While the organometallic

complexes negatively affect the reaction, specifically by exhibiting a lower current density, a

later onset for carbon products, and a higher selectivity for H2, our observation indicates that

organometallic catalysts can potentially be “tuned” with respect to product selectivity.

5.4 Conclusions

Amines have previously been shown to influence both product selectivity and kinetics for

the electrochemical reduction of CO2 to CO via incorporating amines in the electrolyte solution.

Here, we report that development of amine based novel organometallic complexes for the

electrochemical reduction of CO2 to CO, which reduce the onset potential by ~0.1V and increase

the current densities as compared to a more traditional catalyst, Ag. Specifically, AgDAT,

AgPc, and AgPz have comparable or better current densities as compared to Ag with increased

selectivity for CO. Furthermore, when comparing the performance relative to the silver loading,

the amine based organometallic catalysts outperform the silver catalyst by roughly 20 fold.

A major obstacle to the broader application of the electrochemical reduction of CO2 lies

in the challenge of simultaneously achieving high current densities and energetic efficiencies.

Here we demonstrate a major step towards improving both the energetic efficiencies and current

densities of the electrochemical reduction of CO2 to CO on a Ag cathode. In particular, we

observe a drastic improvement in the current densities achieved on a per-metal loading.

Furthermore, we demonstrate the potential to influence the selectivity of other catalysts.

Specifically, CuDAT and other amines influence the current density and product selectivity for a

copper catalyst, which makes a broad range of products. This work demonstrates that

organometallic complexes can be used to “tune” catalyst selection for the desired products.

Page 99: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

90

The work reported here suggest that further exploration of the effect of amine-based

catalysis on the electrochemical conversion of CO2 to CO, or to other products, may lead to

further improvements in performance.

5.5 References

1. Brian A. Rosen, Amin Salehi-Khojin, Michael R. Thorson, Wei Zhu, Devin T. Whipple,

Paul J. A. Kenis and Richard I. Masel, Ionic Liquid–Mediated Selective Conversion of

CO2 to CO at Low Overpotentials, Science, 2011, 334 (6056), p. 643-644.

2. D. T. Whipple and P. J. A. Kenis, Prospects of CO2 Utilization via Direct Heterogeneous

Electrochemical Reduction, Journal of Physical Chemistry Letters, 2010, 1 (24), p. 3451-

3458.

3. Y. Hori, Electrochemical CO2 Reduction on Metal Electrodes, Modern Aspects of

Electrochemistry, 2008, 42 p. 89-189.

4. M. R. Dubois and D. L. Dubois, Development of Molecular Electrocatalysts for CO2

Reduction and H2 Production/Oxidation, Accounts of Chemical Research, 2009, 42 (12),

p. 1974-1982.

5. Narasi Sridhar, Carbon Dioxide Utilization: Electrochemical Conversion of CO2 -

Opportunities and Challenges. Det Norske Veritas, 2011. 07 1-20.

6. A. T. Bell, Basic Energy Needs: Catalysis for Energy. Department of Energy, 2008.

PNNL17214

7. G. Centi and S. Perathoner, Opportunities and Prospects in the Chemical Recycling of

Carbon Dioxide to Fuels, Catalysis Today, 2009, 148 (3-4), p. 191-205.

8. C. S. Song, Global Challenges and Strategies for Control, Conversion and Utilization of

CO2 for Sustainable Development Involving Energy, Catalysis, Adsorption and Chemical

Processing, Catalysis Today, 2006, 115 (1-4), p. 2-32.

9. Yoshio Hori, Hidetoshi Wakebe, Toshio Tsukamoto and Osamu Koga, Electrocatalytic

process of CO selectivity in electrochemical reduction of CO2 at metal electrodes in

aqueous media, Electrochimica Acta, 1994, 39 (11–12), p. 1833-1839.

10. Masashi Azuma, Kazuhito Hashimoto, Masahiro Hiramoto, Masahiro Watanabe and

Tadayoshi Sakata, Electrochemical Reduction of Carbon Dioxide on Various Metal

Electrodes in Low-Temperature Aqueous KHCO[sub 3] Media, Journal of the

Electrochemical Society, 1990, 137 (6), p. 1772-1778.

11. Nagakazu Furuya, Tsuyoshi Yamazaki and Masami Shibata, High performance RuPd

catalysts for CO2 reduction at gas-diffusion electrodes, Journal of Electroanalytical

Chemistry, 1997, 431 (1), p. 39-41.

12. Matthew S Thorum, Jessica Yadav and Andrew A Gewirth, Oxygen Reduction Activity of

a Copper Complex of 3,5-Diamino-1,2,4-triazole Supported on Carbon Black,

Angewandte Chemie International Edition, 2009, 48 (1), p. 165-167.

13. Fikile R. Brushett, Matthew S. Thorum, Nicholas S. Lioutas, Matthew S. Naughton,

Claire Tornow, Huei-Ru “Molly” Jhong, Andrew A. Gewirth and Paul J. A. Kenis, A

Carbon-Supported Copper Complex of 3,5-Diamino-1,2,4-triazole as a Cathode Catalyst

Page 100: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

91

for Alkaline Fuel Cell Applications, Journal of the American Chemical Society, 2010,

132 (35), p. 12185-12187.

14. J. O. Bockris and J. C. Wass, The Photocatalytic Reduction of Carbon-Dioxide, Journal

of the Electrochemical Society, 1989, 136 (9), p. 2521-2528.

15. J.O.M. Bockris and A.K.N. Reddy, Modern electrochemistry, Plenum Press, 1998.

16. K. Chandrasekaran and L. O'M Bockris, In-situ spectroscopic investigation of adsorbed

intermediate radicals in electrochemical reactions: CO2− on platinum, Surface Science,

1987, 185 (3), p. 495-514.

17. E. B. Cole, P. S. Lakkaraju, D. M. Rampulla, A. J. Morris, E. Abelev and A. B. Bocarsly,

Using a One-Electron Shuttle for the Multielectron Reduction of CO2 to Methanol:

Kinetic, Mechanistic, and Structural Insights, Journal of the American Chemical Society,

2010, 132 (33), p. 11539-11551.

18. A. J. Morris, R. T. McGibbon and A. B. Bocarsly, Electrocatalytic Carbon Dioxide

Activation: The Rate-Determining Step of Pyridinium-Catalyzed CO2 Reduction,

Chemsuschem, 2011, 4 (2), p. 191-196.

19. G. Seshadri, C. Lin and A. B. Bocarsly, A New Homogeneous Electrocatalyst for the

Reduction of Carbon-Dioxide to Methanol at Low Overpotential, Journal of

Electroanalytical Chemistry, 1994, 372 (1-2), p. 145-150.

20. C. Delacourt, P. L. Ridgway, J. B. Kerr and J. Newman, Design of an Electrochemical

Cell Making Syngas (CO+H2) from CO2 and H2O Reduction at Room Temperature,

Journal of the Electrochemical Society, 2008, 155 (1), p. B42-B49.

21. S. Kaneco, H. Katsumata, T. Suzuki and K. Ohta, Electrochemical Reduction of CO2 to

Methane at the Cu Electrode in Methanol with Sodium Supporting Salts and its

Comparison with Other Alkaline Salts, Energy Fuels, 2006, 20 (1), p. 409-414.

22. R. L. Cook, R. C. Macduff and A. F. Sammells, High-Rate Gas-Phase CO2 Reduction to

Ethylene and Methane Using Gas-Diffusion Electrodes, Journal of the Electrochemical

Society, 1990, 137 (2), p. 607-608.

23. N. Furuya and S. Koide, Electroreduction of Carbon-Dioxide by Metal Phthalocyanines,

Electrochimica Acta, 1991, 36 (8), p. 1309-1313.

24. T. Yamamoto, D. A. Tryk, A. Fujishima and H. Ohata, Production of Syngas Plus

Oxygen from CO2 in a Gas-Diffusion Electrode-Based Electrolytic Cell, Electrochimica

Acta, 2002, 47 (20), p. 3327-3334.

25. S. Kaneco, K. Iiba, H. Katsumata, T. Suzuki and K. Ohta, Electrochemical Reduction of

High Pressure CO2 at a Cu Electrode in Cold Methanol, Electrochimica Acta, 2006, 51

(23), p. 4880-4885.

26. S. Kaneco, K. Iiba, H. Katsumata, T. Suzuki and K. Ohta, Electrochemical Reduction of

High Pressure Carbon Dioxide at a Cu Electrode in Cold Methanol with CsOH

Supporting Salt, Chemical Engineering Journal, 2007, 128 (1), p. 47-50.

27. S. Kaneco, R. Iwao, K. Iiba, K. Ohta and T. Mizuno, Electrochemical Conversion of

Carbon Dioxide to Formic Acid on Pb in KOH/Methanol Electrolyte at Ambient

Temperature and Pressure, Energy, 1998, 23 (12), p. 1107-1112.

Page 101: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

92

28. Kendra P. Kuhl, Etosha R. Cave, David N. Abram and Thomas F. Jaramillo, New insights

into the electrochemical reduction of carbon dioxide on metallic copper surfaces, Energy

& Environmental Science, 2012, p. in press.

29. William J. Durand, Andrew A. Peterson, Felix Studt, Frank Abild-Pedersen and Jens K.

Nørskov, Structure effects on the energetics of the electrochemical reduction of CO2 by

copper surfaces, Surface Science, 2011, 605 (15–16), p. 1354-1359.

30. Andrew A. Peterson, Frank Abild-Pedersen, Felix Studt, Jan Rossmeisl and Jens K.

Norskov, How copper catalyzes the electroreduction of carbon dioxide into hydrocarbon

fuels, Energy & Environmental Science, 2010, 3 (9), p. 1311-1315.

31. Andrew A. Peterson and Jens K. Nørskov, Activity Descriptors for CO2 Electroreduction

to Methane on Transition-Metal Catalysts, The Journal of Physical Chemistry Letters,

2012, 3 (2), p. 251-258.

32. W. Tang, A. A. Peterson, A. S. Varela, Z. P. Jovanov, L. Bech, W. J. Durand, S. Dahl, J.

K. Norskov and I. Chorkendorff, The importance of surface morphology in controlling

the selectivity of polycrystalline copper for CO2 electroreduction, Physical chemistry

chemical physics : PCCP, 2012, 14 (1), p. 76-81.

33. Norberto Masciocchi, Massimo Moret, Paolo Cairati, Angelo Sironi, G. Attilio Ardizzoia

and Girolamo La Monica, The Multiphase Nature of the Cu(pz) and Ag(pz) (Hpz =

Pyrazole) Systems: Selective Syntheses and Ab-Initio X-ray Powder Diffraction

Structural Characterization of Copper(I) and Silver(I) Pyrazolates, Journal of the

American Chemical Society, 1994, 116 (17), p. 7668-7676.

34. R. Sleat and R. A. Mah, Quantitative Method for Colorimetric Determination of Formate

in Fermentation Media, Applied and Environmental Microbiology, 1984, 47 (4), p. 884-

885.

35. D. T. Whipple, E. C. Finke and P. J. A. Kenis, Microfluidic Reactor for the

Electrochemical Reduction of Carbon Dioxide: The Effect of pH, Electrochemical and

Solid State Letters, 2010, 13 (9), p. D109-D111.

36. H. Li and C. Oloman, Development of a Continuous Reactor for the Electro-reduction of

Carbon Dioxide to Formate - Part 2: Scale-up, Journal of Applied Electrochemistry,

2007, 37 (10), p. 1107-1117.

37. M. N. Mahmood, D. Masheder and C. J. Harty, Use of Gas-Diffusion Electrodes for

High-Rate Electrochemical Reduction of Carbon-Dioxide .1. Reduction at Lead, Indium-

Impregnated and Tin-Impregnated Electrodes, Journal of Applied Electrochemistry,

1987, 17 (6), p. 1159-1170.

Page 102: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

93

Chapter 6

Design Rules for Electrode Arrangement in an

Air-breathing Alkaline Direct Methanol LFFC1

6.1 Introduction

An ever increasing demand for high power density, portable electronic devices (e.g.,

smartphones, laptops) with a fast recharge has motivated the development of many micro-scale

fuel cells [1-5]. Small-scale fuel cells demonstrate superior energy densities compared to

rechargeable batteries; offering smaller, and lighter alternatives for applications requiring

portable power sources [6]. Specifically, developments in both novel fuel cell catalysts and

electrode assemblies, and advances in fabrication have enabled the miniaturization of fuel cells

capable of integration in small portable applications [2-4,7,8]. Furthermore, liquid organic feed

sources, such as methanol, have a distinct advantage compared to gaseous hydrogen feed sources

as the fuel can be stored safely at low pressure, in an energy dense form. In comparison to

traditional batteries, the liquid nature of methanol also enables nearly instantaneous recharging

via refilling or changing-out of a fuel cartridge. Additionally, the use of a membraneless laminar

flow fuel cell (LFFC) enables the use of alkaline media, avoiding the membrane-associated

issues typically encountered in alkaline fuel cells. Furthermore, in alkaline media, the reaction

kinetics are better at both the anode [9] and the cathode [10], and high performance can be

achieved with abundant and cheap catalysts such as Ag [11-15].

LFFCs have several further attractive characteristics. The flowing liquid electrolyte

stream mitigates water management issues and enables fuel flexibility [16]. Alternative fuel feed

1 Part of this work has been published: Thorson, M. R, Brushett, F. B., Kenis, P. J. A., “Design Rules for Electrode

Arrangement in an Air-breathing Alkaline Direct Methanol LFFC.” Journal of Power Sources 2012

Page 103: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

94

mechanisms [17,18], as well as modifications to electrode location and method of catalyst

deposition [19-23] have led to improved current densities. Alternative cell designs [24-26] (e.g.,

F, Y, and T channel geometries) and further modifications to LFFC configurations [27] (e.g.,

herringbone mixers, multiple inlets) have improved both fuel conversion and current density,

while helping to mitigate fuel crossover. Recently, nanoporous separators have been employed

to minimize the interfacial area between the fuel and electrolyte streams, thereby drastically

reducing fuel crossover [28].

A major limitation to obtaining higher current densities in present LFFC configurations is

the depletion of fuel along the electrode surface. Several computational studies have reported on

this boundary layer depletion issue. Specifically, using a CFD model of the formic acid reaction

within a membraneless LFFC, Bazylak et al. showed that a ~15 µm thick boundary layer formed

within 1 mm, which drastically increased the mass transport limitations [29]. Yoon, et al.

modeled a similar scenario to examine the electrode performance for formic acid oxidation with

a membraneless LFFC [27] in which he showed that depletion quickly becomes a major

limitation within the fast forming boundary layer and that the current density will decrease by a

factor of 4 within the first centimeter of the electrode length. Additionally, Khabbazi et al.

modeled the influence of electrode length (8 to 30 mm) on electrode performance and showed

that the power density could be increased by ~80 % by reducing the electrode length [30]. These

aforementioned models all assume well developed flow profiles with smooth electrodes. Actual

electrode surfaces will have inherent roughness, which will influence the diffusion of fuel to

reactive sites on the electrode. Moreover, the local geometries and the orientation of the inlets

for the electrolyte and fuel streams will affect the formation of the boundary layer on the fuel

stream and thereby, change the influence of electrode length on performance.

Page 104: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

95

Here, we experimentally studied the influence of electrode length on performance in an

air-breathing alkaline direct methanol LFFC with the aim of formulating design rules for

multichannel LFFCs with optimum specific energy. The effects of aspect ratio (length-to-width)

on electrode performance, and thus overall fuel cell performance, were investigated.

Furthermore, similar experiments were used to develop an empirical model, which predicts the

optimal electrode configurations for a multichannel LFFC.

6.2 Experimental

6.2.1 Fuel Cell Assembly

Fig. 6.1 shows the LFFC we used in this study, a configuration that is similar to those we

have reported on previously [28,31]. Instead of graphite, we used two stainless steel backing

plates (6 x 3 cm2) as current collectors. We chose to use stainless steel plates backing plates to

improve the structural strength (prevent cracking of the graphite), reduce contact resistance

between the wires from the potentiostat and the backing plate, and improve the pressure

distribution of the bolts. In agreement with prior work [32], we did not observe corrosion of the

steel backing plates of our setup. Wires with banana clips plugged directly into precision-

machined 3/16” holes in the anode and cathode current collectors connected the cell with an in-

house built load box. The cathode current collector had a precision-machined 3-mm wide by 40-

mm long window that enabled diffusion of oxygen from air to the cathode for an air-breathing

configuration. Two 1-mm thick poly(methyl methacrylate) (PMMA) sheets with precision

machined 3-mm wide by 50-mm long windows served as the the channels for the electrolyte

flow field within the fuel cell. We used Teflon sheets with precision cut windows (3-mm wide

and 2.25, 4.5, 9, 18, or 36-mm long), placed between the electrode and the flow channel, to vary

the area of the cathode and anode exposed to air and fuel, respectively. A polycarbonate track-

Page 105: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

96

etch membrane (6-µm thick, 0.05-µm pore size, PCT0059030 Sterlitech Corporation) was placed

between the two PMMA flow channels to reduce the liquid-liquid interfacial area thereby

reducing the effects of fuel crossover [28]. The cell was held together by four insulated bolts

(McMaster Carr).

6.2.2 Electrode Preparation

The anode was prepared via painting a 1:1 (V/V) H2O:2-propanol solution containing

dispersed catalyst onto the hydrophobized side of a sheet of Toray carbon paper (TGP-H-120,

Fuel Cell Earth) such that the final catalyst loading was 2 mg cm-2

of unsupported Pt/Ru alloy

catalyst (50:50 wt %, Alfa Aesar) and 0.1 mg cm-2

of Nafion binder (LIQUION LG-1105, Ion

Power). Similarly, the anode was prepared via painting a catalyst solution on Toray carbon

paper resulting in a catalyst loading of 2 mg cm-2

of Pt/C catalyst (50 % mass on Vulcan carbon,

E-Tek) and 0.1 mg cm-2

of Nafion binder. More details on this catalyst deposition procedure

have been reported previously [33].

6.2.3 Fuel Cell Testing

The fuel cell assembly was tested under alkaline conditions using an anolyte comprised

of fuel + 1 M potassium hydroxide (KOH, Sigma Aldrich), and a catholyte comprised of 1 M

KOH. The KOH electrolyte solutions are not expected to damage the gas diffusion electrodes

via carbonate poisoning because the application of a flowing electrolyte removes carbonates

from the electrolyte system [34].

Four methanol (Sigma Aldrich) concentrations (0.375, 0.75, 1.5, and 3.0 M) were used in

the anolyte. The flow rates of the fuel and electrolyte streams within the fuel cell were regulated

with a syringe pump (Harvard Instruments, PHD 2200). Fuel cell measurements were conducted

at 0.05 V intervals using an in-house built load box. The current was measured once steady-state

Page 106: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

97

had been achieved at a given potential. The geometric surface areas used to calculate the

reported current and power densities were equal to the exposed electrode areas (0.0625, 0.125,

0.25, 0.5, and 1 cm2 for the 3-mm wide and 2.25, 4.5, 9, 18, 36-mm long Teflon windows).

After flowing through the reactor, the electrolyte stream exited the fuel cell through a plastic tube

(Cole Palmer, i.d. = 1.57 mm) and was collected in a glass beaker. As shown in previous work,

no significant potential drop occurs between the fuel cell and the collection beaker [35].

Individual electrode potentials were measured using multimeters (Fluke 8 III) relative to an

Ag/AgCl reference electrode (saturated NaCl, BAS, West Lafayette, IN) in the collection beaker

for the fuel stream. Fuel cell measurements were conducted at 0.05 V intervals using an in-

house built load box. The current was measured once steady-state had been reached for a given

potential, which typically took between 3 and 5 minutes. The geometric surface areas used to

calculate the reported current and power densities were equal to the exposed electrode areas

(0.0625, 0.125, 0.25, 0.5, and 1 cm2 for the 3-mm wide and 2.25, 4.5, 9, 18, 36-mm long Teflon

windows). All conversions are calculated based on the current recorded by the potentiostat, and

consequently, do not include Faradaic losses due to fuel crossover in the conversion calculation.

6.2.4 Modeling

Based on the fuel conversion in the presence of various length electrodes in a single

LFFC unit, we modeled the anticipated conversion for a fuel cell stack comprised of a series of

electrodes. Several assumptions went into the development of this model. First, performance

losses associated with fuel crossover were not take into account because the use of a separator

drastically reduces the crossover effect as we previously reported [28]. Second, we assumed

that, for the conditions tested, the current density of the fuel cell can be interpolated in a linear

fashion relative to the inlet fuel concentration (see Appendix A). Third, we assumed that fuel

Page 107: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

98

streams between consecutive cells had the opportunity to restore any concentration depletion

layers generated on the electrodes in the previous cell. In other words, a well-mixed fuel stream,

but of lower fuel concentration, will enter each successive cell.

For this model, a full factorial experimental design was developed in which we varied the

methanol concentration of the fuel stream (0.75, 1.5, and 3.0 M), the flow rates of the methanol

and electrolyte streams (0.1, 0.2, 0.4, and 0.8 ml min-1

), and the length of the exposed electrode

(9, 18, and 36 mm). In each experiment, the KOH concentration in the electrolyte and fuel

streams was 1 M. For each condition, a polarization curve was obtained to identify a power

output for a given current density. For each electrode length and flow rate tested, the measured

current was plotted as a function of the three methanol concentration. Then, after interpolation

over the regions 0.75 to 1.5 M and 1.5 to 3.0 M, the slope m of these interpolated lines is

calculated using the following equation:

(6.1)

where i is the current density at respectively the high and low methanol concentration, c, of each

region (so 3.0 and 1.5 M, or 1.5 and 0.75 M). Now equations to calculate an estimate current

density (iest) for any inlet feed concentration (cinlet) in each of the two regions can be obtained:

0.75 < [MeOH] < 1.5 (6.2a)

1.5 < [MeOH] < 3.0 (6.2b)

Using the calculated values for iest, the outlet concentration (coutlet) for an individual cell

can be calculated using:

(6.3)

where F is Faraday’s constant, and Q is the volumetric flow rate of the methanol stream.

These aforementioned equations were used in a repetitive fashion to simulate several

electrodes in series until coutlet, was 0.75 M, which corresponds to a conversion of 75 %.

Page 108: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

99

Specifically, in the empirical model, the inlet methanol concentration, cinlet, for the first electrode

was 3 M. The current from the first electrode, iact, was estimated based on calculations using Eq.

(1) and Eq. (2). The outlet concentration, coutlet, for the first electrode was calculated using Eq.

(3). This process was repeated for additional electrodes with the inlet concentrations for the

subsequent electrodes being defined by the outlet concentration of the previous electrode. Once

the final outlet concentration, coutlet, was below 0.75 M, the total electrode surface area was

calculated by multiplying the number of electrodes required for the given conversion with the

surface area of the given electrode (e.g., 16 electrodes × 1.08 cm2 per electrode = 17.3 cm

2).

6.3 Results and Discussion

6.3.1 Influence of Electrode Length on Performance

The influence of electrode length on power and current output was investigated by

Figure 6.1: Schematic of an alkaline, air-breathing, direct methanol, laminar flow fuel cell (LFFC). Two 1-mm thick

PMMA windows provide the flow fields for the electrolyte and fuel streams between the current collectors and the

nanoporous separator.

MeOH

Electrolyte

Air

Track-Etched Membrane

(porous filter to minimize crossover)

Electrolyte

MeOH

Cathode

Anode

Air Air

Methanol to

Reference Electrode

Electrolyte

Cathode

Air Air Air

Electrolyte

Methanol

Anode

Track-Etched Membrane

Page 109: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

100

changing the exposed electrode length (2.25, 4.5, 9, and 18 mm) of both the anode and cathode

in a direct methanol LFFC (Fig. 6.1). Peak power outputs of 7.5, 4.3, 3.1, and 1.7 mW at current

outputs of 24, 19.2, 14.4, and 8.1 mA were observed for 18, 9, 4.5, and 2.25-mm length

electrodes, respectively (Fig. 6.2a). These results corresponded to peak power, and max current

densities of 13.9, 16.1, 23.1, and 24.9 mW·cm-2

and 45, 71, 107, and 121 mA·cm-2

, respectively

for 18, 9, 4.5, 2.25-mm length electrodes (Fig. 6.2b). While the power output consistently drops

with electrode length, the first eighth (2.25 mm) and quarter (4.5 mm) of the electrode produces

23 % and 41 % of the total power output obtained when using the full 18-mm electrode length,

respectively. Both the maximum current and power densities at electrode lengths of 4.5 and

2.25 mm are similar to or better than the current and power densities typically observed for

Nafion-based DMFCs and miniaturized conventional DMFCs (100 to 150 mA cm-2

and 15 to

20 mW cm-2

, respectively [2,36-39]).

Fig. 6.2c shows the individual electrode polarization plots for the aforementioned

experiments. In general, both anode and cathode performance per unit area improved with

decreasing electrode length due to the reduced boundary layer effects. The 2.25- and 4.5-mm

length anodes had later onsets of mass transport losses, around 150 mA·cm-2

, as compared to the

18- and 9-mm length anodes, which had onsets of mass transport losses in the range of 50 to

75 mA·cm-2

(Fig. 6.2c). However, little difference was observed between the 4.5- and 2.25-mm

length anodes. This observation suggests that other factors may dominate in the first 4.5 mm of

the electrode length. Specifically, catalyst roughness, typically on the order of 10 to 30 μm [40],

may hinder diffusion more than boundary layer depletion for the first 4.5 mm, given that the

boundary layer thickness is on the order of 15 μm [29]. The roughness within the catalyst layer

may explain why the power density curves for 4.5- and 2.25-mm lengths anodes are similar,

Page 110: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

101

whereas simulations predict significantly

higher power density curves for even shorter

electrodes [27,29].

6.3.2 Optimal Aspect Ratio for a Single

Electrode

The experiments presented above show

that shortening electrode length leads to higher

power densities, but the absolute amount of

power generated is reduced, as expected (Fig.

6.2a). This result suggests that using wider

electrodes will be beneficial to achieve both

higher power densities as well as higher

absolute power. Changing the electrode aspect

ratio (i.e., by widening the flow channel),

while keeping the total electrode area and the

volumetric flow rate constant, affects the

development of the boundary layer in two

ways: (1) The lower linear velocity increases

the boundary layer thickness at the end of the

electrode; and (2) shorter electrodes result in a

less developed (thinner) boundary layer.

To systematically study the influence

of electrode aspect ratio, as defined as the ratio

Figure 6.2: (a) Polarization and power curves of an air

breathing LFFC for electrode lengths 18, 9, 4.5, and

2.25 mm. (b) Corresponding polarization and power density

curves and (c) anode and cathode polarization curves

normalized to the electrode area. In all experiments were

performed with [MeOH] = 2 M, [KOH] = 1 M, and a flow

rate for each stream of 0.1 ml min-1

.

(a)

(b)

(c)

Page 111: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

102

of the electrode length-to-width, on

both current and power output, we

relied on geometric and dynamic

similarities between the

aforementioned LFFC and various

aspect ratio electrodes and to

empirically model electrodes of

various aspect ratios via experimental

data with the aforementioned LFFC.

So, rather than building several new

cell designs, we equipped this LFFC

with 4.5, 9, and 18 mm electrodes to

experimentally simulate electrode

aspect ratios of 1:2 (4.5 x 9 mm2), 2:1

(9 x 4.5 mm2), and 4:1 (18 x 2.25

mm2). The electrode surface area is

identical for these three cases (0.4

cm2). The flow rate of the electrodes

simulated (0.15 ml min-1

) was

achieved by flowing the fuel streams

at 0.2, 0.1, and 0.05 ml min-1

in the actual experiments performed in cells with 18, 9, and 4.5 mm

electrodes, respectively (Fig. 6.3a), and thus making these experimental conditions equivalent to

the conditions we are actually interested in. The flow channel width was assumed to have a

Figure 6.3: (a) Schematic of the conditions tested and their

correlation to simulated scenarios. (b) Polarization and power

density curves of an air breathing LFFC for electrode lengths 18, 9,

and 4.5 mm operated at fuel and electrolyte flow rates of 0.2, 0.1,

and 0.05 ml min-1

, respectively. In all experiments, [MeOH] = 0.

75 M and [KOH] = 1 M.

Qi=0.05 ml/min

Qi=0.1 ml/min

Qi=0.2 ml/min

1:2

4:1

2:1

Correlation to

Aspect Ratio

Tested

Condition(a)

(b)

l:9 mm / w:3 mm

l:18 mm / w:3 mm

l:4.5 mm / w:3 mml:4.5 mm

w:9 mm

l:9 mm

w:4.5 mm

l:18 mm

w:2.25 mm

Page 112: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

103

minimal influence on performance because the flow profile is primarily affected by the distance

between electrodes, which is the critical dimension for the development of the flow profiles,

which allowed us to use a fixed-width channel of 3 mm to simulate electrodes with three

different widths.

Fig. 6.3b shows the normalized cell performance for electrode lengths of 18, 9, and

4.5 mm operated at 0.2, 0.1, and 0.05 ml min-1

to simulate electrode areas of 4.5 x 9 mm2 (1:2), 9

x 4.5 mm2 (2:1), and 18 x 2.25 mm

2 (4:1). Maximum power densities of 24, 15, and 14 mW·cm

-

2 at current densities of 120, 64, and 54 mA·cm

-2 were observed for simulated electrode areas of

4.5 x 9 mm2, 9 x 4.5 mm

2, and 18 x 2.25 mm

2, respectively. The shorter and wider aspect ratios

exhibit much higher current densities and fuel conversion for a given flow rate and electrode

area. Specifically, the electrode performance increased by 70 % with the low aspect ratio (1:2)

configuration compared to a high aspect ratio (4:1) configuration. Most microfluidic LFFCs

reported in the literature [5] use electrodes with high aspect ratios (long and narrow) as they

easily develop uniform flow fields and facilitate high fuel conversion at reasonable flow

velocities. Our data suggests that a low aspect ratio (short and wide) would significantly

improve the performance of a single electrode LFFC. The results obtained here, which show that

shorter and wider electrodes will improve the performance of an air-breathing alkaline fuel cell

configuration, will be applicable to the design of other single cell microfluidic systems for

electrochemical processes, not just fuel cells, with different arrangements and reactants.

6.3.3 Optimal Electrode Length for Multichannel Configurations

Above, we reported how to optimize the performance of a single electrode configuration.

Scaling to achieve the power output needed for specific applications will require the creation of

Page 113: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

104

multichannel configurations. Through a combined simulation and experimental approach, we

study how to optimize fuel conversion when using different multi-electrode arrangements.

To model conversion for a series of electrodes, the current density across a single

electrode was measured for various methanol concentrations (0.75, 1.5, and 3.0 M), flow rates

(0.0125, 0.025, 0.05, 0.1, 0.2, 0.4, and 0.8 ml min-1

), and electrode lengths (9, 18, and 36 mm) at

the potential which corresponds to peak power (0.2 V). The main effects and interaction plots

may be found in Appendix A. By interpolation of the experimental data over the concentration

range of 0.75 to 1.5 M and 1.5 to 3.0 M, as explained in Section 2.4 of the Experimental, we can

estimate the current for any inlet concentration (Eq. 2a and 2b), and by using the estimated

current, we can calculate the outlet concentration (Eq. 3). Next, we used the outlet concentration

of the first electrode as the inlet concentration of the second to calculate (again with equations 2

and 3) the outlet concentration of the second electrode, and so on, until the outlet concentration is

only 25 % of the original inlet concentration. So for each of the three electrode lengths, we

determined the number of them needed to accomplish a conversion of 75 %. By adding the area

of the electrodes required for the specified conversion, the total electrode area needed was

calculated for each electrode length and flow rate tested (Table 6.1).

A decrease in the electrode length necessitates more electrodes to accomplish the

aforementioned 75 % conversion. However, as Table 6.1 shows, the corresponding total

electrode surface area decreases. This decrease in normalized electrode area is more pronounced

Table 6.1. Absolute electrode area required for a multiple electrode arrangement for 75 % conversion of a methanol

fuel stream

Electrode

Length (mm) Flow Rate (ml/min)

0.8 0.4 0.2 0.1 0.05 0.025 0.0125

36 228.96 113.4 59.4 31.32 17.28 11.88 8.64

18 184.68 92.34 46.98 23.76 12.42 7.02 4.32

9 187.38 91.53 44.82 23.22 11.61 6.21 3.51

Page 114: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

105

at slower flow rates. Specifically, at a flow rate of 0.0125 ml min-1

, the necessary electrode area

for 75 % conversion was reduced by 61 % (3.5 vs. 8.6 cm2) by using thirteen 9-mm electrodes as

compared to eight 36-mm electrodes. In contrast, at a flow rate of 0.8 ml min-1

, the necessary

electrode area was only reduced 18 % (187.4 vs. 229.0 cm2) by using 694 9-mm electrodes as

compared to 212 36-mm electrodes.

Next, we wished to compare these datasets with respect to surface reaction rates to better

enable comparison of the total electrode areas needed to accomplish 75 % conversion at different

flow rates. Figure 4 shows methanol reaction rates normalized to the total electrode surface area

(in mmol/min·mm2) for a 3.0 M methanol fuel stream converted across a series of electrodes

(each 9, 18, or 36-mm long) to 0.75 M (75 % conversion) for each flow rate tested.

Figure 6.4: Reaction rate of methanol within the fuel stack relative to the total electrode area, predicted based on a

model for the use of a fuel cell stack for the conversion of methanol from 3.0 M to 0.75 M with variable length

electrodes (9, 18, and 36 mm) in a stack arrangement.

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

0.8 0.4 0.2 0.1 0.05 0.025 0.0125

No

rmal

lize

d r

eac

tio

n r

ate

(m

mo

l/m

in·m

m2)

Flow Rate (ml/min)

9 mm 18 mm 36 mm

Page 115: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

106

While the shorter electrodes universally improve the normalized reaction rates, the

improvement is more pronounced for slower flow rates because the boundary layer becomes

thicker. Specifically, at a flow rate of 0.0125 ml min-1

, the resultant normalized reaction rate for

75 % conversion increases by 146 % (1.07 vs. 0.43 mmol/min·mm2) by using 9-mm electrodes

instead of 36-mm electrodes. In contrast, at a higher flow rate of 0.8 ml min-1

, the necessary

electrode area only increases by 22 % (1.28 vs. 1.05 mmol/min·mm2) by using 9-mm electrodes

instead of 36-mm electrodes.

The more pronounced influence of electrode length on conversion at slow flow rates

stems from an increase in reactant depletion due to slower fuel flow rates. This effect can be

explained by applying the Blasius solution for boundary layer thickness (Eq. 4):

(4)

where is the boundary layer thickness as defined by the point where the fluid velocity has

come within 1 % of the ‘free stream’ velocity, , is the kinematic viscosity, and is the

distance down the electrode [40]. This equation shows that the boundary layer becomes thicker

upon decreasing the flow rate. As a result, the cell becomes diffusion limited at slower flow

rates, a limitation that can be overcome by using many short electrodes.

The results here show that using several “short” electrodes in series results in much

higher conversion normalized to the surface area than fewer “longer” electrodes, especially at

slow flow rates. Consequently, applications that require low power for long periods of time, for

which fuel utilization is important, would benefit from stack designs with multiple electrodes. In

contrast, applications requiring large amounts of power for short periods of time need to operate

at higher flow rates and do not benefit as much from series of short electrodes.

Page 116: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

107

One practical aspect that we did not discuss above is the need to homogenize the fuel

concentration in the fuel stream before it enters the inlet of the next cell, after emerging from the

outlet of the previous cell with a depleted boundary layer. In the simulations above, we assumed

that a stream of uniform, but lower, fuel concentration would enter each successive cell. To

homogenize the fuel stream between successive cells, one could use one of many microfluidic

mixing methods, e.g., a herringbone mixer [27,41] or multiple inlets (or outlets) along a single

electrode to periodically push away (or remove) the depleted boundary layer.[30] However, the

addition of mixers or multiple inlets (outlets) adds complexity to the system and reduces spatial

density of the stack. Consequently, an optimal design for a given application will need to

balance the desire to minimize electrode length with the need for homogenization of the fuel

stream between successive cells, which adds additional volume and complexity to the stack.

6.4 Conclusions

Here we studied the influence of boundary layer depletion on performance in an alkaline,

air-breathing LFFC as a function of electrode length, electrode arrangement, and flow rates.

Both the current and power densities obtained here for LFFCs are similar to or better than

Nafion-based DMFCs and miniaturized conventional DMFCs [2,36-39]. We confirmed that

higher current densities can be achieved when using shorter electrodes. The power density

increased, respectively, by 65 and 88 % for 4.5-mm and 2.25-mm long electrodes compared to

an 18-mm electrode. While shorter electrodes lead to higher current densities, the absolute

current decreases, suggesting that arranged multiple short electrodes in series can improve

overall performance.

To utilize the increased power densities from shorter electrodes, we explored different

electrode arrangements and aspect ratios. Shorter and wider electrodes exhibit notably higher

Page 117: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

108

current densities and conversions for a given flow rate and electrode area. Specifically, a 67 %

increase in performance was observed when switching from a 1:2 (length-to-width) to a 4:1

aspect ratio. So by changing the standard operating aspect ratio, the performance of single

channel electrodes can be improved significantly.

Alternatively, the use of shorter electrodes in series will allow for higher fuel conversion

while maintaining higher current densities. We observed that the improvement in reaction rate

normalized by surface area as a result of shortening the electrode length is much more

pronounced when operating at slow flow rates. Consequently, arrangements comprised of

shorter electrodes will be most suitable for applications that require low power for long periods

of time, for which fuel utilization is important. However, in arrangements comprised of multiple

electrodes, the fuel stream (with depleted boundary layer) must be homogenized between

consecutive cells, which will add additional volume to a stack. The analysis reported here for

LFFCs can also be applied in the development and optimization of other microfluidic

electrochemical reactor designs [42-44].

6.5 References

1. L. Carrette, K. A. Friedrich and U. Stimming, Fuel Cells - Fundamentals and

Applications, ChemPhysChem, 2001, 1 p. 5-39.

2. A. Kundu, J. H. Jang, J. H. Gil, C. R. Jung, H. R. Lee, S. H. Kim, B. Ku and Y. S. Oh,

Micro-fuel cells-Current development and applications, Journal of Power Sources, 2007,

170 (1), p. 67-78.

3. S. Pennathur, J. C. T. Eijkel and A. Van Den Berg, Energy conversion in microsystems:

Is there a role for micro/nanofluidics?, Lab on a Chip - Miniaturisation for Chemistry

and Biology, 2007, 7 (10), p. 1234-1237.

4. C. K. Dyer, Fuel cells for portable applications, Journal of Power Sources, 2002, 106 (1-

2), p. 31-34.

5. E. Kjeang, N. Djilali and D. Sinton, Microfluidic fuel cells: A review, Journal of Power

Sources, 2009, 186 (2), p. 353-369.

6. C. Potera, Beyond batteries: Portable hydrogen fuel cells, Environmental Health

Perspectives, 2007, 115 (1), p. A38-A41.

Page 118: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

109

7. P. O. López-Montesinos, N. Yossakda, A. Schmidt, F. R. Brushett, W. E. Pelton and P. J.

A. Kenis, Design, fabrication, and characterization of a planar, silicon-based,

monolithically integrated micro laminar flow fuel cell with a bridge-shaped

microchannel cross-section, Journal of Power Sources, 2011, 196 (10), p. 4638-4645.

8. N. T. Nguyen and S. H. Chan, Micromachined polymer electrolyte membrane and direct

methanol fuel cells - A review, Journal of Micromechanics and Microengineering, 2006,

16 (4), p. R1-R12.

9. A. V. Tripkovic, K. D. Popovic, B. N. Grgur, B. Blizanac, P. N. Ross and N. M.

Markovic, Methanol electrooxidation on supported Pt and PtRu catalysts in acid and

alkaline solutions, Electrochimica Acta, 2002, 47 (22-23), p. 3707-3714.

10. K. F. Blurton and E. McMullin, A comparison of fuel cell performance in acid and

alkaline electrolyte, Energy Conversion, 1969, 9 (4), p. 141-144.

11. J. R. Varcoe and R. C. T. Slade, Prospects for alkaline anion-exchange membranes in

low temperature fuel cells, Fuel Cells, 2005, 5 (2), p. 187-200.

12. M. Schulze and E. Gülzow, Degradation of nickel anodes in alkaline fuel cells, Journal of

Power Sources, 2004, 127 (1-2), p. 252-263.

13. N. Wagner, M. Schulze and E. Gülzow, Long term investigations of silver cathodes for

alkaline fuel cells, Journal of Power Sources, 2004, 127 (1-2), p. 264-272.

14. Fikile R. Brushett, Matthew S. Thorum, Nicholas S. Lioutas, Matthew S. Naughton,

Claire Tornow, Huei-Ru “Molly” Jhong, Andrew A. Gewirth and Paul J. A. Kenis, A

Carbon-Supported Copper Complex of 3,5-Diamino-1,2,4-triazole as a Cathode Catalyst

for Alkaline Fuel Cell Applications, Journal of the American Chemical Society, 2010,

132 (35), p. 12185-12187.

15. C. Coutanceau, L. Demarconnay, C. Lamy and J. M. Léger, Development of

electrocatalysts for solid alkaline fuel cell (SAFC), Journal of Power Sources, 2006, 156

(1 SPEC. ISS.), p. 14-19.

16. E. R. Choban, L. J. Markoski, A. Wieckowski and P. J. A. Kenis, Microfluidic fuel cell

based on laminar flow, Journal of Power Sources, 2004, 128 (1), p. 54-60.

17. E. Kjeang, R. Michel, D. A. Harrington, D. Sinton and N. Djilali, An alkaline

microfluidic fuel cell based on formate and hypochlorite bleach, Electrochimica Acta,

2008, 54 (2), p. 698-705.

18. Ranga S. Jayashree, Seong Kee Yoon, Fikile R. Brushett, Pedro O. Lopez-Montesinos,

Dilip Natarajan, Larry J. Markoski and Paul J. A. Kenis, On the performance of

membraneless laminar flow-based fuel cells, Journal of Power Sources, 2010, 195 (11),

p. 3569-3578.

19. E. Kjeang, R. Michel, D. A. Harrington, N. Djilali and D. Sinton, A microfluidic fuel cell

with flow-through porous electrodes, Journal of the American Chemical Society, 2008,

130 (12), p. 4000-4006.

20. W. Sung and J. W. Choi, A membraneless microscale fuel cell using non-noble catalysts

in alkaline solution, Journal of Power Sources, 2007, 172 (1), p. 198-208.

21. E. Kjeang, J. McKechnie, D. Sinton and N. Djilali, Planar and three-dimensional

microfluidic fuel cell architectures based on graphite rod electrodes, Journal of Power

Sources, 2007, 168 (2), p. 379-390.

Page 119: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

110

22. K. S. Salloum, J. R. Hayes, C. A. Friesen and J. D. Posner, Sequential flow membraneless

microfluidic fuel cell with porous electrodes, Journal of Power Sources, 2008, 180 (1), p.

243-252.

23. Kamil S. Salloum and Jonathan D. Posner, A membraneless microfluidic fuel cell stack,

Journal of Power Sources, 2011, 196 (3), p. 1229-1234.

24. J. L. Cohen, D. A. Westly, A. Pechenik and H. D. Abruna, Fabrication and preliminary

testing of a planar membraneless microchannel fuel cell, Journal of Power Sources,

2005, 139 (1-2), p. 96-105.

25. A. Li, S. H. Chan and N. T. Nguyen, A laser-micromachined polymeric membraneless

fuel cell, Journal of Micromechanics and Microengineering, 2007, 17 (6), p. 1107-1113.

26. M. H. Chang, F. Chen and N. S. Fang, Analysis of membraneless fuel cell using laminar

flow in a Y-shaped microchannel, Journal of Power Sources, 2006, 159 (2), p. 810-816.

27. S. K. Yoon, G. W. Fichtl and P. J. A. Kenis, Active control of the depletion boundary

layers in microfluidic electrochemical reactors, Lab on a Chip, 2006, 6 (12), p. 1516-

1524.

28. A. S. Hollinger, R. J. Maloney, R. S. Jayashree, D. Natarajan, L. J. Markoski and P. J. A.

Kenis, Nanoporous separator and low fuel concentration to minimize crossover in direct

methanol laminar flow fuel cells, Journal of Power Sources, 2010, 195 (11), p. 3523-

3528.

29. A. Bazylak, D. Sinton and N. Djilali, Improved fuel utilization in microfluidic fuel cells:

A computational study, Journal of Power Sources, 2005, 143 (1-2), p. 57-66.

30. A. Ebrahimi Khabbazi, A. J. Richards and M. Hoorfar, Numerical study of the effect of

the channel and electrode geometry on the performance of microfluidic fuel cells, Journal

of Power Sources, 2010, 195 (24), p. 8141-8151.

31. R. S. Jayashree, L. Gancs, E. R. Choban, A. Primak, D. Natarajan, L. J. Markoski and P.

J. A. Kenis, Air-breathing laminar flow-based microfluidic fuel cell, Journal of the

American Chemical Society, 2005, 127 (48), p. 16758-16759.

32. D. P. Davies, P. L. Adcock, M. Turpin and S. J. Rowen, Stainless steel as a bipolar plate

material for solid polymer fuel cells, Journal of Power Sources, 2000, 86 (1–2), p. 237-

242.

33. F. R. Brushett, W. P. Zhou, R. S. Jayashree and P. J. A. Kenis, Alkaline Microfluidic

Hydrogen-Oxygen Fuel Cell as a Cathode Characterization Platform, Journal of the

Electrochemical Society, 2009, 156 (5), p. B565-B571.

34. Matt S. Naughton, Fikile R. Brushett and Paul J. A. Kenis, Carbonate resilience of

flowing electrolyte-based alkaline fuel cells, Journal of Power Sources, 2011, 196 (4), p.

1762-1768.

35. E. R. Choban, P. Waszczuk and P. J. A. Kenis, Characterization of limiting factors in

laminar flow-based membraneless microfuel cells, Electrochemical and Solid State

Letters, 2005, 8 (7), p. A348-A352.

36. S. C. Kelley, G. A. Deluga and W. H. Smyrl, Miniature methanol/air polymer electrolyte

fuel cell, Electrochemical and Solid-State Letters, 2000, 3 (9), p. 407-409.

Page 120: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

111

37. T. Shimizu, T. Momma, M. Mohamedi, T. Osaka and S. Sarangapani, Design and

fabrication of pumpless small direct methanol fuel cells for portable applications, Journal

of Power Sources, 2004, 137 (2), p. 277-283.

38. T. J. Yen, N. Fang, X. Zhang, G. Q. Lu and C. Y. Wang, A micro methanol fuel cell

operating at near room temperature, Applied Physics Letters, 2003, 83 (19), p. 4056-

4058.

39. V. Baglio, A. Stassi, E. Modica, V. Antonucci, A. S. Aricò, P. Caracino, O. Ballabio, M.

Colombo and E. Kopnin, Performance comparison of portable direct methanol fuel cell

mini-stacks based on a low-cost fluorine-free polymer electrolyte and Nafion membrane,

Electrochimica Acta, 2010, 55 (20), p. 6022-6027.

40. Alex Bauer, Elod L. Gyenge and Colin W. Oloman, Electrodeposition of Pt-Ru

nanoparticles on fibrous carbon substrates in the presence of nonionic surfactant:

Application for methanol oxidation, Electrochimica Acta, 2006, 51 (25), p. 5356-5364.

41. A. D. Stroock, S. K. W. Dertinger, A. Ajdari, I. Mezic, H. A. Stone and G. M.

Whitesides, Chaotic mixer for microchannels, Science, 2002, 295 (5555), p. 647-651.

42. D. T. Whipple, E. C. Finke and P. J. A. Kenis, Microfluidic Reactor for the

Electrochemical Reduction of Carbon Dioxide: The Effect of pH, Electrochemical and

Solid-State Letters, 2010, 13 (9), p. D109-D111.

43. D. T. Whipple and P. J. A. Kenis, Prospects of CO2 Utilization via Direct Heterogeneous

Electrochemical Reduction, Journal of Physical Chemistry Letters, 2010, 1 (24), p. 3451-

3458.

44. Brian A. Rosen, Amin Salehi-Khojin, Michael R. Thorson, Wei Zhu, Devin T. Whipple,

Paul J. A. Kenis and Richard I. Masel, Ionic Liquid–Mediated Selective Conversion of

CO2 to CO at Low Overpotentials, Science, 2011, 334 (6056), p. 643-644.

Page 121: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

112

Chapter 7

Summary of Accomplishments and Future Directions

7.1 Summary of Accomplishments

An electrochemical reactor to reduce CO2 can be used as a means of producing chemical

fuels and feedstocks to store electrical energy from renewable sources such as wind and solar in

chemical form. When combined with a “green” renewable source, this technology provides a

means of (a) reducing dependence on foreign oil via enabling penetration of renewable

technologies into the transportation sector and (b) reducing CO2 emissions by moving away from

fossil fuels.

For large scale implementation of processes for the electrochemical reduction of CO2, the

following need to be improved: (a) energetic efficiency, (b) reaction rate, and (c) product

selectivity. Significant research has focused on catalyst development, reactor operation and

electrolyte optimization, primarily for the production of large hydrocarbons (e.g., CH4, C2H4).

Because the electrochemical reaction to produce CO has much higher energetic efficiencies as

compared to formic acid, methane, ethylene, and other C1, C2, and C3 products, we believe an

electrochemical process for the reduction of CO2 to CO will be most favorable for

commercialization. Consequently, we used a microfluidic platform to study the influence of

electrolytes, reactor conditions and catalysts to improve (1) reaction energetic efficiency,

(2) reaction rate, and (3) reaction selectivity for the electrochemical reduction of CO2 to CO.

In Chapter 2, cation size within electrolyte solutions was observed to play a significant

role in the electrochemical reduction of CO2 into CO. Specifically, small cations (e.g., Li+, and

Na+) favor H2 evolution, and, consequently, have a low Faradaic yield for CO production,

Page 122: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

113

whereas large cations (e.g., Rb+, Cs

+) increase CO production and suppress H2 evolution,

resulting in a high Faradaic yield for CO, here up to 96 %. We observed the highest partial

current density of 175 mA/cm2 for CO at a cell potential of 3.0 V in the presence of RbOH.

Furthermore, we observed strong trends between anion choice and cell performance. Anions

influence both the anode and cathode polarization and consequently, the partial current density

for CO. Specifically, our data suggests that small anions reduce the reaction overpotential and

that the presence of OH- (high pH) further improves performance.

In Chapter 3, a microfluidic platform was modified to look at the onset potential for CO

production from the reduction of CO2 with Ionic Liquids. With an 18 mole% EMIM BF4

solution, CO production was achieved at a total reactor potential of only 1.5 V, which constitutes

a maximum cathode overpotential of 0.17 V. This low of an onset potential constitutes a drastic

reduction in the onset potential for CO production in an arrangement that achieves excellent

selectivity for CO. However, the individual electrode polarization plots clearly showed that

EMIM BF4 poisons the Pt anode. This anode poisoning was drastically reduced via the

development of a dual electrolyte reactor, with a Nafion cation-exchange membrane. While the

use of a Nafion membrane eliminated anode poisoning, the EMIM BF4 solution appeared to

poison the membrane as was evident by an increased reactor resistance when using the

membrane, which resulted in low current densities. The reactor developed here enabled the

analysis of the overall reactor onset potential, however, at the expense of low current densities

due to high reactor resistances.

In Chapter 4, drastic improvements were achieved regarding the partial current density

for CO with EMIM salts. Specifically, substituting of the BF4 anion for either an OH or Cl anion

enabled partial current densities for CO greater than 60 mA/cm2 at a reactor potential of 3 V.

Page 123: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

114

Furthermore, the substitution of BF4 for OH or Cl eliminated anode poisoning, which stemmed

from the BF4 anion, thus enabling the use of a single channel electrolyte stream. The elimination

of both the membrane reduced the reactor resistance from EMIM+ poisoning the Nafion as

discussed in Chapter 3, which led to the drastic improvements in the current densities.

In Chapter 5, amine-based novel organometallic complexes were developed for the

electrochemical reduction of CO2 to CO, which reduce the onset potential by roughly 0.1 V and

more significantly, increased the partial current densities for CO as compared to a more

traditional catalyst, Ag. Specifically, AgDAT, AgPc, and AgPz have comparable or better

current densities with increased selectivity for CO as compared to Ag. Furthermore, when

comparing the performance relative to the silver loading, the amine-based organometallic

catalysts outperform the silver catalyst by more than an order of magnitude. Furthermore, when

using copper-based organometallic catalysts (i.e., CuDAT), the product selectivity changed.

This resultant change in selectivity and current density demonstrates that organometallic

complexes have potential to “tune” catalysts to produce the desired products. Further

exploration of the effect of amine-based catalysis on the electrochemical conversion of CO2 to

CO, or to other products, may lead to further improvements in performance.

Chapter 6 reported on the influence of boundary layer depletion on performance in an

alkaline, air-breathing LFFC as a function of electrode length, electrode arrangement, and flow

rates. Both the current and power densities obtained here for LFFCs are similar to or better than

Nafion-based DMFCs and miniaturized conventional DMFCs [1-5]. We confirmed that higher

current densities can be achieved when using shorter electrodes. The power density increased,

respectively, by 65 and 88 % for 4.5-mm and 2.25-mm long electrodes compared to an 18-mm

electrode. While shorter electrodes lead to higher current densities, the absolute current

Page 124: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

115

decreases, suggesting that arranged multiple short electrodes in series can improve overall

performance. To utilize the increased power densities from shorter electrodes, we explored

different electrode arrangements and aspect ratios. Shorter and wider electrodes exhibit notably

higher current densities and conversions for a given flow rate and electrode area. Specifically, a

67 % increase in performance was observed when switching from a 1:2 (length-to-width) to a 4:1

aspect ratio. So, by changing the standard operating aspect ratio, the performance of single

channel electrodes can be improved significantly. Alternatively, the use of shorter electrodes in

series will allow for higher fuel conversion while maintaining higher current densities. We

observed that the improvement in reaction rate normalized by surface area as a result of

shortening the electrode length is much more pronounced when operating at slow flow rates.

Consequently, arrangements comprised of shorter electrodes will be most suitable for

applications that require low power for long periods of time, for which fuel utilization is

important. However, in arrangements comprised of multiple electrodes, the fuel stream (with

depleted boundary layer) must be homogenized between consecutive cells, which will add

additional volume to a stack. The analysis reported here for LFFCs can also be applied in the

development and optimization of other microfluidic electrochemical reactor designs [6-8].

7.2 Future Directions

Numerous areas of opportunity exist within the field of electrochemical reduction of

CO2. The following potential directions of research appear to hold much promise:

The development of a wide range of organometallic catalysts could open the door to

further stabilizing the CO2- intermediate, which has an overpotential of -1.9 V vs. SHE, and

thereby make the electrochemical reduction process much more efficient from an energetic

standpoint. Furthermore, we showed in the case of silver modified organometallic catalysts, that

Page 125: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

116

organometallic catalysts can increase both current density and product selectivity. Potentially

more interesting is that organometallic catalysts can change product distributions and onset

potentials. Organometallic complexes may be able to change the intermediate adsorbate-surface

bond strengths, and thus change the catalytic activity of other catalysts. The added stabilizing

effect of organometallic complexes could shift volcano curves from Sachtler-Fahrenfort and

Tranaka-Tamaru plot and thereby, enable less expense or more active catalysts.

Tremendous improvements in catalytic activity can potentially be achieved by improving

catalyst utilization. For common fuel cell catalysts (e.g., Pt), carbon supports have been shown

to drastically improve performance. However, in Chapter 5 a carbon support was shown to

decrease the CO production and favor hydrogen evolution. Other catalysts selective for CO, i.e.,

Au and Zn, should be explored with carbon supports. Furthermore, the use of other supports

such as carbon nanotubes should be explored.

While some long term durability experiments were carried out in this study, further

analysis of catalyst stability is necessary at reasonable overpotentials, with a wide range of

electrolytes and conditions.

A commercial process must consider the cost of CO2 capture in an economic feasibility

analysis. Coupling this work with a capture technology (i.e., Rectisol process) could enable

energy savings. Specifically, by combining this technology with a Rectisol process, a methanol

stream could be used to capture CO2 from an emission source and fed directly into an

electrochemical reactor. Using a liquid with a high solubility for CO2 has the potential to

increase the selectivity for CO by increasing the local concentration of CO2 (the reactant) in the

liquid at the electrode interface.

Page 126: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

117

Most of the catalysts used in these studies were lone transition metals (i.e., Cu and Ag)

supported on Nafion. Minimal studies have been carried out regarding the catalytic activity of

metallic alloys for CO2 reduction. In the same fashion that organometallic catalysts improved

the reaction activity, metallic alloys have tremendous potential to improve the overpotential,

reaction rate, and selectivity (the three areas needing improvement).

Co-catalysts have been shown (i.e., EMIM BF4) to reduce the onset potential for CO. A

wider range of co-catalysts and ionic liquids should be explored to influence the reaction onset

potential and current density. Specifically, ionic liquids with high conductivities, diffusivities,

and solubilities for CO2 should be selected to minimize mass transfer and resistance limitations.

Overall, the CO2 reduction process needs to be further improved through catalyst

development, optimization of operating conditions, and developing compatibility with CO2

capture technology. Of all these areas, catalysis has the greatest potential to improve the (1)

onset potential, (2) reaction rate, and (3) selectivity. Consequently, a much wider range of

catalysts and co-catalysts need to be explored for their activity for CO2 reduction.

7.3 References

1. S. C. Kelley, G. A. Deluga and W. H. Smyrl, Miniature methanol/air polymer electrolyte

fuel cell, Electrochemical and Solid-State Letters, 2000, 3 (9), p. 407-409.

2. T. Shimizu, T. Momma, M. Mohamedi, T. Osaka and S. Sarangapani, Design and

fabrication of pumpless small direct methanol fuel cells for portable applications, Journal

of Power Sources, 2004, 137 (2), p. 277-283.

3. T. J. Yen, N. Fang, X. Zhang, G. Q. Lu and C. Y. Wang, A micro methanol fuel cell

operating at near room temperature, Applied Physics Letters, 2003, 83 (19), p. 4056-

4058.

4. A. Kundu, J. H. Jang, J. H. Gil, C. R. Jung, H. R. Lee, S. H. Kim, B. Ku and Y. S. Oh,

Micro-fuel cells-Current development and applications, Journal of Power Sources, 2007,

170 (1), p. 67-78.

5. V. Baglio, A. Stassi, E. Modica, V. Antonucci, A. S. Aricò, P. Caracino, O. Ballabio, M.

Colombo and E. Kopnin, Performance comparison of portable direct methanol fuel cell

mini-stacks based on a low-cost fluorine-free polymer electrolyte and Nafion membrane,

Electrochimica Acta, 2010, 55 (20), p. 6022-6027.

Page 127: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

118

6. D. T. Whipple, E. C. Finke and P. J. A. Kenis, Microfluidic Reactor for the

Electrochemical Reduction of Carbon Dioxide: The Effect of pH, Electrochemical and

Solid-State Letters, 2010, 13 (9), p. D109-D111.

7. D. T. Whipple and P. J. A. Kenis, Prospects of CO2 Utilization via Direct Heterogeneous

Electrochemical Reduction, Journal of Physical Chemistry Letters, 2010, 1 (24), p. 3451-

3458.

8. Brian A. Rosen, Amin Salehi-Khojin, Michael R. Thorson, Wei Zhu, Devin T. Whipple,

Paul J. A. Kenis and Richard I. Masel, Ionic Liquid–Mediated Selective Conversion of

CO2 to CO at Low Overpotentials, Science, 2011, 334 (6056), p. 643-644.

Page 128: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

119

Appendix A

Supplementary Section to Chapter 6

A.1 Linear Interpolation of Current Density relative to Methanol Concentration

To model conversion for a series of electrodes, it is necessary to estimate the current for

intermediate methanol concentrations not tested. The current output for methanol concentrations

between 0.75 and 3.0 M were estimated using linear interpolation (Eq. 6.1 and Eq. 6.2) as

explained in the Experimental section within Chapter 6.

In Figure A1, we show the measured current output at each of the methanol

concentrations and electrode lengths recorded. Furthermore, Figure A1 shows the actual lines

resulting from the interpolation of experimental data (current) over the two concentration regions

(0.75 to 1.5 M and 1.5 to 3.0 M) for each of the seven flow rates tested (0.0125, 0.025, 0.05, 0.1,

0.2, 0.4, 0.8 ml/min) for 9, 18, and 36-mm electrodes. Table A1 shows the calculated values for

the slope, m, from Eq. 6.1 in the range of 0.0125 to 0.8 ml/min interpolated linearly for 9, 18,

and 36-mm electrodes at a cell potential of 0.2 V between [MeOH] of 3.0 M and 1.5 M as well

as 0.75 M and 1.5 M.

Page 129: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

120

Figure A1. Measured current for 9, 18, and 36 mm electrodes at 0.2 V with flow rates between 0.0125 and 0.8

ml/min interpolated linearly between inlet [MeOH] of 3.0 M and 0.75 M.

Table A1. Calculated values for the slope, m, in the interpolation (Eq. 6.1) of the values for the current for 9, 18,

and 36 mm electrode at 0.2 V with flow rates between 0.0125 and 0.8 ml/min interpolated linearly between inlet

[MeOH] of 3.0 and 0.75 M.

Electrode

length (mm) Conc. (M)

Flow rate (ml/min)

0.8 0.4 0.2 0.1 0.05 0.025 0.0125

36 1.5 to 3.0 10.23 6.26 6.71 1.50 0.67 2.67 2.67

0.75 to 1.5 -10.39 -0.36 0.00 2.74 5.33 18.67 18.67

18 1.5 to 3.0 3.57 2.25 3.11 2.53 2.27 1.67 0.94

0.75 to 1.5 14.48 17.79 16.09 8.28 8.92 8.30 11.48

9 1.5 to 3.0 3.00 1.84 1.61 1.20 0.14 0.75 0.60

0.75 to 1.5 -3.78 -0.77 -5.47 0.44 2.13 0.84 2.66

A.2 Main Effects and Interactions from Design of Experiment

Figure A2 shows the main effects plots from the design of experiment discussed in the

modeling section of Chapter 6 based on a 90 % confidence influence. These graphs confirm

many of the trends observed and discussed in the main text of this work. An increase in

methanol concentration, as seen in Figure A2a, increases the current output from the cell.

0

10

20

30

40

50

60

70

80

90

100

0.5 0.75 1 1.25 1.5 1.75 2 2.25 2.5 2.75 3 3.25

Cu

rre

nt

(mA

)

Methanol Concentration (M)

0.8 ml/min

0.4 ml/min

0.2 ml/min

0.1 ml/min

0.05 ml/min

0.025 ml/min

0.0125 ml/min

0.8 ml/min

0.4 ml/min

0.2 ml/min

0.1 ml/min

0.05 ml/min

0.025 ml/min

0.0125 ml/min

0.8 ml/min

0.4 ml/min

0.2 ml/min

0.1 ml/min

0.05 ml/min

0.025 ml/min

0.0125 ml/min

0

10

20

30

40

50

60

70

80

90

100

0.5 0.75 1 1.25 1.5 1.75 2 2.25 2.5 2.75 3 3.25

Cu

rre

nt

(mA

)

Methanol Concentration (M)

0.8 ml/min

0.4 ml/min

0.2 ml/min

0.1 ml/min

0.05 ml/min

0.025 ml/min

0.0125 ml/min

0.8 ml/min

0.4 ml/min

0.2 ml/min

0.1 ml/min

0.05 ml/min

0.025 ml/min

0.0125 ml/min

0.8 ml/min

0.4 ml/min

0.2 ml/min

0.1 ml/min

0.05 ml/min

0.025 ml/min

0.0125 ml/min

0

10

20

30

40

50

60

70

80

90

100

0.5 0.75 1 1.25 1.5 1.75 2 2.25 2.5 2.75 3 3.25

Cu

rre

nt

(mA

)

Methanol Concentration (M)

0.8 ml/min

0.4 ml/min

0.2 ml/min

0.1 ml/min

0.05 ml/min

0.025 ml/min

0.0125 ml/min

0.8 ml/min

0.4 ml/min

0.2 ml/min

0.1 ml/min

0.05 ml/min

0.025 ml/min

0.0125 ml/min

0.8 ml/min

0.4 ml/min

0.2 ml/min

0.1 ml/min

0.05 ml/min

0.025 ml/min

0.0125 ml/min

0

10

20

30

40

50

60

70

80

90

100

0.5 0.75 1 1.25 1.5 1.75 2 2.25 2.5 2.75 3 3.25

Cu

rre

nt

(mA

)

Methanol Concentration (M)

0.8 ml/min

0.4 ml/min

0.2 ml/min

0.1 ml/min

0.05 ml/min

0.025 ml/min

0.0125 ml/min

0.8 ml/min

0.4 ml/min

0.2 ml/min

0.1 ml/min

0.05 ml/min

0.025 ml/min

0.0125 ml/min

0.8 ml/min

0.4 ml/min

0.2 ml/min

0.1 ml/min

0.05 ml/min

0.025 ml/min

0.0125 ml/min

9 mm 18 mm 36 mm

Page 130: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

121

Similarly, an increase in electrode length increases the total current output, at a diminishing rate

for larger electrode lengths as seen in Figure A2b. As seen in Figure A2c, the flow rate has a

drastic effect only at very low flow rates. At higher flow rates, the change in flow rate has only

a small influence on the current output.

Figure A3 shows the interaction plots for (a) the effects of methanol concentration for all

eight flow rates tested on output current, (b) the effects of methanol concentration for all three

electrode lengths on output current, and (c) the effects of methanol concentration for all eight

flow rates tested with respect to the current output. The data in Figure A3a, which shows the

43

44

45

46

47

48

49

50

51

52

53

0.5 1 1.5 2 2.5 3

Cu

rre

nt (m

A)

Methanol concentration (M)

0

10

20

30

40

50

60

70

80

5 10 15 20 25 30 35 40

Cu

rre

nt (m

A)

Electrode Length (mm)

30

35

40

45

50

55

0 0.2 0.4 0.6 0.8

Cu

rre

nt (m

A)

Flow Rate (mL min-1)

(a) (b)

Figure A2: Schematic of an alkaline, air-

breathing, direct methanol, laminar flow fuel cell

(LFFC). Two 1-mm thick PMMA windows

provide the flow fields for the electrolyte and fuel

streams between the current collectors and the

nanoporous separator.

(c)

Page 131: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

122

effect of methanol concentration for the various flow rates, reveals that methanol concentration

has less of an effect at high flow rates compared to low flow rates. This trend can be explained

by the fact that less depletion is observed at the high flow rates. Figure A3b shows that methanol

concentrations are more important for long electrodes than for short electrodes, because

depletion is more significant with long electrodes.

Comparing all three interaction plots shows that electrode length is the most sensitive factor

(Figure A3c). Especially at low flow rates, the effect of depletion on the electrode becomes

more significant, as is evident from a drastic decrease in the improvement when using longer

electrodes.

Page 132: USING MICROSCALE FLOW CELLS TO STUDY THE … · Araz Zarkhah, Chris, Timberg, Karl Siil, and Saman Moniri, who assisted me in data acquisition and whom I was given the privilege to

123

25

30

35

40

45

50

55

60

0.5 1 1.5 2 2.5 3

Cu

rren

t (m

A)

Methanol Concentration (M)

0.8 ml min-1

0.4 ml min-1

0.2 ml min-1

0.1 ml min-1

0.05 ml min-1

0.025 ml min-1

0.0125 ml min-1

20

30

40

50

60

70

80

0.5 1 1.5 2 2.5 3

Cu

rren

t (m

A)

Methanol Concentration (M)

36 mm

18 mm

9 mm

20

30

40

50

60

70

80

90

5 10 15 20 25 30 35 40

Cu

rren

t (m

A)

Electrode Length (mm)

0.8 ml min-1

0.4 ml min-1

0.2 ml min-1

0.1 ml min-1

0.05 ml min-1

0.025 ml min-1

0.0125 ml min-1

Figure A3. Interaction plots for (a) effects of

methanol concentration for all eight flow rates

tested on output current, (b) effects of methanol

concentration for all three electrode lengths on

output current, and (c) effects of methanol

concentration for all eight flow rates tested on

output current.

(a)

(b) (c)