18.11.2012 Views

Methanol Electrochemical Conversion to Formaldehyde over Bulk ...

Methanol Electrochemical Conversion to Formaldehyde over Bulk ...

Methanol Electrochemical Conversion to Formaldehyde over Bulk ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Methanol</strong> <strong>Electrochemical</strong> <strong>Conversion</strong> <strong>to</strong> <strong>Formaldehyde</strong> <strong>over</strong> <strong>Bulk</strong> Metal and<br />

Supported Catalysts<br />

by<br />

Mohsina Islam, B.Sc., M.Sc., M.S.<br />

A DISSERTATION<br />

IN<br />

CHEMISTRY<br />

Submitted <strong>to</strong> the Graduate Faculty<br />

of Texas Tech University in<br />

Partial Fulfillment of<br />

the Requirements for<br />

the Degree of<br />

DOCTOR OF PHILOSOPHY<br />

Approved<br />

Carol Korzeniewski<br />

Chairperson of the Committee<br />

Shaorong Liu<br />

Dominick J. Casadonte Jr.<br />

Accepted<br />

Dean of the Graduate School<br />

May, 2006


Copyright 2006, Mohsina Islam<br />

ii


ACKNOWLEDGMENTS<br />

I would like <strong>to</strong> express my deep appreciation <strong>to</strong> my research advisor Dr.<br />

Carol Korzeniewski, Professor of Chemistry. It is due <strong>to</strong> her scholastic guidance,<br />

strong support and encouragement that ultimately made this work possible. I<br />

always admire and respect her as a research men<strong>to</strong>r. Under her supervision, I<br />

have learned many scientific and precious aspects of research and this will be a<br />

good guideline for my future career. I also would like <strong>to</strong> thank Dr. Dominick J.<br />

Casadonte Jr. and Dr. Shaorong Liu for their assistance and valuable comments<br />

throughout my graduate studies.<br />

I would like <strong>to</strong> express my appreciation <strong>to</strong> all my present and past group<br />

members for their assistance. Special thanks <strong>to</strong> Mr. Brandon J. Sheehan and Mr.<br />

Duan C. Hindes for their ingenuous assistance in machining some very<br />

challenging work in this research.<br />

I like <strong>to</strong> thank my parents for their continuous support and encouragement.<br />

Thank U, Mom for your support and help especially for taking care of my little<br />

daughter during my busy time. Finally, I am especially thankful for the patience<br />

and support of my husband, SM Rahmat Ullah and lots of inspiration from my<br />

daughter Nosheen S. Ullah.<br />

iii


TABLE OF CONTENTS<br />

ACKNOWLEDGMENTS iii<br />

ABSTRACT vii<br />

LIST OF TABLES & SCHEME ix<br />

LIST OF FIGURES x<br />

CHAPTER<br />

I. INTRODUCTION<br />

Basic Fuel Cell Concepts 1<br />

Polymer Electrolyte Membrane Fuel Cell (PEMFC) 4<br />

Direct <strong>Methanol</strong> Fuel Cell (DMFC) 6<br />

<strong>Methanol</strong> Oxidation Pathways 12<br />

Bimetallic Catalysts for <strong>Methanol</strong> Oxidation 15<br />

Project Focus 17<br />

Fluorescence Assay for <strong>Formaldehyde</strong> 19<br />

Summary 20<br />

References<br />

II. INSTRUMENTATION AND GENERAL EXPERIMENTAL<br />

CONDITIONS<br />

Reagents and Catalysts Materials 26<br />

Sonochemistry for Catalysts Synthesis 26<br />

<strong>Bulk</strong> Electrode Preparation 28<br />

iv<br />

1<br />

22<br />

26


Electrochemistry 30<br />

<strong>Formaldehyde</strong> Detection 36<br />

References<br />

III. METHANOL ELECTROCHEMICAL CONVERSION TO<br />

FORMALDEHYDE OVER BULK METAL ELECTRODES<br />

Introduction 41<br />

Experimental 42<br />

<strong>Formaldehyde</strong> Determination 43<br />

Results and Discussion 43<br />

Conclusions 61<br />

References 70<br />

IV. METHANOL ELECTROCHEMICAL CONVERSION TO<br />

FORMALDEHYDE OVER SUPPORTED CATALYST<br />

MATERIALS<br />

Introduction 72<br />

Experimental 73<br />

<strong>Formaldehyde</strong> Determination 73<br />

Results and Discussion 74<br />

Conclusions 91<br />

References<br />

V. A MICRO-VOLUME ELECTROCHEMICAL CELL FOR THE<br />

STUDY OF METHANOL ELECTROOXIDATION PATHWAYS<br />

v<br />

40<br />

41<br />

72<br />

100<br />

102<br />

Introduction 102


Experimental and Instrumentation 103<br />

Reagents 103<br />

Small Volume <strong>Electrochemical</strong> Cell 105<br />

Choice of Cell Material and Cell Fabrication 105<br />

Results and Discussion 108<br />

Cell Characterization in 0.1M H2SO4 108<br />

Cell Characterization using a Reversible Redox Active<br />

Probe<br />

vi<br />

110<br />

Conclusions 117<br />

References<br />

VI. SUMMARY 119<br />

118<br />

References 128


ABSTRACT<br />

The electrochemical oxidation of 1.0 M CH3OH in 0.1 M H2SO4 <strong>over</strong><br />

different types of platinum-ruthenium (PtRu) materials was investigated. Focus<br />

was on the determination of formaldehyde (H2CO) produced as a function of Ru<br />

content in arc-melted bulk alloys and nanometer-scale catalyst materials. A<br />

sensitive fluorescence assay for formaldehyde, which had a detection limit down<br />

<strong>to</strong> 100 nM H2CO, was employed. The reaction potentials, reported in volts<br />

versus the reversible hydrogen electrode reference (VRHE), were in the range of<br />

0.5 VRHE <strong>to</strong> 0.8 VRHE. The lower potentials approach voltages that have<br />

technological relevance <strong>to</strong> fuel cells. Electrolysis was performed on 50 µL<br />

samples for a period of 180 s. Based on the coulometry, the reactant depletion<br />

in the cell is below 1%.<br />

In experiments with bulk PtRu alloys, three samples with respective Ru mole<br />

fraction (XRu) of 0.1, 0.3 and 0.9 were employed. Reactions were also carried out<br />

on bulk polycrystalline Pt for reference. Compared <strong>to</strong> Pt, the H2CO yields were<br />

lower for the oxidation of methanol <strong>over</strong> PtRu. Among the PtRu alloys, the H2CO<br />

yields decreased with increasing XRu, except at the lowest potential studied (0.5<br />

VRHE), where the formaldehyde yield was lowest for the sample with XRu = 0.3.<br />

The finding is consistent with XRu = 0.3 being the most active PtRu composition<br />

for methanol electrochemical oxidation at ambient temperature. The results are<br />

attributed <strong>to</strong> lower reactivity of methanol on the electrode with XRu = 0.9 at 0.5<br />

vii


VRHE due <strong>to</strong> due <strong>to</strong> blocking of initial dissociative chemisorption steps by inhibiting<br />

oxides present at Ru sites.<br />

Compared <strong>to</strong> bulk electrodes, methanol oxidation <strong>over</strong> nanometer-scale<br />

catalyst resulted in H2CO yields below 10 % under the conditions studied. The<br />

following catalyst materials were used on gold electrode: Pt-Black, C/Pt, 10 wt %<br />

on Vulcan XC-72R carbon, PtRu black with XRu = 0.5, PtRu catalyst prepared via<br />

a sonochemical (SC) method with XRu = 0.1, 0.25 and 0.5. High current densities<br />

were achieved during sample electrolysis. The results indicate the nanometer-<br />

scale catalyst converts methanol <strong>to</strong> more complete oxidation products. The<br />

findings are consistent with earlier studies and are attributed <strong>to</strong> readsorption and<br />

complete oxidation of H2CO within multiple catalyst layers, leading <strong>to</strong> lower H2CO<br />

yields. Similar <strong>to</strong> results for smooth, bulk electrodes, the H2CO yield was<br />

significantly higher for methanol oxidation <strong>over</strong> the lowest Ru content nanometer-<br />

scale catalyst (XRu= 0.1) and approached the response for Pt black.<br />

This project also advanced the design of the small volume electrolysis cell<br />

employed for the thesis research by incorporating a machinable MACOR glass<br />

ceramic disk window, which resists oxygen permeation. The cell response was<br />

characterized by observing the reversible, diffusion controlled waves for<br />

hexamine ruthenium trichloride (Ru (NH3)6Cl3) in cyclic voltammetry<br />

measurements.<br />

viii


LIST OF TABLES<br />

1.1 Fuel cells and their characteristics and applications.<br />

2.1 Sample composition in a<strong>to</strong>mic % based on X-ray<br />

fluorescence and X-ray Diffraction.<br />

3.1 Summary of charge and formaldehyde yields from the<br />

oxidation of 1.0 M CH3OH on bulk Pt and PtRu electrode<br />

materials.<br />

4.1 Summary of charge and formaldehyde yields from the<br />

oxidation of 1.0 M CH3OH <strong>over</strong> Pt based nanoscale catalyst<br />

materials.<br />

4.2 Summary of charge and formaldehyde yields from the<br />

oxidation of 1.0 M CH3OH <strong>over</strong> PtRu based nanoscale<br />

catalyst materials.<br />

LIST OF SCHEME<br />

I Mechanism suggested for methanol oxidation <strong>to</strong> H2CO.<br />

ix<br />

5<br />

31<br />

53<br />

81<br />

85<br />

14


LIST OF FIGURES<br />

1.1: Conventional H2/O2 PEM fuel cell.<br />

1.2 (a): Structure of Nafion.<br />

1.2 (b): Three-phase model of Nafion; a fluorocarbon region (A), an<br />

interfacial zone (B) and an ionic cluster region (C).<br />

1.3: Schematic diagram of a DMFC.<br />

1.4: Some pathways of methanol electrochemical oxidation.<br />

1.5: Scheme of methanol oxidation on Pt showing the consecutive<br />

stripping of hydrogen a<strong>to</strong>ms.<br />

1.6: <strong>Methanol</strong> oxidation reactions on Pt.<br />

1.7: <strong>Methanol</strong> oxidation reactions on PtRu.<br />

2.1: Scanned pho<strong>to</strong>graphs of bulk PtRu disks. (a) XRu= 0.1, (b) XRu=<br />

0.3.<br />

2.2: Front portion of micro-volume electrochemical cell used for<br />

sample electrolysis.<br />

2.3: Block diagram of the au<strong>to</strong>mated formaldehyde analyzer.<br />

2.4: Calibration curve for standard H2CO solution. Concentration in<br />

the range of 2x10 -6 -2.5x10 -5 M.<br />

2.5: Fluorescence signal of standard H2CO solution.<br />

3.1: Cyclic voltammograms of bulk electrode materials employed in<br />

the study. Scans were recorded in Ar purged 0.1 M H2SO4.<br />

Polycrystalline Pt electrode (Pt-solid, a); bulk PtRu alloys (PtRusolid)<br />

with Ru mole fraction (XRu) of 0.1 (b), 0.3 (c) and 0.9 (d).<br />

Cyclic voltammetric measurements were performed by scanning<br />

the potential between 0.05 <strong>to</strong> 0.75 VRHE. The scan rate was 50<br />

mV/s.<br />

3.2: Cyclic voltammogram of a polycrystalline Pt electrode in Ar<br />

purged 0.1 M H2SO4 recorded at a scan rate of 50 mV/s.<br />

3.3: H2CO yields from methanol electrochemical oxidation on a<br />

polycrystalline Pt electrode. All experiments involved the<br />

x<br />

7<br />

8<br />

8<br />

11<br />

13<br />

18<br />

18<br />

18<br />

32<br />

34<br />

38<br />

39<br />

39<br />

47<br />

48<br />

54


electrolysis of 1.0 M CH3OH in 0.1 M H2SO4 in a volume of 50 µL<br />

for a period of 180s.<br />

3.4: H2CO yields from methanol electrochemical oxidation on a PtRu-<br />

solid (XRu = 0.1) electrode. All experiments involved the<br />

electrolysis of 1.0 M CH3OH in 0.1 M H2SO4 in a volume of 50 µL<br />

for a period of 180s.<br />

3.5: H2CO yields from methanol electrochemical oxidation on a PtRu-<br />

solid (XRu = 0.3) electrode. All experiments involved the<br />

electrolysis of 1.0 M CH3OH in 0.1 M H2SO4 in a volume of 50 µL<br />

for a period of 180s.<br />

3.6: H2CO yields from methanol electrochemical oxidation on a PtRu-<br />

solid (XRu = 0.9) electrode. All experiments involved the<br />

electrolysis of 1.0 M CH3OH in 0.1 M H2SO4 in a volume of 50 µL<br />

for a period of 180s.<br />

3.7: Comparison of H2CO yields from methanol (1.0 M)<br />

electrochemical oxidation on bulk electrode materials at 0.6 and<br />

0.5 VRHE.<br />

3.8: Plot of the nanomoles H2CO formed vs. the reaction charge<br />

passed following the oxidation of 1.0 M methanol on a<br />

polycrystalline Pt electrode in 0.1 M H2SO4 (Reaction time = 180<br />

s).<br />

3.9: Plot of H2CO formation rate averaged <strong>over</strong> 180 s electrolysis<br />

periods vs. potential for the oxidation of 1.0 M methanol in 0.1 M<br />

H2SO4 on a polycrystalline Pt electrode.<br />

3.10: Plot of the nanomoles H2CO formed vs. the reaction charge<br />

passed following the oxidation of 1.0 M methanol on a bulk alloy<br />

PtRu- solid (XRu = 0.1) electrode in 0.1 M H2SO4 (Reaction time =<br />

180 s).<br />

3.11: Plot of H2CO formation rate averaged <strong>over</strong> 180 s electrolysis<br />

periods vs. potential for the oxidation of 1.0 M methanol in 0.1 M<br />

H2SO4 on a bulk alloy PtRu- solid (XRu = 0.1) electrode.<br />

3.12: Plot of the nanomoles H2CO formed vs. the reaction charge<br />

passed following the oxidation of 1.0 M methanol on a bulk alloy<br />

PtRu- solid (XRu = 0.3) electrode in 0.1 M H2SO4 (Reaction time =<br />

180 s).<br />

3.13: Plot of H2CO formation rate averaged <strong>over</strong> 180 s electrolysis<br />

xi<br />

67<br />

55<br />

56<br />

57<br />

58<br />

62<br />

63<br />

64<br />

65<br />

66


periods vs. potential for the oxidation of 1.0 M methanol in 0.1 M<br />

H2SO4 on a bulk alloy PtRu- solid (XRu = 0.3) electrode.<br />

3.14: Plot of the nanomoles H2CO formed vs. the reaction charge<br />

passed following the oxidation of 1.0 M methanol on a bulk alloy<br />

PtRu- solid (XRu = 0.9) electrode in 0.1 M H2SO4 (Reaction time =<br />

180 s).<br />

3.15: Plot of H2CO formation rate averaged <strong>over</strong> 180 s electrolysis<br />

periods vs. potential for the oxidation of 1.0 M methanol in 0.1 M<br />

H2SO4 on a bulk alloy PtRu- solid (XRu = 0.9) electrode.<br />

4.1: Cyclic voltammogram of a polycrystalline gold electrode in 0.1 M<br />

H2SO4; Scan rate: 50 mV/s.<br />

4.2: Cyclic voltammogram of a polycrystalline gold electrode in 0.1 M<br />

H2SO4 after adsorption of C/Pt, 10 wt% (a) or Pt-black (b) catalyst<br />

films; Scan Rate: 50 mV/s. Arrow represents the stripping wave<br />

for gold oxide.<br />

4.3: Cyclic voltammograms of catalyst materials employed in the<br />

study. Scans were recorded in Ar purged 0.1 M H2SO4 at 50<br />

mV/s. The following nanoscale catalyst materials were studied as<br />

thin films adsorbed on a Au electrode- Johnson Matthey (JM)<br />

PtRu with XRu = 0.5 (a); PtRu catalyst prepared via a<br />

sonochemical (SC) method with XRu = 0.1 (b) and 0.5 (c).<br />

4.4: H2CO yields from methanol electrochemical oxidation in 0.1 M<br />

H2SO4 on Pt -Black catalyst on a polycrystalline Au electrode. All<br />

experiments involved the electrolysis of 1.0 M CH3OH in 0.1 M<br />

H2SO4 in a volume of 50 µL for a period of 180s.<br />

4.5: H2CO yields from methanol electrochemical oxidation in 0.1 M<br />

H2SO4 on C/Pt, 10 wt% catalyst on a polycrystalline Au electrode.<br />

All experiments involved the electrolysis of 1.0 M CH3OH in 0.1 M<br />

H2SO4 in a volume of 50 µL for a period of 180s.<br />

4.6: H2CO yields from methanol electrochemical oxidation in 0.1 M<br />

H2SO4 on SC PtRu, XRu= 0.1 catalyst on a polycrystalline Au<br />

electrode. All experiments involved the electrolysis of 1.0 M<br />

CH3OH in 0.1 M H2SO4 in a volume of 50 µL for a period of 180s.<br />

4.7: H2CO yields from methanol electrochemical oxidation in 0.1 M<br />

H2SO4 on SC PtRu, XRu= 0.1 catalyst, capped by Nafion on a<br />

polycrystalline Au electrode. All experiments involved the<br />

electrolysis of 1.0 M CH3OH in 0.1 M H2SO4 in a volume of 50 µL<br />

xii<br />

68<br />

69<br />

76<br />

77<br />

79<br />

82<br />

82<br />

86<br />

86


for a period of 180s.<br />

4.8: H2CO yields from methanol electrochemical oxidation in 0.1 M<br />

H2SO4 on SC PtRu, XRu= 0.25 catalyst on a polycrystalline Au<br />

electrode. All experiments involved the electrolysis of 1.0 M<br />

CH3OH in 0.1 M H2SO4 in a volume of 50 µL for a period of 180s.<br />

4.9: H2CO yields from methanol electrochemical oxidation in 0.1 M<br />

H2SO4 on SC PtRu, XRu= 0.5 catalyst on a polycrystalline Au<br />

electrode. All experiments involved the electrolysis of 1.0 M<br />

CH3OH in 0.1 M H2SO4 in a volume of 50 µL for a period of 180s.<br />

4.10: H2CO yields from methanol electrochemical oxidation in 0.1 M<br />

H2SO4 on JM PtRu, XRu= 0.5 catalyst on a polycrystalline Au<br />

electrode. All experiments involved the electrolysis of 1.0 M<br />

CH3OH in 0.1 M H2SO4 in a volume of 50 µL for a period of 180s.<br />

4.11: H2CO yield from methanol (1.0 M) electrochemical oxidation on<br />

catalyst materials at 0.5 VRHE.<br />

4.12: Plot of H2CO formation rate averaged <strong>over</strong> 180 s electrolysis<br />

periods vs. potential for the oxidation of 1.0 M methanol in 0.1 M<br />

H2SO4 <strong>over</strong> a Pt -Black catalyst layer adsorbed on a<br />

polycrystalline Au electrode.<br />

4.13: Plot of H2CO formation rate averaged <strong>over</strong> 180 s electrolysis<br />

periods vs. potential for the oxidation of 1.0 M methanol in 0.1 M<br />

H2SO4 <strong>over</strong> a C/Pt, 10 wt% catalyst layer adsorbed on a<br />

polycrystalline Au electrode.<br />

4.14: Plot of H2CO formation rate averaged <strong>over</strong> 180 s electrolysis<br />

periods vs. potential for the oxidation of 1.0 M methanol in 0.1 M<br />

H2SO4 <strong>over</strong> a SC PtRu, XRu= 0.1 catalyst layer adsorbed on a<br />

polycrystalline Au electrode.<br />

4.15: Plot of H2CO formation rate averaged <strong>over</strong> 180 s electrolysis<br />

periods vs. potential for the oxidation of 1.0 M methanol in 0.1 M<br />

H2SO4 <strong>over</strong> a SC PtRu, XRu= 0.1 catalyst layer (capped by<br />

Nafion) adsorbed on a polycrystalline Au electrode.<br />

4.16: Plot of H2CO formation rate averaged <strong>over</strong> 180 s electrolysis<br />

periods vs. potential for the oxidation of 1.0 M methanol in 0.1 M<br />

H2SO4 <strong>over</strong> a SC PtRu, XRu= 0.25 catalyst layer adsorbed on a<br />

polycrystalline Au electrode.<br />

4.17: Plot of H2CO formation rate averaged <strong>over</strong> 180 s electrolysis<br />

xiii<br />

95<br />

87<br />

87<br />

88<br />

89<br />

93<br />

93<br />

94<br />

94<br />

95


periods vs. potential for the oxidation of 1.0 M methanol in 0.1 M<br />

H2SO4 <strong>over</strong> a SC PtRu, XRu= 0.50 catalyst layer adsorbed on a<br />

polycrystalline Au electrode.<br />

4.18: Plot of H2CO formation rate averaged <strong>over</strong> 180 s electrolysis<br />

periods vs. potential for the oxidation of 1.0 M methanol in 0.1 M<br />

H2SO4 <strong>over</strong> a JM PtRu, XRu= 0.50 catalyst layer adsorbed on a<br />

polycrystalline Au electrode.<br />

4.19: Plot of the nanomoles H2CO formed vs. the reaction charge<br />

passed following the oxidation of 1.0 M methanol in 0.1 M H2SO4<br />

<strong>over</strong> a Pt-Black catalyst layer adsorbed on a polycrystalline Au<br />

electrode (Reaction time = 180 s).<br />

4.20: Plot of the nanomoles H2CO formed vs. the reaction charge<br />

passed following the oxidation of 1.0 M methanol in 0.1 M H2SO4<br />

<strong>over</strong> a C/Pt, 10 wt% catalyst layer adsorbed on a polycrystalline<br />

Au electrode (Reaction time = 180 s).<br />

4.21: Plot of the nanomoles H2CO formed vs. the reaction charge<br />

passed following the oxidation of 1.0 M methanol in 0.1 M H2SO4<br />

<strong>over</strong> a SC PtRu, XRu= 0.10 catalyst layer adsorbed on a<br />

polycrystalline Au electrode (Reaction time = 180 s).<br />

4.22: Plot of the nanomoles H2CO formed vs. the reaction charge<br />

passed following the oxidation of 1.0 M methanol in 0.1 M H2SO4<br />

<strong>over</strong> a SC PtRu, XRu= 0.10 catalyst layer (capped by Nafion)<br />

adsorbed on a polycrystalline Au electrode (Reaction time = 180<br />

s).<br />

4.23: Plot of the nanomoles H2CO formed vs. the reaction charge<br />

passed following the oxidation of 1.0 M methanol in 0.1 M H2SO4<br />

<strong>over</strong> a SC PtRu, XRu= 0.25 catalyst layer adsorbed on a<br />

polycrystalline Au electrode (Reaction time = 180 s).<br />

4.24: Plot of the nanomoles H2CO formed vs. the reaction charge<br />

passed following the oxidation of 1.0 M methanol in 0.1 M H2SO4<br />

<strong>over</strong> a SC PtRu, XRu= 0.50 catalyst layer adsorbed on a<br />

polycrystalline Au electrode (Reaction time = 180 s).<br />

4.25: Plot of the nanomoles H2CO formed vs. the reaction charge<br />

passed following the oxidation of 1.0 M methanol in 0.1 M H2SO4<br />

<strong>over</strong> a JM PtRu, XRu= 0.10 catalyst layer adsorbed on a<br />

polycrystalline Au electrode (Reaction time = 180 s).<br />

xiv<br />

96<br />

96<br />

97<br />

97<br />

98<br />

98<br />

99<br />

99


5.1:<br />

Exploded view of electron transfer at an electrode surface in a<br />

thin-layer flow cell.<br />

5.2: Side view and <strong>to</strong>p view of the electrochemical cell window.<br />

5.3: Cyclic voltammograms of a polycrystalline Pt electrode in 0.1 M<br />

H2SO4 recorded in a conventional electrochemical cell (a) and the<br />

electrolysis cell with the electrode mounted far from MACOR<br />

window (b). The scan rate was 50 mV/s.<br />

5.4: Cyclic voltammograms of a polycrystalline Pt electrode in 0.1 M<br />

H2SO4 recorded in an electrolysis cell with the electrode mounted<br />

near the MACOR window. Scan rate was 10 mV/s (a) and 50<br />

mV/s (b).<br />

5.5: Cyclic voltammograms of a polycrystalline Pt electrode in 0.1 M<br />

H2SO4 recorded in an electrolysis cell with the electrode mounted<br />

near the MACOR window (a) and mounted inside the MACOR<br />

window (Arrow represents skewing of the hydrogen adsorption<br />

region) (b). The scan rate was 10 mV/s.<br />

5.6: Cyclic voltammograms of [Ru(NH3)6Cl3] in 0.1 M NaNO3 recorded<br />

in a conventional electrochemical cell (a, b) and the MACOR<br />

cavity of the thin layer electrolysis cell shown in Figure 3.2 (c, d).<br />

The [Ru(NH3)6] 3+ concentration and scan rates were : (a) 5 × 10 -3<br />

M and 50 mV/s , (b) 5 × 10 -3 M and 0.2 mV/s (c) 2.5 × 10 -3 M and<br />

0.2 mV/s (d) 2.5 ×1 0 -2 M and 1 mV/s. The scans started from 0.4<br />

VAg/AgCl and swept <strong>to</strong>ward more negative potentials.<br />

6.1: Bar charts comparing the formaldehyde yields from reaction of<br />

1.0 M CH3OH in 0.1 M H2SO4 during 180 s electrolysis periods for<br />

solid bulk electrodes.<br />

6.2: Bar charts comparing the formaldehyde yields from reaction of<br />

1.0 M CH3OH in 0.1 M H2SO4 <strong>over</strong> nanoscale catalyst materials<br />

during 180 s reaction periods.<br />

xv<br />

104<br />

107<br />

111<br />

112<br />

113<br />

116<br />

123<br />

124


CHAPTER I<br />

INTRODUCTION<br />

Basic Fuel Cell Concepts<br />

Fuel cells have emerged as a promising technology for electric power<br />

generation. Fuel cell development spans more than 150 years. This long period<br />

has been characterized by many ups and downs [1]. In 1839, William Robert<br />

Grove first demonstrated the fuel cell using a simple experiment where a reverse<br />

of water electrolysis was performed <strong>to</strong> generate electricity from hydrogen (H2)<br />

and oxygen (O2) gases. Grove stated a definition for this type of fuel cell that still<br />

applies <strong>to</strong>day –“A Fuel cell is an electrochemical device that continuously<br />

converts chemical energy in<strong>to</strong> electrical energy (and some heat and water)<br />

without any combustion for as long as an oxidant and reactant (fuel) are<br />

supplied” [2]. During the 1960s, the U.S. National Aeronautics and Space<br />

Administration (NASA) used precursors <strong>to</strong> <strong>to</strong>day’s fuel cell technology as power<br />

sources in Gemini and Apollo spacecraft, and fuel cells still provide electricity and<br />

water for the space shuttle [2]. Fuel cells are similar in many ways <strong>to</strong> batteries.<br />

They both share the electrochemical nature of the power generation process.<br />

Furthermore, like internal combustion engines (ICEs), fuel cells will operate<br />

continuously by consuming some type of fuel [2].<br />

The fuel cell consists of an anode and a cathode separated by an<br />

electrolyte membrane. In a hydrogen-oxygen fuel cell (Figure 1.1), hydrogen<br />

flows past the anode, which contains a catalyst coating that facilitates oxidation<br />

1


of the gas <strong>to</strong> form pro<strong>to</strong>ns (H + ) and electrons (e - ). The electrolyte membrane<br />

passes the pro<strong>to</strong>ns <strong>to</strong> the cathode side of the fuel cell, and the electrons released<br />

flow through the external circuit. At the cathode, a catalyst coating facilitates the<br />

reduction of oxygen: pro<strong>to</strong>ns and electrons combine with oxygen a<strong>to</strong>ms on the<br />

catalyst surface <strong>to</strong> produce water (Equations 1-3) [2].<br />

Anode reaction: H2 (g) → 2H + (aq) +2e - (1)<br />

Cathode reaction: ½ O2 (g) + 2H + (aq) + 2e - → H2O (l) (2)<br />

Cell reaction: H2 (g) + ½ O2 (g) → H2O (l) (3)<br />

The oxygen required for fuel cell operation typically comes from air, but<br />

hydrogen is not as readily available. Hydrogen can be produced from different<br />

processes, including natural gas steam reforming, electrolysis, pho<strong>to</strong>electrolysis,<br />

biomass gasification and pyrolysis and pho<strong>to</strong>biological methods. In natural gas<br />

steam reforming, the first step is <strong>to</strong> expose natural gas <strong>to</strong> high temperature<br />

steam <strong>to</strong> produce hydrogen, carbon monoxide, and carbon dioxide. The second<br />

step is <strong>to</strong> convert the carbon monoxide with steam <strong>to</strong> additional hydrogen and<br />

carbon dioxide. In electrolysis, electric energy is used <strong>to</strong> split water <strong>to</strong> form<br />

hydrogen and oxygen gas. Pho<strong>to</strong>electrolysis is a process where sunlight<br />

absorbed by a semiconduc<strong>to</strong>r electrode provides the energy <strong>to</strong> split water in<strong>to</strong><br />

hydrogen and oxygen. The production of hydrogen can result from high<br />

temperature gasification and low temperature pyrolysis of biomass (i.e., wood<br />

chips and agricultural wastes). Pho<strong>to</strong>biological processes produce hydrogen<br />

using light energy through the metabolic activities of certain pho<strong>to</strong>synthetic<br />

microbes.<br />

2


There are some limitations in the use of hydrogen. As hydrogen is a gas,<br />

it is difficult <strong>to</strong> s<strong>to</strong>re and distribute. <strong>Methanol</strong>, natural gas, coal gas,<br />

hydrocarbons (methane, propane) are some additional potential fuels. A<br />

reformer can be used for turning hydrocarbon or alcohol fuels in<strong>to</strong> hydrogen,<br />

which is then fed in<strong>to</strong> the fuel cell [3]. Alternatively, simple alcohols can be<br />

directly used although the reaction kinetics are slower than for hydrogen.<br />

Catalysts are required <strong>to</strong> enable fuel cell reactions. The cost, availability and<br />

performance of catalyst materials are important criteria in the development of a<br />

fuel cell.<br />

Since in a fuel cell, the energy s<strong>to</strong>red in the bonds of chemical reactants is<br />

converted directly <strong>to</strong> electricity, in theory a fuel cell can operate at much higher<br />

efficiencies than ICEs, which require the intermediate production of heat from<br />

chemical reactants [2-4]. Fuel cells are conventionally said <strong>to</strong> be non Carnot-<br />

limited [1] and have potential <strong>to</strong> extract more electricity from the same amount of<br />

fuel compared <strong>to</strong> ICEs. Fuel cells can have zero or ultra low emissions, which<br />

makes them attractive for transport applications. The fuel cell itself has no<br />

moving parts, making it a quiet source of power and not readily susceptible <strong>to</strong><br />

mechanical breakdown. The potential of fuel cells <strong>to</strong> operate efficiently lowers<br />

fuel consumption, which benefits conservation efforts and results in reduced<br />

emission of gaseous pollutants and the green house gas CO2 (the latter is<br />

emitted when hydrocarbons are reformed <strong>to</strong> produce H2).<br />

There are several types of fuel cells, which are mainly classified according<br />

<strong>to</strong> their electrolytes and operating temperatures. These include the polymer<br />

3


electrolyte membrane fuel cell (PEMFC), phosphoric acid fuel cell (PAFC), direct<br />

methanol fuel cell (DMFC), alkaline fuel cell (AFC), molten carbonate fuel cell<br />

(MCFC), and solid oxide fuel cell (SOFC). The characteristics of all fuel cells are<br />

summarized in Table 1.1 [1-3]. Low operating temperatures, rapid start up, light<br />

weight, high power density and simplicity make PEMFCs and DMFCs attractive<br />

for transportation and consumer electronics.<br />

Polymer Electrolyte Membrane Fuel Cells (PEMFCs)<br />

The electrolyte in PEMFCs or SPEFCs (Solid Polymer Electrolyte Fuel<br />

Cells) is a polymeric ion exchange material that enables high pro<strong>to</strong>n conductivity.<br />

Nafion, which is a perfluorosulfonated polymer consisting of<br />

polytetrafluorethylene backbone and perfluoroether side chain with a sulfonate<br />

group is commonly used. The PEM is a thin sheet (50-200 µm) that when<br />

hydrated allows hydrogen ions <strong>to</strong> cross it. A general schematic of a PEMFC is<br />

shown in Figure 1.1. These fuel cells are capable of operating at ambient<br />

pressure and temperatures just below 100 ºC. PEMFCs were the first fuel cells<br />

<strong>to</strong> be used in space exploration missions. The Gemini program employed a 1kW<br />

fuel cell stack as an auxiliary power source [5]. The his<strong>to</strong>rical development of<br />

PEM fuel cells has been described recently [6]. PEMFCs also provided the<br />

astronauts with clean drinking water. Initially a polystyrene sulfonate (PSS)<br />

polymer was employed as a membrane, which proved unstable. This led NASA<br />

<strong>to</strong> rely on AFC systems.<br />

A major breakthrough in the field of PEM fuel cells came with the<br />

development of Nafion membranes by DuPont. These membranes possess a<br />

4


Table 1.1. Fuel cells and their characteristics and applications [1-3].<br />

Fuel cell<br />

type<br />

Electrolyte Charge<br />

carrier<br />

Op.<br />

temp.<br />

(˚C)<br />

5<br />

Efficien<br />

cy<br />

(%)<br />

Life time<br />

(hr)<br />

Power<br />

range/Applications<br />

AFC KOH OH - 60-120 35-55 >10,000 40,000 5-250 kW,<br />

au<strong>to</strong>motive, CHP<br />

(combined heat<br />

and power<br />

generation)<br />

PAFC H3PO4 H + ~220 40 >40,000 200 kW, portable<br />

CHP<br />

MCFC Lithium CO3<br />

and<br />

potassium<br />

carbonate<br />

2- ~650 > 50 >40,000 200 kW-MW range,<br />

CHP & standalone.<br />

SOFC Solid oxide<br />

(ytria,<br />

zirconia)<br />

DMFC H2SO4/<br />

HClO4<br />

O 2- ~1000 > 50 >40,000 2 kW-MW range,<br />

CHP & standalone.<br />

H + 50-100 40-50 >10,000 1-100 kW.<br />

Transportation,<br />

remote power


higher acidity, pro<strong>to</strong>n conductivity and far greater stability than polystyrene<br />

sulfonate. The structure of Nafion is shown Figure 1.2.<br />

As PEMFCs use solid membrane electrolyte and the only liquid product is<br />

water, corrosion problems are minimal. PEMFCs can produce high power<br />

densities and offer the advantages of low weight and small volume. The main<br />

disadvantage of the PEMFC is that it is highly sensitive <strong>to</strong> fuel impurities, such as<br />

carbon monoxide (CO), which is produced as a by-product of hydrocarbon<br />

reforming <strong>to</strong> hydrogen [7]. The CO level of 10-50 ppm poisons the catalyst,<br />

causing severe degradation of cell performance. The effect has become<br />

pronounced when CO c<strong>over</strong>s platinum catalyst sites, preventing the hydrogen<br />

fuel from reacting [3, 7].<br />

Direct <strong>Methanol</strong> Fuel Cells (DMFCs)<br />

These cells are similar <strong>to</strong> the PEMFC in that they both can use a polymer<br />

membrane as electrolyte. DMFCs operate at the same temperatures as<br />

PEMFCs, although slightly higher temperatures are preferable in order <strong>to</strong><br />

improve the power density. In DMFCs, the anode catalyst directly oxidizes<br />

methanol, eliminating the need for a fuel reformer, which simplifies operation.<br />

DMFCs are being targeted for tiny <strong>to</strong> mid size applications like cellular phones<br />

and lap<strong>to</strong>p computers. Compared <strong>to</strong> hydrogen, methanol is an attractive fuel<br />

option. It can be produced from natural gas or renewable biomass resources.<br />

Among the alcohol fuels, methanol is the most easily oxidized fully <strong>to</strong> CO2, and it<br />

is easy <strong>to</strong> s<strong>to</strong>re, transport and distribute. The DMFC can be operated with liquid<br />

or gaseous methanol/water mixtures. It is assumed that the existing<br />

6


Figure 1.1: Conventional H2/O2 PEM fuel cell<br />

[3].<br />

7


Figure 1.2 (a): Structure of Nafion.<br />

Figure 1.2 (b): Three-phase model of Nafion; a fluorocarbon region (A), an<br />

interfacial zone (B) and an ionic cluster region (C) [5].<br />

8


infrastructure for fuels may be adapted <strong>to</strong> methanol, which has the advantage of<br />

a high specific energy density [5]. Thus, methanol has become a potential<br />

candidate <strong>to</strong> power fuel cells [8, 9].<br />

DMFCs were first studied <strong>over</strong> 40 years ago. At that time, an alkaline<br />

electrolyte was used. However, carbonation of the alkaline electrolyte (caused<br />

by the evolved CO2) decreased efficiency by reducing the electrolyte conductivity<br />

and de-polarizing the cathode [10]. In the late 1970s <strong>to</strong> early 1980s, several<br />

researchers explored the use of H2SO4 as an electrolyte [11]. Currently,<br />

researchers focus on the use of a PEM capable of transporting pro<strong>to</strong>n. A PEM<br />

has the same advantages in a DMFC as it does in a PEMFC operating on H2 and<br />

O2 (e.g. lightweight, good pro<strong>to</strong>n conductivity).<br />

Figure 1.3 shows a schematic diagram of a DMFC. Electrons and pro<strong>to</strong>ns<br />

are liberated at the anode with the aid of a catalyst. The electrons and pro<strong>to</strong>ns<br />

travel through the external circuit and PEM, respectively, <strong>to</strong> the cathode<br />

electrocatalyst where they are consumed <strong>to</strong>gether with oxygen in a reduction<br />

reaction (Equations 4-6) [2].<br />

Anode reaction: CH3OH (aq) + H2O (l) → CO2 (g) +6H + (aq) +6e - (4)<br />

Cathode reaction: 6H + (aq) +6e - + 3/2 O2 (g) → 3 H2O (l) (5)<br />

Cell reaction: CH3OH (aq) + 3/2 O2 (g) → CO2 (g) + 2 H2O (l) (6)<br />

DMFCs have limitations, which include slow oxidation kinetics of methanol<br />

and reactant cross<strong>over</strong> through the membrane [2, 12]. Although significant<br />

performance gains have been achieved using PEMs, many problems still exist.<br />

The best available and widely used PEM <strong>to</strong> date is Nafion. However, Nafion is<br />

9


permeable <strong>to</strong> methanol (and water). This permeability enables methanol <strong>to</strong><br />

“cross<strong>over</strong>” from the anode <strong>to</strong> the cathode and lowers cathode performance.<br />

Since methanol will be oxidized at the cathode, it creates a “mixed” electrode<br />

potential that will be lower than the standard potential for oxygen reduction. In<br />

order <strong>to</strong> circumvent this problem, one can modify the catalyst composition. In<br />

addition, several alternatives <strong>to</strong> Nafion are being investigated. One common<br />

approach is <strong>to</strong> create membranes that exhibit less methanol cross<strong>over</strong>.<br />

Membranes based on sulfonated poly (ether ke<strong>to</strong>ne) [13], sulfonated<br />

polysulfones and poly (ether sulfones) [14] have been developed and tested.<br />

While these membranes exhibit less methanol permeability than Nafion, their<br />

pro<strong>to</strong>n conductivities are substantially lower and hence they do not perform as<br />

well. Pickup et al. [15, 16] have developed composite poly (1-<br />

methylpyrrole)/Nafion membranes. These membranes exhibit ca. 50% less<br />

methanol cross<strong>over</strong> without a significant increase in the resistance (or decrease<br />

in conductivity) of the composite membrane.<br />

While methanol is the most common hydrocarbon that is being explored<br />

for direct oxidation in fuel cells, many other hydrocarbons have been either<br />

proposed or studied. Savadogo’s group has investigated running fuel cells<br />

directly on both propane [15] and acetals [17]. Qi, Z. has recently reported good<br />

low-current density performance for cells running on 2-propanol [18]. Ethanol is<br />

also being explored as a possible fuel [19].<br />

10


Figure 1.3: Schematic diagram of a DMFC [3].<br />

11


<strong>Methanol</strong> Oxidation Pathways<br />

The electrochemical-oxidation of methanol has attracted extensive interest<br />

<strong>over</strong> the past few years because of its potential in DMFC applications [20-23]. In<br />

a typical DMFC anode, methanol is oxidized <strong>over</strong> nanometer scale catalyst<br />

particles <strong>to</strong> CO2 through a multi-step six electron process (CH3OH+ H2O → CO2<br />

+ 6H + +6e - ). Some pathways for methanol electrochemical oxidation are shown<br />

in Figure 1.4. The conversion <strong>to</strong> CO2 is complicated, as several pathways are<br />

possible. Furthermore, the involvement of the electrode surface gives rise <strong>to</strong><br />

geometric and electronic effects that are poorly unders<strong>to</strong>od and difficult <strong>to</strong> probe.<br />

The continuous electrochemical oxidation of methanol generally suffers from<br />

poisoning of the catalyst by adsorbed CO. CO poison formation from methanol<br />

and its inhibiting effect have been studied exhaustively [24, 25, 26]. It is also well<br />

established from early studies that methanol oxidation occurs via parallel<br />

pathways giving rise <strong>to</strong> a complex network of reactions involving formaldehyde<br />

(H2CO) and formic acid (HCOOH) as side products and sources of efficiency loss<br />

for DMFC operation. [9, 27, 28, 29, 30-34, 41]. It has been proposed [31] that<br />

the elementary steps leading <strong>to</strong> H2CO accumulation could start with the<br />

adsorption of a methanol molecule via the oxygen a<strong>to</strong>m <strong>to</strong> form a methoxide<br />

species (H3CO)ad. The reaction involves the elimination of one hydrogen a<strong>to</strong>m<br />

(as a H + ) delivering one electron <strong>to</strong> the metal [CH3OH → (H3CO)ad + H + +e - ]. In<br />

a second step (H3CO)ad can eliminate a second H a<strong>to</strong>m from the CH3 group,<br />

forming H2CO which then desorbs [(H3CO)ad → H2CO + H + + e - ]. The<br />

mechanism discussed [31] above is shown in Scheme I. The study of these<br />

12


- +<br />

- 2e , - 2H<br />

- +<br />

CH 3OHCH 2O<br />

- 2e , - 2H<br />

HCOOH<br />

+ H O<br />

- +<br />

- 2e , - 2H<br />

- +<br />

- 4e , - 4H<br />

+ H O<br />

2<br />

CO<br />

- +<br />

- 2e , - 2H<br />

CO<br />

Figure 1.4: Some pathways of methanol electrochemical oxidation.<br />

13<br />

ads<br />

2<br />

2<br />

- H O<br />

2<br />

- +<br />

- 2e , - 2H


eactions especially pathways leading <strong>to</strong> H2CO have been unexplored until<br />

recently due <strong>to</strong> interferences which inhibit detection of the intermediate species<br />

Scheme I: Mechanism suggested for methanol oxidation <strong>to</strong> H2CO [31].<br />

by modern in situ analysis techniques like mass spectrometry [35,36] and<br />

infrared spectroscopy [37-39]. Direct detection of H2CO by mass spectrometry is<br />

generally hindered by the methanol mass fragments at m/z = 28-30 [27, 35, 36].<br />

IR spectroscopy is also incapable of producing H2CO spectra as H2CO in<br />

aqueous solution forms a gem-diol (H2C(OH)2) which <strong>over</strong>laps with water<br />

vibrational bands [37-39]. Recent studies [27, 28, 30, 40] have brought serious<br />

attention <strong>to</strong> the stable by-products of methanol oxidation, as these affect CO2<br />

production and decrease energy conversion efficiency [9, 41-43]. These efforts<br />

mainly involved the use of differential electrochemical mass spectrometry<br />

(DEMS), which can only detect CO2 and methyl formate directly. In their DEMS<br />

studies, assuming that CO2, formic acid and H2CO are the only direct reaction<br />

products, the faradic current for H2CO was found indirectly as the difference<br />

between the <strong>to</strong>tal Faradic charges measured and the partial currents/charges of<br />

CO2 and formic acid formation, calculated using corresponding calibration<br />

constants [27, 28, 40]. In this research project, we are mainly interested in<br />

14


H2CO, which is formed in the first two electron oxidation step (CH3OH → H2CO +<br />

2H + +2e - ). The determination of H2CO produced from methanol oxidation <strong>over</strong><br />

smooth or porous catalyst/electrode materials during short electrolysis periods<br />

requires very sensitive analytical methods, since small amounts of H2CO are<br />

produced. A recent study by the Iwasita group demonstrated H2CO<br />

quantification using a longer electrolysis time (1000s) and smooth polycrystalline<br />

and Pt(III) electrodes with high performance liquid chroma<strong>to</strong>graphy (HPLC) as a<br />

detection method [31]. For the last few years, our group has emphasized a direct<br />

measurement technique for H2CO based on fluorescence spectroscopy [29, 41].<br />

<strong>Methanol</strong> electrochemical oxidation <strong>to</strong> CO2 is complex and significant<br />

improvements in catalyst development are required for low-temperature DMFCs<br />

[9]. Conventionally, Pt based catalyst materials have been employed. The<br />

oxidation of methanol on Pt based catalysts is assumed <strong>to</strong> be proceeded by the<br />

adsorption of the molecule followed by several steps of depro<strong>to</strong>nation which<br />

results the main catalyst poison Pt-COads [5]. A scheme for this process is shown<br />

in Figure 1.5. These CO intermediates are only removed when there are<br />

oxygenated species present on the Pt surface. The mechanisms of these<br />

reactions on Pt are represented in Figure 1.6.<br />

Bimetallic Catalysts for <strong>Methanol</strong> Oxidation<br />

Since Pt surface is extremely susceptible <strong>to</strong> poisoning and rapidly blocked<br />

by adsorbed CO intermediates, methanol electrochemical oxidation reactions on<br />

Pt are comparatively slow [7, 47-49]. To enhance the catalytic activity and<br />

increase the progress of CO removal by oxidation <strong>to</strong> CO2, a number of co-metals<br />

15


have been added <strong>to</strong> Pt catalysts. Studies have shown that the most promising<br />

electro-catalysts are bimetallic materials consisting of Pt and a second transition<br />

metal including Ru, Re, Os, Mo, Pb, Bi and Sn that activates H2O and promote<br />

the oxidation of CO <strong>to</strong> CO2 [47-49]. Of all bimetallic materials mentioned, Pt-Ru<br />

has long been recognized as exhibiting superior catalytic activity, if compared <strong>to</strong><br />

Pt and are by far the most widely used anode catalysts for DMFCs [9, 48, 49]. A<br />

bifunctional mechanism [50-52] is considered <strong>to</strong> be responsible for the<br />

enhancement of catalytic activity. Although Pt is an effective catalyst/electrode<br />

material for methanol dehydrogenation, water activation can be enhanced by<br />

incorporating ruthenium (Ru + H2O → Ru-OHads + H + + e - ), which efficiently<br />

breaks the H-OH bonds in the water molecule and forms the surface oxide<br />

necessary for the reaction <strong>to</strong> promote the CO2 production. It is anticipated that<br />

the addition of Ru lowers the potential for methanol oxidation by facilitating the<br />

removal of CO from the electrode surface, thereby increasing CO2 production [9,<br />

21, 40, 48, 49, 53-57]. The bifunctional mechanism, first proposed by Watanable<br />

and Mo<strong>to</strong>o [50] identifies the capability of Ru <strong>to</strong> nucleate oxygen-containing<br />

species at ca. 300 mV more negative than Pt. Additionally, Ru is a fairly noble<br />

metal and therefore much more stable than other promoters under the conditions<br />

of DMFC operation [5]. The simplified mechanism for methanol oxidation using<br />

Pt-Ru is shown in Figure 1.7.<br />

The catalytic activity of PtRu catalyst/electrode materials is sensitive <strong>to</strong><br />

Pt/Ru composition and structure. The performance of PtRu alloy catalysts with a<br />

certain Pt/Ru ratio in methanol oxidation reaction also depends on the electrode<br />

16


potential as well as on the methanol concentration and reaction temperature [40,<br />

51, 58]. It was observed that adding a small or medium amount of Ru (Ru mole<br />

fraction, XRu = ~ 0.1-0.5) <strong>to</strong> Pt electrode materials greatly increases catalytic<br />

activity <strong>to</strong>ward methanol oxidation [32, 40, 51, 53]. Iwasita noted that a catalyst<br />

composition of Pt/Ru between XRu = ~ 0.1-0.5 provides adequate Pt and Ru sites<br />

for the efficient methanol electrochemical reaction [56].<br />

Project Focus<br />

The goal of this project was <strong>to</strong> evaluate the importance of H2CO as a<br />

stable reaction by-product of methanol electrochemical oxidation, as<br />

accumulation of H2CO can lower efficiency in DMFCs. The study focused on the<br />

measurement of H2CO produced during the oxidation of methanol <strong>over</strong> Pt and<br />

PtRu catalyst materials, including arc-melted, bulk PtRu alloys and nano meter<br />

scale PtRu catalyst with varying Ru surface composition. The bulk alloys were<br />

prepared and characterized at the Lawerence Berkeley National Labora<strong>to</strong>ry and<br />

have become standard reference materials for the study of CO and methanol<br />

electrochemical oxidation reactions in support of PEM fuel cell anode<br />

development. This dissertation reports the first results on H2CO production from<br />

methanol electrochemical oxidation <strong>over</strong> the standard materials. A sensitive<br />

fluorescence assay was employed that enabled H2CO <strong>to</strong> be quantified following<br />

short (180 s) electrolysis periods in solutions that contained 1.0 M methanol in<br />

0.1 M H2SO4. Reactions on bulk Pt and PtRu electrodes were studied first, and<br />

afterward the results were applied <strong>to</strong> understand subsequent measurements of<br />

17


Figure 1.5: Scheme of methanol oxidation on Pt showing the consecutive<br />

stripping of hydrogen a<strong>to</strong>ms [5].<br />

Figure 1.6: <strong>Methanol</strong> oxidation reactions on Pt.<br />

Figure 1.7: <strong>Methanol</strong> oxidation reactions on PtRu.<br />

18


methanol oxidation on alloy-like nanometer-scale catalyst particles of Pt and<br />

PtRu. Quantitative detection of H2CO was critical as the electrochemical<br />

conversion of methanol <strong>to</strong> H2CO is typically low especially with porous catalyst<br />

particles. Our aim was <strong>to</strong> study the H2CO yield with different catalysts and<br />

electrodes <strong>to</strong> identify the fac<strong>to</strong>rs that limit catalyst and electrode performance in<br />

fuel cells.<br />

Fluorescence Assay for <strong>Formaldehyde</strong><br />

The direct detection of H2CO during methanol electrochemical conversion<br />

is very important because H2CO can cause severe deficiency in fuel cell<br />

performance. Disregarding the in-situ IR spectrometry and mass spectrometry<br />

techniques for H2CO analysis, liquid chroma<strong>to</strong>graphy (LC) is a possible<br />

technique [31]. However, the chroma<strong>to</strong>graphic method is slow and requires<br />

longer electrolysis periods <strong>to</strong> attain enough sensitivity for H2CO detection [31].<br />

This method is also restricted by the requirement of bulky, expensive and<br />

elaborate instrumentation. For our purpose, we were interested in the simple<br />

and fast fluorometric flow injection analysis (FIA) method developed by Dasgupta<br />

and Dong [45]. The analysis utilizes the Hantzsch reaction involving the<br />

cyclization of an amine, an aldehyde, and a β-dike<strong>to</strong>ne <strong>to</strong> form a dihydropyridine<br />

derivative. The reaction was introduced by Nash for analytical purposes in<br />

pho<strong>to</strong>metric versions using 2, 4-pentadione [60]. Based on the Hantzsch<br />

reaction, the specific use of 2, 4-pentadione and ammonium acetate (which<br />

serves as a source of NH3 and as a buffering agent) for the determination of<br />

H2CO leads <strong>to</strong> 3, 5-diacetyl-1, 4-dihydrolutidine (DDL):<br />

19


O O<br />

2 + H2CO + NH3<br />

DDL fluoresces optimally at 510 nm and has an excitation maximum at 412 nm<br />

[61]. A number of reagents were studied for fluorometric determinations, and 1,<br />

3-cyclohexadione (CHD) and 5, 5-dimethyl-CHD (dimedone) were found reactive<br />

with all aldehydes [45]. It was also determined that 1, 3-cyclohexadione can<br />

provide substantially more sensitivity than 2, 4- pentadione with acceptable<br />

selectivity <strong>over</strong> other aldehydes if proper reaction time is maintained [44].<br />

In the present work, 1, 3-cyclohexadione was used for au<strong>to</strong>mated real-<br />

time determination of H2CO. A flow injection analysis arrangement was<br />

employed (Chapter II).<br />

60 °C, 5 min<br />

Summary<br />

This dissertation describes investigations of H2CO production during the<br />

electrochemical oxidation of methanol <strong>over</strong> bulk metal electrodes and nanometer<br />

scale catalyst particles through the use of a micro volume electrochemical cell<br />

and a sensitive fluorescence assay for H2CO. In Chapter II, a description of the<br />

general experimental procedures and instrumentation are presented. Chapter II<br />

also includes a brief discussion of the electrochemical cell employed for<br />

methanol electrolysis. Chapters IIII and IV focus on the electrochemical<br />

oxidation of methanol <strong>to</strong> H2CO during reactions <strong>over</strong> Pt and PtRu solid metal<br />

electrodes and nanometer scale catalyst materials, respectively. The<br />

characterization of the electrolysis cell is described in Chapter V. Chapter VI<br />

20<br />

O O<br />

N<br />

+ 3 H2O


provides a summary which compares experimental results for bulk and catalyst<br />

electrode materials.<br />

21


References<br />

1. Blomen, J. M. J. L.; N.M.Mugerwa Fuel Cell Systems; Platinum Press:<br />

New York, 1993.<br />

2. Hoogers, G. Fuel cell technology handbook; CRC press: New York, 2003.<br />

3. Larminie, J.; Dicks, A. Fuel Cells Systems Explained; John Wiley & Sons,<br />

Inc.: New York, 2000.<br />

4. Kartha, S.; Grimes, P. Physics Today 1994, November, 54.<br />

5. Carrette, L.; Friedrich, A. K.; Stimming, U. Fuel Cells 2001 2001, 1, 5.<br />

6. Wright, P. V. Electrochimica Acta 1998, 43, 1137.<br />

7. Gottesfeld, S.; Zawodzinski, T. A., In Advances in <strong>Electrochemical</strong><br />

Science and Engineering; Alkire, R. C., Gerischer, H., Kolb, D. M. and<br />

Tobias, C. W., Ed.; Wiley-VCH: New York, 1997; Vol. 5, pp 195.<br />

8. Hogarth, M. P.; Hards, G. A. Platinum Metals Rev. 1996, 40, 150.<br />

9. Jarvi, T. D.; Stuve, E. M., In Electrocatalysis; Lipkowski, J. and Ross, P.<br />

N., Ed.; Wiley-VCH Publishers: New York, 1998; Vol. Ch. 3, pp 75.<br />

10. Murray, J. N.; Grimes, P. G. In Fuel Cells; American Institute of Chemical<br />

Engineers: New York, 1963.<br />

11. Glazebrook, R. W. J. Power Sources 1982, 7, 15.<br />

12. McNicol, B. D. J. Electroanal. Chem. 1981, 118, 71.<br />

13. Kerres, J.; Urllich, A.; Haring, T.; Preidel, W.; Baldauf, M.; Gebhardt, U. J.<br />

New -Mater. Electrochem. Syst. 2000, 3, 229.<br />

14. Nolte, R.; Ledjeff, K.; Bauer, M.; Mulhaupt, R. J. Membr. Sci. 1993, 82,<br />

221.<br />

15. Jia, N.; Lefebvre, M. C.; Halfyard, J.; Qi, Z.; Pickup, P. G. Electrochem.<br />

Solid-State Lett. 2000, 3, 529.<br />

16. Pickup, P. G.; Qi, Z. Canadian Patent Application No. 2,310,310, PCT/CA<br />

01/00767 2000.<br />

17. Savadogo, O.; Yang, X. J. Appl. Electrochem. 2001, 31, 787.<br />

22


18. Qi, Z.; Hollett, M.; Attia, A.; Kaufman, A. Electrochem. Solid-State Lett.<br />

2002, 5, A129.<br />

19. Ne<strong>to</strong>, A. O.; Giz, M. J.; Perez, J.; Ticianelli, A.; R., G. E. J. Electrochem.<br />

Soc. 2002, 149, A272.<br />

20. Iwasita, T.; Nart, F. C. J. Electroanal. Chem. 1991, 317, 291.<br />

21. Gasteiger, H. A.; Markovic, N.; Ross, P. N., Jr.; Cairns, E. J. J. Phys.<br />

Chem. 1994, 98, 617.<br />

22. Gasteiger, H. A.; Markovic, N. M.; Ross, P. N., Jr. J. Phys. Chem. 1995,<br />

99, 16757.<br />

23. Ianneillo, R.; Schmidt, V. M.; Stimming, U.; Stumper, J.; Wallau, A.<br />

Electrochim. Acta 1994, 39, 1863.<br />

24. Iwasita, T.; Nart, F. C., In Advances in <strong>Electrochemical</strong> Science and<br />

Engineering; Gerischer, H. and Tobias, C., Ed.; VCH Publishers: New<br />

York, 1995; Vol. 4, pp 123.<br />

25. Chang, S. C.; Leung, L. W. H.; Weaver, M. J. J. Phys. Chem. 1990, 94,<br />

6013.<br />

26. Shin, J.; Korzeniewski, C. J. Phys. Chem. 1995, 99, 3419.<br />

27. Jusys, Z.; Kaiser, J.; Behm, R. J. Langmuir 2003, 19, 6759.<br />

28. Jusys, Z.; Behm, R. J. J. Phys. Chem. B 2001, 105, 10874.<br />

29. Korzeniewski, C.; Childers, C. L. J. Phys. Chem. B 1998, 102, 489.<br />

30. Lu, G.-Q.; Chrzanowski, W.; Wieckowski, A. J. Phys. Chem. B 2000, 104,<br />

5566.<br />

31. Batista, E. A.; Malpass, G. R. P.; Motheo, A. J.; Iwasita, T. J. Electroanal.<br />

Chem. 2004, 571, 273.<br />

32. Batista, E. A.; Hoster, H.; Iwasita, T. J. Electroanal. Chem. 2003, 554-555,<br />

265.<br />

33. Iwasita, T. Electrochim. Acta 2002, 47, 3663.<br />

34. Cao, D.; Lu, G.-Q.; Wieckowski, A.; Wasileski, S. A.; Neurock, M. J. Phys.<br />

Chem. B. 2005, 109, 11622.<br />

23


35. Wasmus, S.; Wang, J.-T.; Savinell, R. F. J. Electrochem. Soc. 1995, 142,<br />

3825.<br />

36. Lin, W.-F.; Wang, J.-T.; Savinell, R. F. J. Electrochem. Soc. 1997, 144,<br />

1917.<br />

37. Fan, Q.; Pu, C.; Smotkin, E. S. J. Electrochem. Soc. 1996, 143, 3053.<br />

38. Juan<strong>to</strong>, S.; Beden, B.; Hahn, F.; Leger, J.-M.; Lamy, C. J. Electroanal.<br />

Chem. 1987, 237, 119.<br />

39. Olivi, P.; L.O.S. Bulhoes, J.-M. L.; Hahn, F.; Beden, B.; Lamy, C.<br />

Electrochim. Acta 1996, 41, 927.<br />

40. Jusys, Z.; Kaiser, J.; Behm, R. J. Electrochim. Acta 2002, 47, 3693.<br />

41. Childers, C. L.; Huang, H.; Korzeniewski, C. Langmuir 1999, 15, 786.<br />

42. Markovic, N. M.; Gasteiger, H. A.; Ross, P. N., Jr.; Jiang, X.; Villegas, I.;<br />

Weaver, M. J. Electrochim. Acta 1994, 40, 91.<br />

43. Korzeniewski, C.; Kardash, D. J. Phys. Chem. B 2001, 105, 8663.<br />

44. Fan, Q.; Dasgupta, P. K. Anal. Chem. 1994, 66, 551.<br />

45. Dong, S.; Dasgupta, P. K. Environ. Sci. Technol. 1987, 21, 581.<br />

46. Dasgupta, P. K.; Genfa, Z.; Li, J.; Boring, C. B.; Jambunathan, S.; Al-Horr,<br />

R. Anal. Chem. 1999, 71, 1400.<br />

47. Ralph, T. R.; Hogarth, M. P. Platinum Metals Rev. 2002, 46, 3.<br />

48. Ralph, T. R.; Hogarth, M. P. Platinum Metals Rev. 2002, 46, 117.<br />

49. Hogarth, M. P.; Ralph, T. R. Platinum Metals Rev. 2002, 46, 146.<br />

50. Watanabe, M.; Mo<strong>to</strong>o, S. J. Electroanal. Chem. 1975, 60, 267.<br />

51. Gasteiger, H. A.; Markovic, N.; Ross, P. N., Jr.; Cairns, E. J. J. Phys.<br />

Chem. 1993, 97, 12020.<br />

52. Gasteiger, H. A.; Markovic, N.; Ross, P. N., Jr.; Cairns, E. J. J.<br />

Electrochem. Soc. 1994, 141, 1795.<br />

53. Gao, L.; Huang, H.; Korzeniewski, C. Electrochim. Acta 2004, 49, 1281.<br />

24


54. Long, J. W.; Stroud, R. M.; Swider-Lyons, K. E.; Rolison, D. R. J. Phys.<br />

Chem. B 2000, 104, 9772.<br />

55. Tong, Y. Y.; Kim, H. S.; Babu, P. K.; Waszczuk, P.; Wieckowski, A.;<br />

Oldfield, E. J. Am. Chem. Soc. 2002,124, 468.<br />

56. Iwasita, T.; Hoster, H.; John-Anacker, A.; Lin, W. F.; Vielstich, W.<br />

Langmuir 2000, 16, 522.<br />

57. Korzeniewski, C.; Basnayake, R.; Vijayaraghavan, G.; Li, Z.; Xu, S.;<br />

Casadonte, D. J., Jr. Surf. Sci. 2004,573, 100.<br />

58. Kardash, D.; Korzeniewski, C.; Markovic, N. J. Electroanal. Chem. 2001,<br />

500, 518.<br />

59. Huang, H.; Korzeniewski, C.; Vijayaraghavan, G. Electrochim. Acta<br />

2002,42, 3675.<br />

60. Nash, T. Biochem.J.1953, 55, 416.<br />

61. Belman, S. Anal. Chim. Acta 1963, 29, 120.<br />

25


CHAPTER II<br />

INSTRUMENTATION AND GENERAL EXPERIMENTAL CONDITIONS<br />

Reagents and Catalysts Materials<br />

The following metal salts and catalyst materials were obtained from<br />

Johnson Matthey (JM) (Ward Hill, MA): PtRu black (50 at. % Pt, 50 at. % Ru, JM<br />

PtRu, XRu - 0.5); Pt at 10 wt % metal loading on Vulcan XC-72R carbon (C/Pt, 10<br />

wt %); Pt-black, PtBr4 (99.99 %), RuCl3. x H2O (99.9 % purity, 42 wt % Ru<br />

content) and hexaammine ruthenium (III) trichloride. (Ru(NH3)6Cl3, 32.6%, Ru<br />

content). Concentrated H2SO4 and HClO4 were obtained in 99.999% purity from<br />

Aldrich (Milwaukee, WI). Potassium hydroxide (Aldrich), 1, 3- cyclohexadione<br />

(CHD, Aldrich), ammonium acetate (Aldrich), HCl (Mallinckrodt, Hazelwood, MO)<br />

and sodium nitrate (Mallinckrodt chemicals) were reagent grade or better and<br />

used as received. <strong>Methanol</strong> (99.9 % purity, Fisher Scientific, Hous<strong>to</strong>n, TX) was<br />

washed <strong>over</strong> alumina, filtered, distilled, and s<strong>to</strong>red refrigerated. Aqueous<br />

solutions containing methanol were prepared fresh the day of each experiment.<br />

Argon (Trinity Gases, Dallas, TX) was of ultrahigh purity. All solutions were<br />

prepared with deionized water (18 MΩ cm) from a four-cartridge Nanopure<br />

Infinity System (Barnstead, Dubuque, IA).<br />

Sonochemistry for Catalysts Synthesis<br />

PtRu alloy nanoparticles (SC PtRu) with various Ru surface compositions<br />

(XRu = 0.1, 0.25 and 0.5) were synthesized using a sonochemical (SC) method<br />

reported earlier [1]. The procedure is outlined here for reference. A Sonics and<br />

26


Materials Vibra Cell TM immersion sonica<strong>to</strong>r operating with 600 W electrical input<br />

power and calorimetrically measured 17 W/cm 2 acoustic powers [2,3] was used.<br />

The optimization of the sonication cell was performed either by using KI [4] or an<br />

8891 ultrasound cleaning bath (Cole-Parmer Instrument Company, operating<br />

frequency: 47 kHz). The cell temperature was maintained at 5º C while the<br />

cooling bath temperature was - 20 ºC.<br />

In the sonochemically prepared PtRu alloy synthesis procedure, there is a<br />

reaction between a mixture of Pt/Ru metal salts (PtBr4 and RuCl3) and lithium<br />

metal occurring in the sonication cell. Prior <strong>to</strong> synthesis, the cell was dried in an<br />

oven at 160º C, then purged with nitrogen and charged with tetrahydrofuran<br />

(THF, 15 mL), lithium metal cuttings (4.32 mmol) and naphthalene (2.81 mmol).<br />

A solution of 15 mL THF was prepared by dissolving an appropriate amount of<br />

PtBr4 and RuCl3 (0.05, 0.125 mol Ru 3+ and 0.45, 0.375 mol Pt 4+ for particles with<br />

XRu ≈ 0.1 and XRu ≈ 0.25 respectively and 0.25 mol in each metal for particles<br />

with XRu ≈ 0.5) in a previously dried (at 160º C for more than 6 hrs) Schlenk flask.<br />

After the dissolution of both metal halides, the resulting suspension looked black<br />

red-brown. The cell was sonicated for 2 hrs after cooling at -20º C for 15 min.<br />

During the sonication of cell containing lithium metal cutting, the RuCl3/PtBr4<br />

mixture was slowly introduced <strong>to</strong> the dark purple-blue lithium naphthalide THF<br />

solution passing through a cannula. The reaction mixture bubbled during the<br />

addition and ultimately turned black. The sonication was continued for 4 more<br />

hours.<br />

The reaction mixture was transferred <strong>to</strong> a 250 mL beaker in air. A black<br />

27


precipitate was collected after centrifuging the mixture. For washing purposes,<br />

the precipitate was rinsed with THF until it appeared colorless and odorless. The<br />

precipitate was then washed with 20–25 mL of deionized water. After 5–10 min<br />

stirring, the powder was finally filtered and centrifuged. The centrifuge tube was<br />

half filled with the suspension, and then filled in full by THF. When THF was<br />

added <strong>to</strong> the suspension, a few bubbles were observed. A compact black color<br />

precipitate with yellow–green supernatant was obtained after centrifuging for<br />

approximately 2 min. The precipitate was repeatedly washed by THF until a<br />

colorless supernatant was observed. The resulting metal black was collected with<br />

THF (no more than 30 mL), and transferred in<strong>to</strong> a vial equipped with a small<br />

magnetic stirring bar. The suspension was continually stirred in order <strong>to</strong> prevent it<br />

from nucleating. The procedure produced particles with diameters in the range of<br />

2–6 nm as observed by transmission electron microscopy.<br />

<strong>Bulk</strong> Electrode Preparation<br />

The following bulk materials were employed as working electrodes:<br />

polycrystalline platinum (Pt-solid, 0.81 cm 2 ), PtRu alloys (PtRu-solid) with Ru<br />

mole fractions (XRu) of 0.1 (0.71 cm 2 ), 0.3 (0.66 cm 2 ) and 0.9 (0.69 cm 2 ) and<br />

polycrystalline gold (Au) (0.67 cm 2 ). The preparation methods of the PtRu bulk<br />

alloys were described previously [5, 6]. In their work, the bulk alloys with same<br />

Ru compositions similar <strong>to</strong> our samples were observed <strong>to</strong> have the chemical<br />

composition and lattice constants which are shown in Table 2.1 [5]. The PtRu<br />

(PtRu-solid, XRu = 0.1 and 0.9) metal disks were received as polished with<br />

irregular oval shape. After receiving, the PtRu-solid, XRu = 0.3 disk was polished<br />

28


with emery paper (400 and 600 grit size) and then mirror polished with 9.0, 3.0,<br />

1.0, 0.3 and 0.05 µm alumina powder (Buehler) <strong>to</strong> get a smooth crystal faces.<br />

The Kel-F rod (1.25 cm diameter, Boedeker Plastics, Shiner, TX) was used as a<br />

supporting material for sealing the disks for fabricating electrodes. Intended for<br />

proper fastening, one end of Kel-F rod needed <strong>to</strong> be drilled according <strong>to</strong> the<br />

exact shape of the metal disks. For the asymmetric oval shaped disks (PtRu-<br />

solid, XRu = 0.1 and 0.3), a complicated procedure was followed <strong>to</strong> seal the edges<br />

of the electrodes for the present work. In the first step, the PtRu-solid, XRu = 0.1<br />

and 0.3 disks were scanned with a high resolution (600 dpi) scanner <strong>to</strong> get an<br />

identical model picture (Figure 2.1). From these scanned models, exact<br />

sketches that specify precise geometric dimensions of the peripheral part of the<br />

uneven disks were drawn with the help of a special 3D computer program<br />

(AUTODESK INVENTOR 1). This computer program efficiently scaled the<br />

original pho<strong>to</strong>graphs of the disks. Then the CNC (COMPUTER NUMERICAL<br />

CONTROL) milling machine was programmed with scaled dimension <strong>to</strong> cut the<br />

Kel-F rod. A hollow cavity that matched the original profile of irregular disk<br />

resulted. The disks were wrapped with Teflon tape (Zeus, Orangeburg, SC) <strong>to</strong><br />

<strong>over</strong>come any mismatch between the drilled Kel-F rod and the disk exterior<br />

structure. Each metal disk (either regular or irregular shaped) was pressure<br />

sealed at the edges in<strong>to</strong> the end of drilled Kel-F rod with the polished face<br />

exposed. In addition, a thin film of silicone sealant was applied <strong>to</strong> the disk edges<br />

<strong>to</strong> avoid further leaking of electrolyte solution on<strong>to</strong> the backside of the electrode.<br />

Then the electrodes were soaked in a stirred solution of 0.1M H2SO4 <strong>over</strong>night <strong>to</strong><br />

29


leach out the organic compounds which can possibly deteriorate the<br />

performances of the electrodes. The opposite end of rod was drilled <strong>to</strong> permit<br />

positioning of a Pt wire against the backside of the electrode disk. A film of silver<br />

epoxy (Dynaloy, Indianapolis, IN) secured the contact between the wire and disk.<br />

The Pt and Au disk working electrodes were periodically polished<br />

mechanically with alumina from 1.0 µm <strong>to</strong> down <strong>to</strong> 0.05 µm followed by<br />

sonication <strong>to</strong> remove debris. The PtRu surfaces were polished gently with 0.05<br />

µm alumina, cleaned by sonication in 5 M KOH, dipped briefly in 5 M H2SO4, and<br />

finally rinsed in ultra pure water prior <strong>to</strong> electrochemical measurements [7].<br />

Electrochemistry<br />

Reported cyclic voltammograms of the bulk PtRu and nanometer scale<br />

catalyst materials were recorded using a conventional glass electrochemical cell.<br />

The working and counter electrodes and reversible hydrogen electrode (RHE)<br />

reference were held in the same compartment and immersed in approximately 20<br />

mL of 0.1 M H2SO4.<br />

Electrolysis experiments were carried out using a electrochemical cell<br />

(Figure 2.2) that accommodated a small volume (50 µL) of sample [8]. The<br />

details of the cell and cell performances will be described in Chapter V. Briefly, a<br />

spectroelectrochemical cell was employed with a window that enabled the<br />

formation of a 50 µL thin layer. The window was constructed from a MACOR<br />

glass ceramic disk with the dimensions 2.54 cm diameter by 0.635 cm thickness<br />

(Accuratus, Phillipsburg, NJ). A cavity of 50 µL volume was milled in<strong>to</strong> the center<br />

of one face. A 2 mm recess cut in<strong>to</strong> the <strong>to</strong>p edge of the well allowed the working<br />

30


Table 2.1: Sample composition in a<strong>to</strong>mic % based on X-ray fluorescence and Xray<br />

Diffraction [5].<br />

Pt PtRu (XRu = 0.1) PtRu (XRu = 0.3) PtRu (XRu = 0.9)<br />

Ru % 0.0 9.7 29.8 90.5<br />

Structure fcc fcc fcc hcp<br />

a (Å) 3.9231 3.9166 3.8907 2.7178<br />

fcc-face centered cubic; hcp-hexagonal close packed.<br />

31


(b)<br />

(a)<br />

Figure 2.1: Scanned pho<strong>to</strong>graphs of bulk PtRu disks. (a) XRu= 0.1, (b) XRu= 0.3.<br />

32


electrode <strong>to</strong> be positioned reproducibly against the window with the metal disk<br />

facing the well and in contact with solution [8]. The electrode geometric area<br />

exposed <strong>to</strong> solution during electrocatalysis was approximately 0.74 cm 2 . A hole<br />

drilled through the disk center accommodated the transfer of solution in<strong>to</strong> a gas<br />

tight syringe (Hamil<strong>to</strong>n Company, Reno, Nevada) through a 23 gauge needle.<br />

The counter electrode consisted of a ring of Pt-wire and was positioned in the<br />

center of the cell behind the working electrode. The reference electrode was<br />

held in a separate compartment behind a wetted s<strong>to</strong>pcock that connected <strong>to</strong> the<br />

main body of the cell through a segment of glass tubing. A KCl saturated silver-<br />

silver chloride electrode (Ag/AgCl) was typically employed as the reference for<br />

electrolysis experiments. Potentials measured versus the Ag/AgCl reference<br />

were converted <strong>to</strong> the RHE scale and reported as volts versus RHE (VRHE).<br />

The response of the electrolysis cell (Figure 2.2) used for this work was<br />

characterized by performing cyclic voltammetry measurements using hexamine<br />

ruthenium (III) trichloride as a probe. Cyclic voltammograms of Ru(NH3)6Cl3 in<br />

0.1 M NaNO3 were recorded in a conventional electrochemical cell and in the thin<br />

layer cell. A polycrystalline gold electrode (0.481 cm 2 ) was used as the working<br />

electrode for these experiments. The Au was pressure sealed in<strong>to</strong> the end of<br />

Kel-F rod (Boedeker Plastics, Shiner, TX). The voltammetric measurements<br />

were performed by cycling between -0.3 and 0.4 VAg/AgCl with varying<br />

concentrations and scan rates. Cell performance <strong>to</strong> pure polycrystalline Pt<br />

electrode material (0.81 cm 2 ) were checked by recording cyclic voltammograms<br />

between 0.0 <strong>to</strong> 1.5 VRHE at different scan rates and changing the position of the<br />

33


Cell Body<br />

Working Electrode<br />

34<br />

Clamp<br />

50 L Cavity<br />

µ<br />

Gas Tight<br />

Syringe<br />

Macor Window<br />

Figure 2.2: Front portion of micro-volume electrochemical cell used for sample<br />

electrolysis.


electrode with respect <strong>to</strong> cell window (MACOR) position. The cell<br />

characterization will be described thoroughly in Chapter V.<br />

Prior <strong>to</strong> all electrochemical experiments, the solution in the<br />

electrochemical cell was degassed by bubbling with Ar for ca. 30 min. During<br />

measurements, Ar flowed in the space above the solution continuously. The<br />

procedure for sample electrolysis involved first degassing the 0.1M H2SO4<br />

electrolyte with the working electrode pulled away from the window, followed by<br />

cycling the electrode between 0.05 VRHE - 0.75 VRHE <strong>to</strong> verify cleanliness. The<br />

working electrode was then inserted in<strong>to</strong> the cavity with the potential held at 0.05<br />

VRHE. Using a gas tight syringe, 25 µL of solution was withdrawn from the cavity<br />

and replaced with an equal volume of solution containing 2.0 M CH3OH in 0.1 M<br />

H2SO4 <strong>to</strong> bring the CH3OH concentration in the cavity <strong>to</strong> 1.0 M. Following a 5 s<br />

equilibration period, the electrode was stepped <strong>to</strong> the reaction potential and held<br />

for 180 sec. At the end of the electrolysis period, 50 µL of solution was<br />

withdrawn from the cavity and diluted (by a fac<strong>to</strong>r 2-5) in 0.1 M H2SO4 <strong>to</strong> bring<br />

the H2CO concentration in<strong>to</strong> the range of 2-25 µM.<br />

<strong>Electrochemical</strong> cell potentials were maintained with a three electrode<br />

potentiostat (PC4/300, Gamry Instruments, Warminster, PA).<br />

For all bulk electrodes, surface cleanliness was checked by observing<br />

expected features in the hydrogen adsorption and oxide regions of a cyclic<br />

voltammogram recorded at 50 mV/s between 0.05 VRHE and 1.5 VRHE for Pt-solid;<br />

0.05 VRHE and 0.75 VRHE for PtRu alloys; 0.05 VRHE and 1.8 VRHE for Au<br />

electrode.<br />

35


Experiments with catalyst were performed by depositing a thin film of<br />

material on a cleaned, polished and optically flat Au electrode according <strong>to</strong><br />

literature procedures [9, 10]. Catalyst was suspended in ultra pure water (or THF<br />

for SC-PtRu material) <strong>to</strong> a known concentration of 2-4 mg/mL and dispensed with<br />

a micro-pipette on<strong>to</strong> the Au disk. If needed, the suspension (for the catalysts<br />

dispersed in water ) was evenly dispersed on the surface by adding additional<br />

water droplets and then the electrode surface was allowed <strong>to</strong> air dry for<br />

<strong>over</strong>night. The catalyst films were allowed <strong>to</strong> air dry for <strong>over</strong>night. For all cases<br />

the resulting electrode surface contained a thin catalyst layer. To remove<br />

surface contaminants, the electrode was cycled briefly between -0.2 VRHE and<br />

+1.5 VRHE (for C/Pt, 10 wt%; Pt- black), or 0.0 VRHE and 0.9 VRHE (for PtRu<br />

catalysts) until the voltammetry reached a steady state and waves characteristic<br />

of Pt and Ru appeared. The sonochemically prepared materials were activated<br />

prior <strong>to</strong> cycling <strong>to</strong> remove THF derived adsorbates by stepping the potential <strong>to</strong><br />

1.4 VRHE (PtRu, XRu = 0.1) or 1.1 VRHE (PtRu, XRu = 0.5) for 6 s.<br />

<strong>Formaldehyde</strong> Detection<br />

Dissolved H2CO was determined with a fluorescence assay [11].<br />

Following electrolysis, sample (50 µL) was reacted with 1, 3- cyclohexanedione<br />

in an ammonia/ammonium acetate buffer using a flow injection arrangement<br />

(Figure 2.3) [12, 13]. The reagent solution was prepared by mixing 25 mg of<br />

CHD, 16.5 g of ammonium acetate and 6.5 mL concentrated HCl (12 M) in 200<br />

mL of deionized water and then diluted <strong>to</strong> 250 mL. Reproducible results were<br />

obtained by allowing the reagent <strong>to</strong> stand <strong>over</strong>night. As discussed earlier [12],<br />

36


time is required <strong>to</strong> attain ke<strong>to</strong>/enol tau<strong>to</strong>meric equilibrium is attained. After<br />

injection, the sample (50 µL) and reagent were incubated for 3 minutes at 95 ºC<br />

and then transported <strong>to</strong> the flow- through fluorometer. The fluorometer was<br />

constructed around a liquid-core waveguide [14] utilizing Teflon-AF 2400 tubing<br />

(Biogeneral Fiber technology, 0.040" o.d, 0.034" i.d.). The sample was excited<br />

by a UV lamp (BF350-UV1, JKL components, Pacoima, CA) with maximum<br />

intensity at 365 nm. The fluorescence emission was collected at 460 nm through<br />

a 30 nm band-pass filter on the pho<strong>to</strong>detec<strong>to</strong>r (Intro, Socorro, NM) [14].<br />

The calibration curve for formaldehyde was recorded with freshly prepared<br />

standard solutions in the range of 2-25 µM. The linear calibration plot in Figure<br />

2.4 demonstrates the assay sensitivity and precision. The limit of detection as<br />

near 100 nM, which is an order of magnitude more sensitive than the lowest<br />

formaldehyde level, measured in methanol oxidation experiments. Figure 2.5<br />

shows the fluorescence signal for standard H2CO solution.<br />

37


CHD/water<br />

Reagent Pump<br />

Sample<br />

Injection Valve<br />

To Waste<br />

Thermostatted<br />

Reaction Coil<br />

Teflon AF<br />

(waveguide)<br />

PD<br />

Water Bath<br />

Optical Fiber<br />

Sample in<br />

38<br />

UV Lamp<br />

(365 nm)<br />

Flow-through<br />

Fluorometer<br />

Waste<br />

Figure 2.3: Block diagram of the au<strong>to</strong>mated H2CO analyzer.<br />

Flow<br />

Restric<strong>to</strong>r


Peak Current (nA)<br />

Figure 2.4: Calibration curve for standard<br />

H2CO solution. Concentration in the range of<br />

2x10 -6 -2.5x10 -5 M.<br />

Peak Current (nA)<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0<br />

y = 171.91x + 111.09, R 2 = 0.9987<br />

0 10 20 30<br />

Concentration (µM)<br />

200<br />

Time (Seconds)<br />

39<br />

400<br />

2 µM<br />

5 µM<br />

10 µM<br />

20 µM<br />

25 µM<br />

Figure 2.5: Fluorescence signal of standard H2CO solution.


References<br />

1. Korzeniewski, C.; Basnayake, R.; Vijayaraghavan, G.; Li, Z.; Xu, S.;<br />

Casadonte, D. J., Jr. Surf. Sci. 2004, 573, 100.<br />

2. Kotranarou, A.; Mills, G.; Hoffman, M. R. Environmental Science<br />

Technology 1992, 26, 2420.<br />

3. Mason, T. J., In Practical Sonochemistry- Users Guide <strong>to</strong> Applications in<br />

Chemistry and Chemical Engineering; Ellis Horwood Publishers, Ltd: New<br />

York, 1991, pp 46.<br />

4. Suslick, K. S.; Casadonte, D. J. J. Am. Chem. Soc. 1987, 109, 3459.<br />

5. Gasteiger, H. A.; Markovic, N.; Ross, P. N., Jr.; Cairns, E. J. J. Phys.<br />

Chem. 1993, 97, 12020.<br />

6. Gasteiger, H. A.; Markovic, N.; Ross, P. N., Jr.; Cairns, E. J. J.<br />

Electrochem. Soc. 1994, 141, 1795.<br />

7. Markovic, N. M.; Gasteiger, H. A.; Ross, P. N., Jr.; Jiang, X.; Villegas, I.;<br />

Weaver, M. J. Electrochim. Acta 1994, 40, 91.<br />

8. Huang, H.; Korzeniewski, C.; Vijayaraghavan, G. Electrochim. Acta 2002,<br />

42, 3675.<br />

9. Friedrich, K. A.; Henglein, F.; Stimming, U.; Unkauf, W. Colloids Surf. A.:<br />

Physiochem. Eng. Asp. 1998, 134, 193.<br />

10. Solla-Gullon, J.; Montiel, V.; Aldaz, A.; Clavilier, J. J. Electroanal. Chem.<br />

2000, 491, 69.<br />

11. Childers, C. L.; Huang, H.; Korzeniewski, C. Langmuir 1999, 15, 786.<br />

12. Fan, Q.; Dasgupta, P. K. Anal. Chem. 1994, 66, 551.<br />

13. Dong, S.; Dasgupta, P. K. Environ. Sci. Technol. 1987, 21, 581.<br />

14. Dasgupta, P. K.; Genfa, Z.; Li, J.; Boring, C. B.; Jambunathan, S.; Al-Horr,<br />

R. Anal. Chem. 1999, 71, 1400.<br />

40


CHAPTER III<br />

METHANOL ELECTROCHEMICAL CONVERSION TO FORMALDEHYDE<br />

OVER BULK METAL ELECTRODES<br />

Introduction<br />

Parallel reactions (Figure 1.4) during methanol electrooxidation at bulk<br />

metal electrodes lead <strong>to</strong> the formation of considerable amounts of H2CO <strong>to</strong>gether<br />

with carbon dioxide (CO2) [1, 2]. In this reports, H2CO yields from methanol<br />

electrochemical oxidation <strong>over</strong> smooth Pt and PtRu electrode materials at<br />

potentials in the range of 0.5-0.8 VRHE. The H2CO generated from 1.0 M<br />

methanol in 0.1 M H2SO4 at fixed potentials was determined with a small volume<br />

electrolysis arrangement following short electrolysis periods (180s) using a<br />

sensitive fluorescence assay [3, 4].<br />

With Pt electrodes, the H2CO yields approached 78% at 0.5 VRHE and<br />

decreased <strong>to</strong> 13% with increasing potentials (0.8 VRHE). At present, there is a<br />

general consensus that PtRu alloy materials are the most promising among all<br />

the binary catalysts for methanol oxidation. Exploring the bulk PtRu alloy<br />

materials is very important as this material is novel in the area of methanol<br />

electrochemical conversion <strong>to</strong> principal product (CO2) or side products (H2CO,<br />

formic acid, methyl formate, etc). The study demonstrates that the H2CO yield is<br />

lower (41%, 13% and 39% for XRu = 0.1, 0.3 and 0.9, respectively at 0.5<br />

VRHE) for PtRu electrodes compared <strong>to</strong> the Pt electrode. On bulk PtRu<br />

electrodes, the H2CO yields reached lowest values <strong>over</strong> the potential range<br />

41


studied (0.8 V – 0.5 VRHE) for the XRu = 0.3 alloy. The enhancement effect of Ru<br />

on Pt catalysts for the anodic methanol oxidation can be explained through<br />

bifunctional mechanism and already being described in Chapter I. The higher<br />

H2CO yield in bulk electrode materials relative <strong>to</strong> porous nanometer scale<br />

catalysts (Chapter IV) are believed <strong>to</strong> arise from the diffusion of partial oxidation<br />

products <strong>to</strong> the solution from smooth metal surfaces without further oxidization of<br />

H2CO <strong>to</strong> CO2.<br />

Experimental<br />

The polycrystalline Pt and PtRu bulk metal electrodes with three different<br />

Ru composition (XRu = 0.1, 0.3 and 0.9) were employed <strong>to</strong> investigate the percent<br />

yield of H2CO from methanol electrochemical oxidation. Just prior <strong>to</strong> the<br />

experiment, all the electrodes were mechanically and electrochemically polished<br />

and alloys were cleaned according <strong>to</strong> the procedures described in Chapter II. A<br />

KClsat Ag/AgCl electrode was employed for electrolysis experiment and a<br />

reversible hydrogen electrode was used for cyclic voltammetric measurementts.<br />

Short period (6s) conditioning of PtRu alloy electrodes were done at 1.4 VRHE<br />

until features appear in hydrogen adsorption desorption region. Cyclic<br />

voltammetric measurements were performed by cycling between 0.05 <strong>to</strong> 0.75<br />

VRHE and 0.05 <strong>to</strong> 1.5 VRHE for PtRu and Pt electrodes, respectively. The scan<br />

rate for cyclic voltammetry measurements was 50 mV/s. Electrolysis<br />

experiments were carried out for 180 s in a 50 µL, argon purged, electrochemical<br />

cell (Figure 2.2). The electrode was held at 0.0 VRHE during methanol injection.<br />

The studied potential range was 0.5-0.8 V RHE. Immediately after electrolysis, 50<br />

42


µL of solution was removed from the cell and diluted (dilution fac<strong>to</strong>r 2-5) with 0.1<br />

M H2SO4 <strong>to</strong> bring the H2CO concentration in<strong>to</strong> the range of 2-25 µM. Other<br />

experimental details are described in Chapter II.<br />

<strong>Formaldehyde</strong> Determination<br />

H2CO was determined fluorometrically following reaction with 1, 3-<br />

cyclohexadione (see Chapter II) using the flow injection arrangement shown in<br />

Figure 2.3.<br />

Results and Discussion<br />

Figures 3.1 and 3.2 show cyclic voltammograms of the PtRu and Pt bulk<br />

electrode materials, respectively, recorded in a conventional three electrode<br />

electrochemical cell. The voltammograms recorded in 0.1M H2SO4 demonstrate<br />

the background response of the materials in aqueous acid solution and the<br />

cleanliness of the cell, electrolyte and electrode surfaces. Figure 3.1 also<br />

represents the comparative cyclic voltammograms of Pt (Figure 3.1a) and PtRu<br />

with XRu = 0.1, 0.3 and 0.9 (Figure 3.1 b, c, d). The positive potential was<br />

restricted <strong>to</strong> 0.75 VRHE <strong>to</strong> minimize oxidation-induced changes in the surface Ru<br />

composition of PtRu materials [5]. In cyclic voltammetry studies, mainly<br />

highlighted are (i) the hydrogen region, which contains the hydrogen adsorption<br />

and hydrogen desorption peaks in the potential range 0.0-0.3 VRHE, and (ii)<br />

double layer region in the potential range of 0.3-0.7 VRHE. These potential<br />

regions are important and the most meaningful in the context of cyclic<br />

voltammetry of pure Pt [6]. Therefore, the voltammogram (Figure 3.1 a) recorded<br />

43


with a polycrystalline Pt electrode can be considered as a model <strong>to</strong> describe<br />

features in these key potential regions. Sharp peaks due <strong>to</strong> hydrogen adsorption<br />

(Pt + H + + e - →Pt-H) and hydrogen desorption (Pt-H → Pt + H + + e - ) are observed<br />

between ~ 0.05 and 0.35 VRHE and are a characteristic of the presence of Pt.<br />

The double layer charging region occurs between about 0.35 and 0.7 VRHE and is<br />

narrow for Pt. Figure 3.2 shows a cyclic voltammogram of Pt when the positive<br />

potential was extended <strong>to</strong> 1.5 VRHE. In addition <strong>to</strong> the hydrogen region and<br />

double layer charging region, features typical for polycrystalline Pt appear in the<br />

(0.0 <strong>to</strong> 0.4 VRHE) oxide potential region (0.8 <strong>to</strong> 1.4 VRHE). Here, an oxide stripping<br />

peak was observed on the negative sweep at about 0.75 VRHE, and at about 1.0<br />

VRHE on the forward scan, a shoulder wave associated with Pt oxidation (Pt +<br />

H2O → PtOH + H + + e - ) appears.<br />

For clarity, the above mentioned potential regions will be referred <strong>to</strong> in the<br />

discussion of cyclic voltammetry of Pt-Ru alloys. Cyclic voltammograms for PtRu<br />

alloy electrode materials are shown in Figure 3.1 (b), (c), (d). The responses<br />

demonstrate the effect of increasing the bulk and surface composition of Ru in<br />

the PtRu alloy electrodes. The cyclic voltammetry results enabled comparisons<br />

of the hydrogen regions and the double layer regions for the bulk alloy electrodes<br />

and the pure Pt electrode. While the Pt electrode displayed the typical hydrogen<br />

adsorption and desorption peaks, as well-defined surface oxidation and reduction<br />

peaks (Figure 3.1 a and Figure 3.2), the intermetallic alloy electrodes typically<br />

showed very little, if any, evidence of hydrogen adsorption and desorption<br />

depending on the Pt/Ru ratio. For all PtRu alloy electrodes, the hydrogen regions<br />

44


(at ~ 0.05 <strong>to</strong> 0.35 VRHE) and double layer regions (0.35 <strong>to</strong> 0.7 VRHE) were<br />

observed in the same scan range as the polycrystalline Pt electrode. The<br />

addition of Ru is accompanied by major changes, which include an increase in<br />

the double layer capacitance, a decrease in the hydrogen adsorption peaks and<br />

a change in the oxide stripping peak. The PtRu alloy with XRu = 0.1 (Figure 3.1b)<br />

displayed features that are most similar <strong>to</strong> those of polycrystalline Pt among the<br />

alloy samples studied [6, 7]. The incorporation of a small amount of Ru (XRu =<br />

0.1) caused the hydrogen adsorption and hydrogen desorption peaks <strong>to</strong><br />

diminished slightly and the double layer region <strong>to</strong> broaden compared <strong>to</strong> bulk Pt<br />

electrode. Increasing the Ru amount (XRu = 0.3) (Figure 3.1c), lead <strong>to</strong> a larger<br />

double layer thickness and a significant decline in the sharpness of the hydrogen<br />

adsorption area. The voltammetry of the alloy with XRu = 0.9 composition (Figure<br />

3.1d) does not display distinct peaks in the hydrogen adsorption region. It is<br />

instead dominated by a large capacitive current that extends in<strong>to</strong> the double-<br />

layer region. These findings are consistent with observations for bulk PtRu<br />

electrodes with different Ru surface compositions (XRu = 0.07, 0.33 and 0.46) and<br />

have been attributed <strong>to</strong> the presence of ruthenium oxide on the electrode<br />

surface, and the adsorption of oxygen like species extending progressively<br />

<strong>to</strong>ward more negative potentials with increasing Ru composition[6-8]. This<br />

phenomenon plays a significant role in the conversion efficiency of methanol <strong>to</strong><br />

H2CO, which will be addressed in discussions that follow.<br />

All results are summarized in Tables 3.1 and Figure 3.3 <strong>to</strong> Figure 3.6.<br />

Table 3.1 lists the charge passed, the H2CO quantity produced and the resulting<br />

45


H2CO yield for the experiments. Below 0.5 VRHE, the % yield value could not be<br />

determined because the reaction current, and thus the amount of charge passed,<br />

was <strong>to</strong>o low <strong>to</strong> be measured. Results for the studied potentials are explained<br />

below. Special emphasis will be on the lowest positive potentials (0.6 and 0.5<br />

VRHE) studied as one of the goals of this work was <strong>to</strong> investigate the reaction<br />

pathways at low methanol oxidation <strong>over</strong> potentials. The computation of percent<br />

conversion was performed by taking the ratio of the charge required <strong>to</strong> produce<br />

the quantity of H2CO measured with the fluorescence technique <strong>to</strong> the <strong>to</strong>tal<br />

charge passed during electrolysis and multiplying the result by 100. In general,<br />

the H2CO is one of the most dominant by-products of methanol oxidation for<br />

reactions on bulk metal electrodes with smooth surfaces [9-13].<br />

Our study with bulk electrode materials (Table 3.1) shows that the H2CO<br />

yield generally decreases as the reaction potential is stepped positive. Figure<br />

3.3, 3.4, 3.5 and Figure 3.6 show the graphical representation of the H2CO yield<br />

with respect <strong>to</strong> potential for the Pt-solid and the PtRu-solid with XRu = 0.1, 0.3<br />

and 0.9 bulk metal electrodes respectively. The <strong>to</strong>pmost entries in Table 3.1 are<br />

for the Pt- solid electrode (see also Figure 3.3), which showed the highest H2CO<br />

yields for the considered working potentials (0.8 <strong>to</strong> 0.5 VRHE) than those of the<br />

PtRu- solid metal electrodes (2 nd , 3 rd and 4 th entries of Table 3.1). For pure Pt,<br />

the lowest H2CO yield (13%) was observed at 0.8 VRHE. The formation of H2CO<br />

<strong>over</strong> Pt gradually increased <strong>to</strong>ward lower electrode potentials up <strong>to</strong> 0.5 V where it<br />

attained 78% of the <strong>to</strong>tal electrolysis charge. The H2CO yield started <strong>to</strong> become<br />

most significant at potentials lower than 0.7 VRHE. This general behavior was<br />

46


Current<br />

0<br />

0<br />

0<br />

0<br />

0.0<br />

0.02 mA / cm 2<br />

E / V vs. RHE<br />

47<br />

(a) Pt - solid<br />

(b) PtRu - solid (XRu = 0.1)<br />

(c) PtRu - solid (XRu = 0.3)<br />

(d) PtRu - solid (XRu = 0.9)<br />

0.2<br />

0.4<br />

0.6 0.8<br />

Figure 3.1: Cyclic voltammograms of bulk electrode materials employed in the<br />

study. Scans were recorded in Ar purged 0.1 M H2SO4. Polycrystalline Pt<br />

electrode (Pt-solid, a); bulk PtRu alloys (PtRu-solid) with Ru mole fraction (XRu) of<br />

0.1 (b), 0.3 (c) and 0.9 (d). Cyclic voltammetric measurements were performed<br />

by scanning the potential between 0.05 <strong>to</strong> 0.75 VRHE. The scan rate was 50 mV/s.


Current<br />

Polycrystalline Pt-solid Electrode<br />

in 0.1 M H2SO4<br />

0.05 mA / cm 2<br />

0<br />

0.0 0.4 0.8 1.2<br />

E / V vs. RHE<br />

Figure 3.2: Cyclic voltammogram of a polycrystalline Pt electrode in Ar purged<br />

0.1 M H2SO4 recorded at a scan rate of 50 mV/s.<br />

48<br />

1.6


observed previously by our research group [9, 14]. These studies employed a<br />

lower methanol concentration, but the maximum H2CO formation was near 30%<br />

at 0.5 VRHE and dropped <strong>to</strong> less than 10% at higher potentials [9, 14]. Using<br />

HPLC, which is similar <strong>to</strong> fluorescence in its ability <strong>to</strong> quantify H2CO in this<br />

application, Iwasita and co-workers measured percent yield of 71% and 81% for<br />

H2CO produced at 0.6 VRHE at polycrystalline Pt and Pt(111) electrodes,<br />

respectively [2]. Furthermore, using the DEMS (Differential <strong>Electrochemical</strong><br />

Mass Spectrometry) technique Wang and co-workers observed H2CO yields of<br />

up <strong>to</strong> 72% in reactions at 0.75 VRHE on a smooth polycrystalline Pt electrode in<br />

0.5 M H2SO4 containing 0.01M methanol [10, 13]. Our results demonstrate, the<br />

percent conversion of methanol <strong>to</strong> H2CO on a polycrystalline Pt electrode<br />

increases as the reaction potential decreases between 0.8V and 0.5 VRHE. The<br />

result is consistent with conclusions drawn by the Wieckowski group recently<br />

[15], in a study that confirmed methanol decomposition on Pt proceeds effectively<br />

through the formation of soluble products at low potentials and that methanol<br />

dehydrogenation <strong>to</strong> H2CO can accounts for a majority of the oxidation charge.<br />

More<strong>over</strong>, the decreasing yield of H2CO observed in our study at potentials<br />

greater than 0.5 VRHE (Table 3.1 and Figure 3.3) is attributable <strong>to</strong> the facile water<br />

activation which takes place at these potentials on Pt and can accelerate CO2<br />

formation from methanol [16].<br />

In contrast <strong>to</strong> Pt materials, PtRu bulk alloys show evidence of much lower<br />

secondary product (H2CO) formation, particularly below 0.6 VRHE. The superior<br />

catalytic performance of PtRu binary alloy materials relative <strong>to</strong> Pt is well known<br />

49


and it is explained by the term “bi-functional mechanism” [6, 17, 18]. This term<br />

put emphasis on the joint activities of both metals, Pt being the one adsorbing<br />

methanol and cleaving the C-H bond and Ru, the one that promotes the<br />

dissociation of water and oxidizes the adsorbed residues at low potentials<br />

compared <strong>to</strong> pure Pt. Similar <strong>to</strong> Pt- solid, the PtRu - solid (XRu = 0.1) electrode<br />

(Figure 3.4 and 2 nd entries of Table 3.1) displays an increase in H2CO yield in<br />

progressing from higher <strong>to</strong> lower electrode potentials (from 0.8 VRHE <strong>to</strong> 0.5 VRHE).<br />

In the case of the lowest working potential (0.5 VRHE), the H2CO yield (41%) is<br />

reduced substantially from that of the bulk platinum electrode (78%). A very low<br />

yield of H2CO (~6.2%) and higher reaction charge were found for the highest<br />

reaction potential (0.8 VRHE). At 0.6 VRHE, relatively lower yield (32%) of H2CO<br />

was achieved if compared <strong>to</strong> the yield on pure Pt (38%) but the reaction yield<br />

remains almost the same for both electrodes at 0.7 VRHE. The results indicate<br />

that incorporating a small amount of Ru in<strong>to</strong> Pt enhances the catalytic activity of<br />

the surface with respect <strong>to</strong> the oxidation of methanol beyond H2CO. This result<br />

will be discussed in greater detail later in the chapter.<br />

Of all the bulk alloys, PtRu solid (XRu = 0.3) showed the best catalytic<br />

performance and lowest H2CO yields (3 rd entries of Table 3.1 and Figure 3.5) for<br />

all studied working potentials (0.8 <strong>to</strong> 0.5 VRHE). On the lowest Ru content alloy<br />

(XRu = 0.1), the methanol surface chemistry was similar <strong>to</strong> that of polycrystalline<br />

Pt, and there is only a small change of H2CO yield above 0.5 VRHE. With the<br />

inclusion of a medium amount of Ru (XRu = 0.3), drastic changes were observed<br />

in reaction charges, H2CO quantity and H2CO yields. Comparing the yield (see<br />

50


Figure 3.7) for the lowest two working potentials (0.6 and 0.5 VRHE), PtRu solid<br />

with XRu = 0.3 produced appreciably lower yields (8.4% and 13% for 0.6 and 0.5<br />

VRHE, respectively) than those of PtRu solid with XRu = 0.1 (32% and 41% for 0.6<br />

and 0.5 VRHE, respectively) and pure Pt (38% and 78% for 0.6 and 0.5 VRHE,<br />

respectively). Here the role of Ru-ada<strong>to</strong>ms is accelerating the methanol<br />

oxidation rate on smooth Pt at low potentials and the high c<strong>over</strong>age of oxides on<br />

Ru sites is expected <strong>to</strong> enable the conversion of H2CO <strong>to</strong> products with greater<br />

oxygen content, such as HCOOH and CO2. Also, the PtRu-solid (XRu = 0.3)<br />

electrode used in our study showed excellent catalytic activity in the H2CO yield<br />

(~3-5%) for higher electrode potentials (0.8 and 0.7 VRHE). The very low yields<br />

for higher electrode potentials are due <strong>to</strong> the combined effect of higher driving<br />

force for the reaction and enhancement of water activation by Ru a<strong>to</strong>ms.<br />

The measurement with PtRu-solid, XRu = 0.9 showed higher average<br />

H2CO yields (4 th entries of Table 3.1 and Figure 3.6) compared <strong>to</strong> that of PtRu-<br />

solid, XRu = 0.3 electrode for all potentials (0.8 <strong>to</strong> 0.5 VRHE). The H2CO yields<br />

were significantly higher at 0.6 and 0.5 VRHE for PtRu-solid, XRu = 0.9 (16% and<br />

39% respectively) relative <strong>to</strong> PtRu-solid, XRu = 0.3 (8.4% and 13% respectively).<br />

The lower activity of the PtRu solid with XRu = 0.9 <strong>to</strong>wards methanol oxidation<br />

relative <strong>to</strong> the PtRu solid with XRu = 0.3 can be explained by the lower availability<br />

of Pt a<strong>to</strong>ms for methanol dehydrogenation, which requires an ensemble, namely,<br />

a specific number of Pt sites for methanol decomposition [7, 19]. The alloy with<br />

higher Ru mole fraction contains a passivating surface layer, probably Ru oxides,<br />

which can inhibit methanol dissociative chemisorption at ambient temperature<br />

51


[20]. The jump in H2CO yield measured in reactions <strong>over</strong> PtRu, XRu = 0.9 at 0.5<br />

V may reflect a decrease in the rate of surface oxide attack on carbon containing<br />

fragments at the low electrochemical driving force <strong>to</strong>gether with an inhibition of<br />

C-H bond breaking steps by the oxides. These effects would create a trap for<br />

accumulation of H2CO by reducing the fraction of reacting CH3OH molecules that<br />

progress from H2CO <strong>to</strong> HCOOH, or from H2CO <strong>to</strong> CO and subsequently CO2.<br />

These findings indicate that the bulk alloys with the high Pt or Ru composition is<br />

not as active <strong>to</strong>wards the methanol oxidation as the alloy containing a medium<br />

(XRu = 0.3) amount of Ru.<br />

The data in Table 3.1 indicate that the PtRu solid with small amount of Ru<br />

(XRu = 0.1) behaves almost like pure Pt with respect <strong>to</strong> H2CO yields, whereas<br />

incorporating somewhat more Ru (XRu = 0.3), improves the catalytic activity of<br />

the alloy <strong>to</strong>ward methanol oxidation significantly. In literature reports, at ambient<br />

temperatures PtRu electrodes with low Ru surface concentration (XRu = 0.1- 0.3)<br />

show the best kinetics for the electrooxidation of methanol (ranging from 0.005 <strong>to</strong><br />

0.5 M) [6, 7]. Our PtRu - solid (XRu = 0.1) is highly active at the most positive<br />

potentials and also showed better performances for lowest electrode potentials<br />

regarding the H2CO yield compared <strong>to</strong> pure Pt electrode. The current results<br />

indicate that the PtRu solid with XRu = 0.3 is the best performing electrode among<br />

those studied for 1.0 M methanol oxidation due <strong>to</strong> the small H2CO yields shown<br />

at low reaction potentials in combination with high reaction currents. The Pt/Ru<br />

ratio in alloy materials is not the only fac<strong>to</strong>r for attaining the best catalytic activity.<br />

Reactant concentration also plays a role. In earlier work by our research group,<br />

52


Table 3.1: Summary of charge and formaldehyde yields from the oxidation of 1.0<br />

M CH3OH on bulk Pt and PtRu electrode materials a .<br />

Electrode Materials b Potential c<br />

(E/V vs.<br />

RHE)<br />

Average<br />

Charge<br />

(mC)<br />

53<br />

Average H2CO<br />

Quantity (nmol)<br />

H2CO<br />

Yield %<br />

Pt-solid 0.8 3.08 2.12 13 ± 4<br />

0.7 4.33 4.5 20 ± 3<br />

0.6 5.69 10.7 38 ± 7<br />

0.5 0.33 1.33 78 ± 4<br />

PtRu – solid (XRu = 0.1) 0.8 8.63 2.89 6.2 ± 3.1<br />

0.7 2.39 2.82 23 ± 3<br />

0.6 1.77 3.14 32 ± 7<br />

0.5 0.41 0.85 41 ± 8<br />

PtRu – solid (XRu = 0.3) 0.8 25.95 4.65 3.5±0.1<br />

0.7 10.3 1.69 3.8±1.8<br />

0.6 14.1 3.86 8.4 ± 5.3<br />

0.5 3.86 2.10 13 ± 4<br />

PtRu – solid (XRu = 0.9) 0.8 4.11 1.27 6.5 ± 2.1<br />

0.7 7.02 2.06 5.8 ± 0.7<br />

0.6 3.20 2.43 16 ± 5<br />

0.5 1.51 3.02 39 ± 2<br />

a All experiments involved the electrolysis of 1.0 M CH3OH in 0.1 M H2SO4 in a volume of 50 µL for a period<br />

of 180s. The values reported for the charge and H2CO amount are averages calculated from results of 3-5<br />

trials. The yield values were computed from the ratio of the charge required <strong>to</strong> produce H2CO <strong>to</strong> the <strong>to</strong>tal<br />

charge passed during the electrolysis period. These yield values were averaged <strong>to</strong> give the reported<br />

percentages and uncertainties. The uncertainties represent 95.5% confidence limits based on two times the<br />

sample standard deviation. b The following electrode materials were used: Polycrystalline Pt electrode (Ptsolid);<br />

bulk PtRu alloys (PtRu-solid) with Ru mole fraction (XRu) of 0.1, 0.3 and 0.9. c The reference electrode<br />

was KCl saturated Ag/AgCl. For reporting purposes, all the measured potentials vs. Ag/AgCl were converted<br />

<strong>to</strong> the reversible hydrogen electrode (RHE) scale and expressed as volts versus RHE (VRHE).


% H 2CO yield<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

77.85<br />

37.67<br />

54<br />

19.81<br />

12.61<br />

0<br />

0.45 0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85<br />

E/V vs. RHE<br />

Figure 3.3: H2CO yields from methanol electrochemical oxidation on a<br />

polycrystalline Pt electrode. All experiments involved the electrolysis of 1.0 M<br />

CH3OH in 0.1 M H2SO4 in a volume of 50 µL for a period of 180s.


% H 2CO yield<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

41.42<br />

32.32<br />

55<br />

23.09<br />

0<br />

0.45 0.5 0.55 0.6 0.65<br />

E/V vs. RHE<br />

0.7 0.75 0.8 0.85<br />

Figure 3.4: H2CO yields from methanol electrochemical oxidation on a PtRu-<br />

solid (XRu = 0.1) electrode. All experiments involved the electrolysis of 1.0 M<br />

CH3OH in 0.1 M H2SO4 in a volume of 50 µL for a period of 180s.<br />

6.21


% H 2CO yield<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

13<br />

8.41<br />

0<br />

3.8<br />

3.52<br />

0.45 0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85<br />

E/V vs. RHE<br />

Figure 3.5: H2CO yields from methanol electrochemical oxidation on a PtRu-<br />

solid (XRu = 0.3) electrode. All experiments involved the electrolysis of 1.0 M<br />

CH3OH in 0.1 M H2SO4 in a volume of 50 µL for a period of 180s.<br />

56


% H 2CO yield<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

38.73<br />

15.65<br />

0<br />

0.45 0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85<br />

E/V vs. RHE<br />

Figure 3.6: H2CO yields from methanol electrochemical oxidation on a PtRu-<br />

solid (XRu = 0.9) electrode. All experiments involved the electrolysis of 1.0 M<br />

CH3OH in 0.1 M H2SO4 in a volume of 50 µL for a period of 180s.<br />

57<br />

5.8<br />

6.5


% H 2CO yield<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

Pt-solid PtRu-solid (XRu=0.1) PtRu-solid (XRu=0.3) PtRu-solid (XRu=0.9)<br />

Electrode Composition<br />

58<br />

0.6 V vs.RHE<br />

0.5 V vs.RHE<br />

Figure 3.7: Comparison of H2CO yields from methanol (1.0 M) electrochemical<br />

oxidation on bulk electrode materials at 0.6 and 0.5 VRHE.


it was shown that the surface c<strong>over</strong>age of adsorbed residues from methanol<br />

decomposition, especially CO, appears <strong>to</strong> increase on Pt surfaces with<br />

increasing methanol concentration. Poisoning by the adsorbed species is<br />

assumed <strong>to</strong> be responsible for limiting CO2 formation below about 0.8 VRHE [21].<br />

In addition, the Ru composition influences the balance between CO2 and H2CO<br />

yields from methanol and Iwasita and co-workers [22] reported that an improved<br />

catalytic effect can be obtained if the final oxidation step ((CO)ad +<br />

(OH)ad→CO2+H + +e - ) occurs between CO adsorbed at Pt and OH adsorbed at Ru<br />

(Pt(CO)ad + Ru(OH)ad→CO2+H + +e - ) [22]. In our studies of H2CO formation at the<br />

PtRu (XRu = 0.1) alloy, 10% Ru a<strong>to</strong>ms seems inadequate <strong>to</strong> provide sufficient<br />

hydroxide for oxidizing the high amount of adsorbed residues expected. Our<br />

results are in good agreement with the work of Iwasita, et al, which showed that<br />

the rate of methanol oxidation <strong>to</strong> CO2 (Figure 1.4) is faster at XRu = 0.5 than at<br />

XRu = 0.15 PtRu alloy [23, 24] even though the XRu = 0.15 surface gave higher<br />

current densities for methanol oxidation. These workers concluded that the lower<br />

Ru content alloy produced a greater yield of H2CO, consistent with our findings.<br />

With our techniques for electrolysis and H2CO determinations, it is<br />

possible <strong>to</strong> make repetitive measurements simply and rapidly. Figure 3.8 and 3.9<br />

summarize results of the analysis of data for polycrystalline Pt from a series of<br />

repetitive electrolysis experiments at fixed potentials. Figure 3.8 shows the plot<br />

of nanomoles of H2CO formed versus the reaction charge passed following the<br />

sample electrolysis. The Y-axis of the plot in Figure 3.8 gives the H2CO quantity<br />

determined from the intensity of H2CO bands in comparison <strong>to</strong> a calibration curve<br />

59


(see Figure 2.4). At each potential, the error bars represent 95% confidence<br />

intervals on the mean of 3 or 4 repetitive measurements. Figure 3.9 reports the<br />

H2CO formation rate as a function of potential. The ability <strong>to</strong> make such<br />

determinations enables the significance of reaction rate changes due <strong>to</strong> the<br />

different electrode properties (i.e., composition, surface structure) <strong>to</strong> be assessed<br />

statistically, and uncertainties <strong>to</strong> be assigned <strong>to</strong> kinetic and thermodynamic<br />

parameters derived from the data. In the same manner, this type of data<br />

analysis is summarized for the PtRu alloy electrodes in Figure 3.10 and Figure<br />

3.11 (XRu = 0.1); Figure 3.12 and Figure 3.13 (XRu = 0.3); and Figure 3.14 and<br />

Figure 3.15 (XRu = 0.9).<br />

In general Figure 3.8 (Pt electrode), Figure 3.10 (PtRu-solid, XRu = 0.1<br />

electrode), Figure 3.12 (PtRu-solid, XRu = 0.3) and Figure 3.14 (PtRu-solid, XRu =<br />

0.9) compare uncertainty in H2CO determinations <strong>to</strong> the error in corresponding<br />

coulometry measurements. The errors relative <strong>to</strong> the averages are larger for the<br />

coulometry than the fluorescence spectroscopy measurements. Under the<br />

conditions of the experiments, the precision of the fluorescence technique<br />

outperforms that of the electrochemistry.<br />

From the graphs of the rate of H2CO formation versus electrode potential<br />

shown in Figure 3.9 (Pt electrode), Figure 3.11 (PtRu-solid, XRu = 0.1 electrode),<br />

Figure 3.13 (PtRu-solid, XRu = 0.3) and Figure 3.15 (PtRu-solid, XRu = 0.9), the<br />

general trend is that the rate of formation of H2CO is highest at the lower<br />

electrode potentials, 0.5 and 0.6 VRHE.<br />

60


Conclusions<br />

Although the in situ IR and DEMS methods are highly efficient in the study<br />

of competing methanol electrooxidation pathways, they are incapable of the<br />

direct detection of H2CO. As we [9, 14] and others [2, 10, 13] have shown, H2CO<br />

can be generated at low potentials in reactions on bulk electrode surfaces and<br />

should be considered as an important intermediate in mechanistic work. The<br />

reported fluorescence assay had sufficient sensitivity <strong>to</strong> track H2CO produced in<br />

a micro volume cell employing polycrystalline Pt and bulk PtRu alloy working<br />

electrodes <strong>over</strong> short electrolysis periods (180 s). The catalytic activity of PtRu<br />

electrodes <strong>to</strong>ward methanol oxidation was superior <strong>to</strong> polycrystalline Pt, but the<br />

yields of H2CO were strongly dependent on the Ru content. The lowest H2CO<br />

yield was attained for the alloy electrode with average Ru composition (XRu = 0.3)<br />

<strong>over</strong> the potential range studied (0.8 V – 0.5 VRHE). The PtRu- solid (XRu = 0.3)<br />

electrode is considered <strong>to</strong> be in the range of optimum composition for the<br />

electrochemical efficient oxidation of methanol in PEM type fuel cells.<br />

61


<strong>Formaldehyde</strong> quantity(nmol)<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0.5 V RHE<br />

0<br />

0 1 2 3 4 5 6 7<br />

charge (mC)<br />

62<br />

0.8 V RHE<br />

0.7 V RHE<br />

0.6 V RHE<br />

Figure 3.8: Plot of the nanomoles H2CO formed vs. the reaction charge passed<br />

following the oxidation of 1.0 M methanol on a polycrystalline Pt electrode in 0.1<br />

M H2SO4 (Reaction time = 180 s).


Rate (nmol/s/cm 2 )<br />

0.08<br />

0.07<br />

0.06<br />

0.05<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

0<br />

0.4 0.45 0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85<br />

E/V vs. RHE<br />

Figure 3.9: Plot of H2CO formation rate averaged <strong>over</strong> 180 s electrolysis periods<br />

vs. potential for the oxidation of 1.0 M methanol in 0.1 M H2SO4 on a<br />

polycrystalline Pt electrode.<br />

63


<strong>Formaldehyde</strong> quantity (nmol)<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0.5 V RHE<br />

0.6 V RHE<br />

0.7 V RHE<br />

0<br />

0 1 2 3 4 5<br />

Charge (mC)<br />

6 7 8 9 10<br />

64<br />

0.8 V RHE<br />

Figure 3.10: Plot of the nanomoles H2CO formed vs. the reaction charge passed<br />

following the oxidation of 1.0 M methanol on a bulk alloy PtRu- solid (XRu = 0.1)<br />

electrode in 0.1 M H2SO4 (Reaction time = 180 s).


Rate (nmol/s/cm 2 )<br />

0.025<br />

0.02<br />

0.015<br />

0.01<br />

0.005<br />

0<br />

0.4 0.45 0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85<br />

E/V vs. RHE<br />

Figure 3.11: Plot of H2CO formation rate averaged <strong>over</strong> 180 s electrolysis periods<br />

vs. potential for the oxidation of 1.0 M methanol in 0.1 M H2SO4 on a bulk alloy<br />

PtRu- solid (XRu = 0.1) electrode.<br />

65


<strong>Formaldehyde</strong> Quantity (nmol)<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0.5 V RHE<br />

0.7 V RHE<br />

0.6 V RHE<br />

0<br />

0 5 10 15<br />

Charge (mC)<br />

20 25 30<br />

66<br />

0.8 V RHE<br />

Figure 3.12: Plot of the nanomoles H2CO formed vs. the reaction charge passed<br />

following the oxidation of 1.0 M methanol on a bulk alloy PtRu- solid (XRu = 0.3)<br />

electrode in 0.1 M H2SO4 (Reaction time = 180 s).


Rate (nmol/s/cm 2 )<br />

0.04<br />

0.035<br />

0.03<br />

0.025<br />

0.02<br />

0.015<br />

0.01<br />

0.005<br />

0<br />

0.4 0.45 0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85<br />

E/V vs RHE<br />

Figure 3.13: Plot of H2CO formation rate averaged <strong>over</strong> 180 s electrolysis periods<br />

vs. potential for the oxidation of 1.0 M methanol in 0.1 M H2SO4 on a bulk alloy<br />

PtRu- solid (XRu = 0.3) electrode.<br />

67


<strong>Formaldehyde</strong> quantity (nmol)<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0.5 V RHE<br />

0.6 V RHE<br />

0<br />

0 1 2 3 4<br />

Charge (mC)<br />

5 6 7 8<br />

68<br />

0.8 V RHE<br />

0.7 V RHE<br />

Figure 3.14: Plot of the nanomoles H2CO formed vs. the reaction charge passed<br />

following the oxidation of 1.0 M methanol on a bulk alloy PtRu- solid (XRu = 0.9)<br />

electrode in 0.1 M H2SO4 (Reaction time = 180 s).


Rate (nmol/s/cm 2 )<br />

0.03<br />

0.025<br />

0.02<br />

0.015<br />

0.01<br />

0.005<br />

0<br />

0.45 0.5 0.55 0.6 0.65<br />

E/V vs. RHE<br />

0.7 0.75 0.8 0.85<br />

Figure 3.15: Plot of H2CO formation rate averaged <strong>over</strong> 180 s electrolysis periods<br />

vs. potential for the oxidation of 1.0 M methanol in 0.1 M H2SO4 on a bulk alloy<br />

PtRu- solid (XRu = 0.9) electrode.<br />

69


References<br />

1. Cao, D.; Lu, G.-Q.; Wieckowski, A.; Wasileski, S. A.; Neurock, M. J.<br />

Phys.Chem. B. 2005 , 109 , 11622.<br />

2. Batista, E. A.; Malpass, G. R. P.; Motheo, A. J.; Iwasita, T. J. Electroanal.<br />

Chem. 2004 , 571 , 273.<br />

3. Dong, S.; Dasgupta, P. K. Environ. Sci. Technol. 1987 , 21 , 581.<br />

4. Nash, T. Biochem. J. 1953 , 55 , 416.<br />

5. Markovic, N. M.; Gasteiger, H. A.; Ross, P. N., Jr.; Jiang, X.; Villegas, I.;<br />

Weaver, M. J. Electrochim. Acta 1994 , 40 , 91.<br />

6. Gasteiger, H. A.; Markovic, N.; Ross, P. N., Jr.; Cairns, E. J. J. Phys.<br />

Chem. 1993 , 97 , 12020.<br />

7. Gasteiger, H. A.; Markovic, N.; Ross, P. N., Jr.; Cairns, E. J. J. Phys.<br />

Chem. 1994 , 98 , 617.<br />

8. Gasteiger, H. A.; Markovic, N. M.; Ross, P. N., Jr. J. Phys. Chem. 1995 ,<br />

99 , 16757.<br />

9. Childers, C. L.; Huang, H.; Korzeniewski, C. Langmuir 1999 , 15 , 786.<br />

10. Wang, H.; Wingender, C.; Baltruschat, H.; Lopez, M.; Reetz, M. T. J.<br />

Electroanal. Chem. 2001 , 509 , 163.<br />

11. Jusys, Z.; Behm, R. J. J. Phys. Chem. B 2001 , 105 , 10874.<br />

12. Jusys, Z.; Kaiser, J.; Behm, R. J. Langmuir 2003 , 19 , 6759.<br />

13. Wang, H.; Loffler, T.; Baltruschat, H. J. Appl. Electrochem. 2001 , 31 ,<br />

759.<br />

14. Korzeniewski, C.; Childers, C. L. J. Phys. Chem. B 1998 , 102 , 489.<br />

15. Lu, G.-Q.; Chrzanowski, W.; Wieckowski, A. J. Phys. Chem. B 2000 , 104<br />

, 5566.<br />

16. Corrigan, D. S.; Weaver, M. J. J. Electroanal. Chem. 1988 , 241 , 143.<br />

17. Watanabe, M.; Mo<strong>to</strong>o, S. J. Electroanal. Chem. 1975 , 60 , 267.<br />

70


18. Hamnett, A., In Interfacial Electrochemistry. Theory, Experiment, and<br />

Applications ; Wieckowski, A., Ed.; Marcel Dekker: New York, 1999, pp<br />

843.<br />

19. Jarvi, T. D.; Stuve, E. M., In Electrocatalysis ; Lipkowski, J. and Ross, P.<br />

N., Ed.; Wiley-VCH Publishers: New York, 1998; Vol. Ch. 3, pp 75.<br />

20. Gasteiger, H. A.; Markovic, N.; Ross, P. N., Jr.; Cairns, E. J. J.<br />

Electrochem. Soc. 1994 , 141 , 1795.<br />

21. Kardash, D.; Korzeniewski, C. Langmuir 2000 , 16 , 8419.<br />

22. Iwasita, T. Electrochim. Acta 2002 , 47 , 3663.<br />

23. Iwasita, T.; Hoster, H.; John-Anacker, A.; Lin, W. F.; Vielstich, W.<br />

Langmuir 2000 , 16 , 522.<br />

24. Batista, E. A.; Hoster, H.; Iwasita, T. J. Electroanal. Chem. 2003 , 554-555<br />

, 265.<br />

71


CHAPTER IV<br />

METHANOL ELECTROCHEMICAL CONVERSION TO FORMALDEHYDE<br />

OVER SUPPORTED CATALYST MATERIALS<br />

Introduction<br />

The DMFC is considered <strong>to</strong> have a great deal of potential due <strong>to</strong> its<br />

simplified system design and direct use of liquid fuel. But the present<br />

performance levels are not yet sufficient for commercial applications. The limited<br />

performance of the cell is due in part <strong>to</strong> poor kinetics of the anode reaction.<br />

Improvements in methanol oxidation catalysts are desired [1, 2]. Pt-based<br />

materials are typically employed. On pure Pt, methanol oxidation is strongly<br />

inhibited by poison formation. Bimetallic catalysts, such as PtRu, have much<br />

greater activity and resist poisoning [3, 4].<br />

In contrast <strong>to</strong> bulk metals, the electrodes in fuel cell are comprised of<br />

nanometer-scale metal particles adsorbed on<strong>to</strong> either micron-scale carbon<br />

particles or porous carbon supports [2]. The catalytic, physical, electrochemical<br />

and electronic properties of nanostructure catalyst materials can be quite<br />

different from those of bulk electrode materials [2, 5]. Reaction rates have been<br />

shown <strong>to</strong> be sensitive <strong>to</strong> the size and structure of the nanoparticle catalyst [6],<br />

and phase separation can occur in bi-metallic nanoparticles [7, 8]. In the study of<br />

methanol electrochemical conversion <strong>to</strong> H2CO <strong>over</strong> bulk electrodes (Chapter III),<br />

a significant amount of H2CO was produced. For comparison purposes, an<br />

attempt was made <strong>to</strong> study the methanol electrochemical oxidation reaction<br />

72


pathways <strong>over</strong> nanometer scale catalyst materials, which have more practical<br />

applications than bulk electrodes. In contrast <strong>to</strong> the solid metal electrodes<br />

studied, the yields of H2CO <strong>over</strong> catalyst particles were low (below 10%) during<br />

the electrochemical oxidation of 1.0 M methanol in 0.1M H2SO4 for short<br />

electrolysis periods (180 s). Materials examined include: Pt- black; C/Pt, 10 wt%;<br />

JM PtRu, XRu = 0.5; and SC PtRu catalyst in the following Ru compositions XRu =<br />

0.1, 0.25, and 0.5. In fuel cells, PtRu materials with XRu= 0.5 are favored for<br />

DMFC operations [2]. However, in nanometer scale PtRu catalysts, phase<br />

separation can occur <strong>to</strong> produce Ru-rich and Pt-rich regions [8, 9]. This structure<br />

contrasts bulk PtRu, which form random alloys. In addition, the high surface area<br />

particles create a porous bed that enables reacting molecules <strong>to</strong> undergo<br />

encounters with multiple catalyst sites and thereby facilitate more complete<br />

oxidation of reactant. This chapter presents the yield data for H2CO produced in<br />

reactions of methanol <strong>over</strong> nanometer-scale catalyst materials in the MACOR<br />

based small volume (50 µL) electrochemical cell described earlier (Chapter II).<br />

The oxidation of 1.0 M methanol at potentials in the range of 0.5-0.8 VRHE was<br />

examined.<br />

Experimental<br />

The details of the catalysts preparation and the electrochemical techniques<br />

employed were described in Chapter II.<br />

<strong>Formaldehyde</strong> Determination<br />

H2CO was determined fluorometrically following reaction with 1, 3-<br />

73


cyclohexadione (see Chapter II) using the flow injection arrangement shown in<br />

Figure 2.3.<br />

Results and Discussion<br />

Cyclic voltammograms of a polycrystalline gold (Au) electrode before<br />

(Figure 4.1) and after modification with an ultra-thin layer of Pt (Figure 4.2) or<br />

PtRu based catalyst materials (Figure 4.3) are displayed. The voltammograms<br />

were recorded in a conventional three electrode electrochemical cell. The cyclic<br />

voltammogram in Figure 4.1 for a mirror polished Au electrode in Ar purged 0.1<br />

M H2SO4 between H2 and O2 evolution potentials provides information about the<br />

cleanliness of surface. The shape and peak potentials for the waves that appear<br />

are in good agreement with prior reports [10]. The oxide formation peaks on the<br />

positive going sweep between 1.2 and 1.4 VRHE (associated with Au + H2O →<br />

AuOH + H + + e - ) and the oxide stripping peak (AuOH + H + + e - → Au + H2O) on<br />

the negative going sweep near 1.0 VRHE, <strong>to</strong>gether with the featureless double<br />

layer region (0.0 <strong>to</strong> 0.85 VRHE) represent the basic features of a clear,<br />

polycrystalline gold electrode. There are no obvious faradic electrochemical<br />

processes occurring in the wide potential region between 0.0-1.0 VRHE. This<br />

potential region provides a broad window <strong>over</strong> which the characteristics features,<br />

for instance, hydrogen adsorption and desorption peaks (0.0-0.3 VRHE) and<br />

double layer charging (0.3-0.7 VRHE) from adsorbed Pt and PtRu nanoparticles<br />

can be observed.<br />

Figure 4.2 shows the cyclic voltammogram of the Au electrode after<br />

adsorption of carbon supported platinum catalyst (C/Pt, 10 wt %) (Figure 4.2a) or<br />

74


a Pt-black film (Figure 4.2b) in 0.1 M H2SO4. The faradic current relative <strong>to</strong> the<br />

charging current in each case suggests the catalyst film c<strong>over</strong>age is a few<br />

monolayers [11, 12]. The cyclic voltammograms of the C/Pt, 10 wt% and Pt-<br />

black film (Figure 4.2 a and b) have waves characteristic of platinum similar <strong>to</strong><br />

Figure 3.2 in Chapter III. In acid electrolyte solutions, features associated with<br />

hydrogen adsorption and oxide formation and stripping are present, but are<br />

considerably weaker for C/Pt, 10 wt% than for bulk polycrystalline Pt. The current<br />

and peak potentials for C/ Pt, 10 wt% catalyst are consistent with prior reported<br />

voltammograms of the catalyst recorded on Au in 0.05 M H2SO4 [12, 13]. In<br />

contrast <strong>to</strong> the carbon supported Pt material, which are affected by the<br />

capacitance of the carbon, particles of Pt-black adsorbed on Au (Figure 4.2 b)<br />

display waves in the clear 0.1 M H2SO4 electrolyte solution that are more typical<br />

of bulk polycrystalline Pt. For Pt-black, the defined features in the hydrogen<br />

adsorption and oxide formation and strippinng regions are analogous <strong>to</strong> those<br />

observed for Pt-black catalyst immobilized Au or carbon [14,15, 16] and<br />

moderate <strong>to</strong> high metal loading (≥ 20%) carbon supported Pt catalysts on Au [11-<br />

13, 17]. The large oxide stripping peaks were found on the negative going<br />

sweep near 0.7 VRHE and 0.65 VRHE for Pt-black and C/Pt ,10 wt% respectively.<br />

For both cases the hydrogen adsorption/desorption region in between 0.0 <strong>to</strong> 0.3<br />

VRHE indicate the presence of Pt on the gold surface. On the positive scan, the<br />

current rise near 0.8 VRHE and the Pt-oxide formation features are observed in<br />

the range of 0.9 <strong>to</strong> 1.5 VRHE. The clear definitions of all the waves indicate the Pt<br />

particles have a high degree of cleanliness. The arrow in both cyclic<br />

75


Current<br />

Polycrystalline Au electrode<br />

in 0.1 M H SO<br />

2 4<br />

0<br />

0.02 mA / cm 2<br />

0.0 0.4 0.8 1.2 1.6 2.0<br />

E / V vs. RHE<br />

Figure 4.1: Cyclic voltammogram of a polycrystalline gold electrode in 0.1 M<br />

H2SO4; Scan rate: 50 mV/s.<br />

76


Current<br />

Au / catalyst film<br />

in 0.1 M H2SO4<br />

200 µA<br />

0<br />

0<br />

-0.1 0.5 1.1 1.7<br />

E / V vs. RHE<br />

77<br />

(a) C/Pt, 10wt%<br />

(b) Pt- black<br />

Figure 4.2: Cyclic voltammogram of a polycrystalline gold electrode in 0.1 M<br />

H2SO4 after adsorption of C/Pt, 10 wt% (a) or Pt-black (b) catalyst films; Scan<br />

Rate: 50 mV/s. Arrow represents the stripping wave for gold oxide.


voltammograms represents the stripping wave for gold oxide, which indicates<br />

that gold is exposed at the surface between Pt particles.<br />

Figure 4.3 shows the representative cyclic voltammograms of a Au<br />

electrode after modification with an ultra thin layer of nanometer scale PtRu<br />

catalyst materials (JM PtRu, XRu = 0.5; SC PtRu, XRu = 0.1 and SC PtRu, XRu =<br />

0.5) used in this study. The response of the catalyst particles can be correlated<br />

<strong>to</strong> the Ru content in the particles by comparison <strong>to</strong> results for bulk PtRu alloy<br />

electrodes (Chapter III). For the sonochemically prepared nanoparticles with XRu<br />

= 0.1 in 0.1 M H2SO4 (Figure 4.3b, central voltammogram), waves appear at<br />

0.05-0.3 VRHE characteristic of hydrogen adsorption and hydrogen desorption [11,<br />

12]. The features for the SC PtRu, XRu = 0.1 nanocatalysts are almost similar <strong>to</strong><br />

those of Pt electrode (Figure 3.2) except some lessening of wave current in the<br />

hydrogen region reflecting the presence of Ru. Like the bulk PtRu alloy electrode<br />

with XRu = 0.07 studied earlier [18, 19], the SC PtRu, XRu = 0.1 electrode also<br />

displayed a much narrower double layer charging region compared <strong>to</strong> the Ru rich<br />

electrode. Our PtRu bulk electrode (Figure 3.1) and catalyst (Figure 4.3 b) with<br />

similar Ru content (XRu = 0.1) display the same properties, which is encouraging.<br />

The upper and bot<strong>to</strong>m voltammograms in Figure 4.3 represent the commercial<br />

JM PtRu, XRu = 0.5 (Figure 4.3a) and SC PtRu, XRu = 0.5 (Figure 4.3c)<br />

nanoparticle materials, where the multiple hydrogen adsorption and desorption<br />

peaks are completely suppressed. In these cases, incorporating more Ru (XRu =<br />

0.5) in<strong>to</strong> the particles, increases the Ru surface composition and the hydrogen<br />

78


Current<br />

Au / catalyst film<br />

in 0.1 M H2SO4<br />

0<br />

0<br />

0<br />

200 µA<br />

(a) JM PtRu, XRu = 0.5<br />

(b) SC PtRu, XRu = 0.1<br />

(c) SC PtRu, XRu = 0.5<br />

0.0 0.2 0.4 0.6 0.8<br />

E / V vs. RHE<br />

Figure 4.3. Cyclic voltammograms of catalyst materials employed in the study.<br />

Scans were recorded in Ar purged 0.1 M H2SO4 at 50 mV/s. The following<br />

nanoscale catalyst materials were studied as thin films adsorbed on a Au<br />

electrode- Johnson Matthey (JM) PtRu with XRu = 0.5 (a); PtRu catalyst<br />

prepared via a sonochemical (SC) method with XRu = 0.1 (b) and 0.5 (c).<br />

79<br />

X 2<br />

X 4


adsorption and desorption features become less resolved and <strong>over</strong>lap<br />

increasingly with the double layer region. The broad envelopes typical for high<br />

Ru content materials are noticeable in the range of 0.3-0.7 VRHE, the double layer<br />

charging region. This section is featureless typical of the pseudocapacitive<br />

properties of Ru oxide [9]. The same findings have been reported by other<br />

researchers where the formation of surface oxide occurs at progressively lower<br />

potentials as the Ru c<strong>over</strong>age on a pure Pt catalyst was increased [18, 20-22].<br />

The H2CO produced with these and a few additional catalysts on a gold<br />

electrode was determined as a function of potential (0.5-0.8 VRHE). Table 4.1 and<br />

4.2 report reaction charge and the quantity of H2CO produced during short<br />

periods (180s) of methanol electrochemical oxidation <strong>over</strong> the nanoscale<br />

catalysts. Figure 4.4 <strong>to</strong> 4.11 graphically represent the H2CO yields with respect<br />

<strong>to</strong> electrode potentials for all the catalysts studied. The percent conversion was<br />

computed from the ratio of the charge required <strong>to</strong> produce the quantity of H2CO<br />

detected fluorometrically <strong>to</strong> the <strong>to</strong>tal charge passed during electrolysis. In Table<br />

4.1, the results for Pt based catalysts (Pt-black and C/Pt, 10 wt %) are<br />

summarized. For Pt-black the greatest yield (5.28%) was found at 0.6 VRHE and<br />

decreased at higher potentials. The largest yield for C/Pt, 10 wt% catalyst was<br />

7.9% for 0.5 VRHE and decreased continuously <strong>to</strong> 0.5 % with increasing<br />

potentials. High reaction charges were observed for these catalysts at almost all<br />

potentials compared <strong>to</strong> a solid platinum electrode (Chapter III, Table 3.1). The<br />

response reflects both the larger surface area of the catalyst and the strong<br />

poisoning that occurred for the solid platinum at the potentials employed. Similar<br />

80


Table 4.1 Summary of charge and formaldehyde yields from the oxidation of 1.0<br />

M CH3OH <strong>over</strong> Pt based nanoscale catalyst materials a .<br />

Electrode Materials b Potential c<br />

(E/V vs.<br />

RHE)<br />

Average<br />

Charge<br />

(mC)<br />

81<br />

Average<br />

H2CO<br />

Quantity<br />

(nmol)<br />

H2CO<br />

Yield %<br />

Pt- Black 0.8 121.21 13.00 2.0 ± 0.5<br />

0.7 64.30 10.68 3.3 ± 0.5<br />

0.6 35.08 9.76 5.3 ± 0.7<br />

0.5 17.01 3.70 4.8 ± 0.5<br />

C/Pt, 10 wt% 0.8 121.21 3.39 0.5 ± 0.1<br />

0.7 65.15 7.57 2.2 ± 0.3<br />

0.6 33.76 3.95 2.3 ± 0.2<br />

0.5 14.78 5.70 7.9 ± 0.6<br />

a All experiments involved the electrolysis of 1.0 M CH3OH in 0.1 M H2SO4 in a volume of 50 µL<br />

for a period of 180s. The values reported for the charge and H2CO amount are averages<br />

calculated from results of 3-5 trials. The yield values were computed from the ratio of the charge<br />

required <strong>to</strong> produce H2CO <strong>to</strong> the <strong>to</strong>tal charge passed during the electrolysis period. These yield<br />

values were averaged <strong>to</strong> give the reported percentages and uncertainties. The uncertainties<br />

represent 95.5% confidence limits based on two times the sample standard deviation. b The<br />

following catalyst materials were used on a gold electrode: Pt-Black (from JM); C/Pt, 10 wt% =<br />

Pt at 10 wt% on Vulcan XC-72R carbon (from JM). c The reference electrode was KCl saturated<br />

Ag/AgCl. For reporting purposes, all the measured potentials vs. Ag/AgCl were converted <strong>to</strong> the<br />

reversible hydrogen electrode (RHE) scale and expressed as volts versus RHE, (VRHE).


% H 2CO Yield<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

4.762<br />

82<br />

5.281<br />

3.251<br />

1.973<br />

0.4 0.5 0.6 0.7 0.8 0.9<br />

E/V vs. RHE<br />

Figure 4.4: H2CO yields from methanol electrochemical oxidation in 0.1 M H2SO4<br />

on Pt -Black catalyst on a polycrystalline Au electrode. All experiments involved<br />

the electrolysis of 1.0 M CH3OH in 0.1 M H2SO4 in a volume of 50 µL for a period<br />

of 180s.<br />

% H 2CO Yield<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

7.85<br />

2.29<br />

2.24<br />

0.54<br />

0.4 0.5 0.6 0.7 0.8 0.9<br />

E/V vs RHE<br />

Figure 4.5: H2CO yields from methanol electrochemical oxidation in 0.1 M H2SO4<br />

on C/Pt, 10 wt% catalyst on a polycrystalline Au electrode. All experiments<br />

involved the electrolysis of 1.0 M CH3OH in 0.1 M H2SO4 in a volume of 50 µL for<br />

a period of 180s.


esults have been observed with lower methanol concentration and different<br />

catalyst materials [23, 24]. With the higher methanol concentrations employed in<br />

this study, the yield of partial oxidation products is expected <strong>to</strong> increase as the<br />

decrease in interfacial water activity lowers the rate of surface oxide formation<br />

(H2O → OHads + H + + e - ) [25-27].<br />

To enable comparisons among the catalyst systems, JM PtRu, XRu = 0.5,<br />

and sonochemically prepared PtRu catalysts with various Ru surface<br />

compositions (XRu = 0.1, 0.25 and 0.5), were chosen. Similar <strong>to</strong> PtRu bulk<br />

electrodes, PtRu catalysts have considerably higher <strong>to</strong>lerance for carbon<br />

monoxide than pure Pt materials [18]. The PtRu catalysts are predicted <strong>to</strong> have<br />

a higher catalytic activity for methanol oxidation than platinum due <strong>to</strong> the<br />

bifunctional mechanism (See Chapter III) [3, 4, 18]. In addition, incorporating Ru<br />

in<strong>to</strong> the Pt lattice is believed <strong>to</strong> weaken the Pt-COads bond strength, reducing the<br />

–COads c<strong>over</strong>age [28]. Compared <strong>to</strong> Pt catalyst, the PtRu materials showed<br />

superior activity <strong>to</strong>ward CH3OH oxidation, and the average yields for H2CO are<br />

lower than for the reactions on PtRu materials particularly at 0.5 VRHE, except for<br />

SC PtRu, XRu = 0.1 (Table 4.2 and Figure 4.7- 4.11). Similar <strong>to</strong> the bulk<br />

materials, the H2CO yield for the SC PtRu, XRu = 0.1 catalyst more closely<br />

resembles that for clean Pt, as they both produce a larger quantity of H2CO (yield<br />

~7.8%) at the lowest working potentials (0.5VRHE). Although all catalysts in our<br />

analysis generate very small yields, and there is no sharp difference in yield<br />

values for all catalysts across the studied potential range (0.5-0.8 VRHE), it is<br />

interesting <strong>to</strong> note that at 0.5 VRHE the H2CO yield declined from 7.4% <strong>to</strong> 0.5%<br />

83


with increasing Ru content (Figure 4.11). PtRu catalysts with XRu = 0.5 showed<br />

the lowest H2CO yields, and JM PtRu, XRu = 0.5 and SC PtRu, XRu = 0.5 results<br />

are comparable <strong>to</strong> each other. At the lowest and the most important electrode<br />

potential (0.5 VRHE) from the standpoint of practical fuel cell applications, the<br />

H2CO yield did not decrease significantly with the incorporation of a small amount<br />

of Ru (~7.4% yield for XRu = 0.1) compared <strong>to</strong> Pt based catalysts (~8% yield for<br />

C/Pt, 10 wt%). However, the H2CO yield decreased with further increases in Ru<br />

<strong>to</strong> 1.4% for SC PtRu, XRu = 0.25 and 0.5-0.7% for both SC PtRu, XRu = 0.5 and<br />

JM PtRu, XRu = 0.5. Similar trends were observed in this project for bulk<br />

electrodes (Chapter III) and for nanometer scale catalyst by other researchers<br />

[20, 29]. DEMS results for PtRu catalysts showed that the alloy materials<br />

containing nearly equal Pt and Ru content (XRu = 0.3 <strong>to</strong> 0.5) exhibited higher<br />

activity <strong>to</strong>ward CO2 production from methanol at ~ 0.4- 0.5 VRHE compared <strong>to</strong><br />

lower Ru containing catalysts [20]. The observed result demonstrated that the<br />

surface area normalized methanol oxidation activity of PtRu catalysts and<br />

electrodes is highest for low Ru contents around XRu = 0.15 at more positive<br />

potentials (~ 0.7 VRHE) [20]. These activities refer <strong>to</strong> the inherent chemical<br />

activity obtained by normalizing the oxidation current <strong>to</strong> the active surface area<br />

determined by COad- stripping <strong>to</strong> CO2 [20]. The DEMS study with smooth Pt and<br />

PtRu (XRu = 0.35 - 0.45) materials by Wang and co-workers showed that the<br />

current efficiency of CO2 increased with the incorporation of Ru as it promotes<br />

the methanol oxidation on Pt at low potentials [29]. At room temperature, PtRu<br />

alloy catalysts having a Ru composition XRu= 0.3-0.5, at present the best<br />

84


Table 4.2: Summary of charge and formaldehyde yields from the oxidation of 1.0<br />

M CH3OH <strong>over</strong> PtRu based nanoscale catalyst materials a .<br />

Electrode Materials b Potential c<br />

Average H2CO H2CO Catalyst<br />

(E/V vs. Charge Quantity Yield % loading<br />

RHE) (mC) (nmol)<br />

JM PtRu, XRu = 0.5 0.8 28.28 3.00 2.2 ± 0.6 62µg/cm 2<br />

0.7 24.06 2.99 1.6 ± 0.5<br />

0.6 8.81 0.39 1.1 ± 0.7<br />

0.5 5.58 0.16 0.5 ± 0.1<br />

SC PtRu, XRu = 0.5 0.8 27.47 0.59 0.7 ± 0.5 86µg/cm 2<br />

0.7 18.87 0.51 2.0 ± 0.7<br />

0.6 15.83 0.45 2.1 ± 0.6<br />

0.5 14.92 0.22 0.7 ± 0.6<br />

SC PtRu, XRu = 0.25 0.8 43.39 9.92 4.4 ± 0.2 86µg/cm 2<br />

0.7 34.92 7.29 2.8 ± 0.3<br />

0.6 8.17 2.34 5.5 ± 0.1<br />

0.5 7.28 0.47 1.4 ± 0.5<br />

SC PtRu, XRu = 0.1 0.8 54.73 13.04 4.6 ± 0.4 100µg/cm 2<br />

SC PtRu, XRu = 0.1<br />

with Nafion<br />

0.7 37.00 6.81 3.5 ± 0.4<br />

0.6 19.04 4.78 5.2 ± 0.6<br />

0.5 7.43 2.79 7.4 ± 0.7<br />

0.8 153.17 5.25 0.7 ± 0.3 96 µg/cm 2<br />

0.7 118.50 6.34 1.1 ± 0.5<br />

0.6 84.57 4.40 1.0 ± 0.1<br />

0.5 18.72 1.76 2.0 ± 0.5<br />

b Johnson Matthey (JM) PtRu with Ru mole fraction (XRu) 0.5; PtRu catalyst prepared via a sonochemical<br />

(SC) method with XRu = 0.1, 0.25 and 0.5. SC PtRu, XRu = 0.1 with Nafion refers <strong>to</strong> the catalyst film capped<br />

by a Nafion coating. a,c Other conditions are the same with table 4.1<br />

85


% H 2CO yield<br />

10<br />

8<br />

6<br />

4<br />

2<br />

7.38<br />

5.19<br />

86<br />

3.53<br />

4.61<br />

0<br />

0.4 0.5 0.6 0.7 0.8 0.9<br />

E/V vs RHE<br />

Figure 4.6: H2CO yields from methanol electrochemical oxidation in 0.1 M H2SO4<br />

on SC PtRu, XRu= 0.1 catalyst on a polycrystalline Au electrode. All experiments<br />

involved the electrolysis of 1.0 M CH3OH in 0.1 M H2SO4 in a volume of 50 µL for<br />

a period of 180s.<br />

% H 2CO yield<br />

4<br />

3<br />

2<br />

1<br />

1.95<br />

1.004<br />

1.11<br />

0.71<br />

0<br />

0.4 0.5 0.6 0.7 0.8 0.9<br />

E/V vs RHE<br />

Figure 4.7: H2CO yields from methanol electrochemical oxidation in 0.1 M H2SO4<br />

on SC PtRu, XRu= 0.1 catalyst, capped by Nafion on a polycrystalline Au<br />

electrode. All experiments involved the electrolysis of 1.0 M CH3OH in 0.1 M<br />

H2SO4 in a volume of 50 µL for a period of 180s.


% H 2CO yield<br />

6<br />

4<br />

2<br />

0<br />

1.36<br />

5.52<br />

87<br />

2.8<br />

4.43<br />

0.4 0.5 0.6 0.7 0.8 0.9<br />

E/V vs RHE<br />

Figure 4.8: H2CO yields from methanol electrochemical oxidation in 0.1 M H2SO4<br />

on SC PtRu, XRu= 0.25 catalyst on a polycrystalline Au electrode. All experiments<br />

involved the electrolysis of 1.0 M CH3OH in 0.1 M H2SO4 in a volume of 50 µL for<br />

a period of 180s.<br />

% H 2CO yield<br />

5<br />

3<br />

1<br />

-1<br />

0.66<br />

2.09<br />

1.96<br />

0.67<br />

0.4 0.5 0.6 0.7 0.8 0.9<br />

E/V vs RHE<br />

Figure 4.9: H2CO yields from methanol electrochemical oxidation in 0.1 M H2SO4<br />

on SC PtRu, XRu= 0.5 catalyst on a polycrystalline Au electrode. All experiments<br />

involved the electrolysis of 1.0 M CH3OH in 0.1 M H2SO4 in a volume of 50 µL for<br />

a period of 180s.


% H 2CO yield<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

0.53<br />

1.11<br />

88<br />

1.57<br />

2.15<br />

0.4 0.5 0.6 0.7 0.8 0.9<br />

E/V vs RHE<br />

Figure 4.10: H2CO yields from methanol electrochemical oxidation in 0.1 M<br />

H2SO4 on JM PtRu, XRu= 0.5 catalyst on a polycrystalline Au electrode. All<br />

experiments involved the electrolysis of 1.0 M CH3OH in 0.1 M H2SO4 in a<br />

volume of 50 µL for a period of 180s.


% H 2CO yield<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

Pt-black C/Pt, 10 wt% SC PtRu,<br />

XRu=0.1<br />

1 M <strong>Methanol</strong>; 0.5 V vs. RHE<br />

SC PtRu,<br />

XRu=0.1, with<br />

Nafion<br />

89<br />

SC PtRu,<br />

XRu=0.25<br />

Electrode Composition<br />

JM PtRu,<br />

XRu=0.5<br />

SC PtRu,<br />

XRu=0.5<br />

Figure 4.11: H2CO yield from methanol (1.0 M) electrochemical oxidation on<br />

catalyst materials at 0.5 VRHE.


catalysis for CO2 production from methanol, probably due <strong>to</strong> the high surface<br />

c<strong>over</strong>age of oxides on Ru. The bot<strong>to</strong>m entry of Table 5.2 indicates that the<br />

methanol oxidation occurs more efficiently when catalyst (SC PtRu, XRu = 0.1)<br />

was immobilized in Nafion. For every reaction potentials, the H2CO yield for SC<br />

PtRu, XRu = 0.1, with Nafion, became much less than that of SC PtRu, XRu = 0.1,<br />

without Nafion. At the most highlighted investigated potential (0.5 VRHE), the<br />

H2CO yield and H2CO quantity were reduced <strong>to</strong> a greater extent as the catalyst<br />

layer and the products and byproducts produced, are entrapped by Nafion film.<br />

The reaction charges were also increased a great deal compared <strong>to</strong> the catalyst<br />

without a Nafion film c<strong>over</strong>ing.<br />

In the platinum based catalysts (Table 4.1), the amount of H2CO produced<br />

at 0.5 VRHE are 3.70 nmol (Pt-black) and 5.70 nmol (C/Pt, 10 wt %) which were<br />

somewhat large relative <strong>to</strong> the amount (~ 0.1-0.4 nmol) in the PtRu based<br />

catalysts with average or higher Ru composition (XRu = 0.25 -0.5). This<br />

difference in H2CO quantity formed probably arises from the Ru incorporation<br />

and availability of oxides <strong>to</strong> transform H2CO <strong>to</strong> higher oxidation products. The<br />

more ruthenium is alsol catalyzing the CO oxidation, thus favoring the CO2<br />

formation rather than forming much H2CO [3, 4, 18]. A more useful value <strong>to</strong><br />

analyze is the H2CO percent yield because the quantity is corrected for surface<br />

c<strong>over</strong>age effects. With both the Pt-black and the Pt-Ru deposits, the percent<br />

yield is very low which is below 5% for 0.5 VRHE. Since the percent yield values<br />

are well below the solid Pt and PtRu electrodes (Chapter III), these experiments<br />

indicate that the small nanoparticles grouping of Pt or PtRu catalysts better drive<br />

90


the reaction beyond H2CO. In the nano-scale catalyst films partial oxidation<br />

product like H2CO generally desorbs from catalyst surfaces and diffuses within<br />

the film where it can meet additional catalyst for more complete oxidation of<br />

methanol [24].<br />

The analysis of data and the results for repetitive measurements are<br />

summarized graphically in Figure 4.12 <strong>to</strong> Figure 4.25 for all the studied catalysts.<br />

Figure 4.12 <strong>to</strong> Figure 4.18 report H2CO formation rate as a function of potential.<br />

Figure 4.19 <strong>to</strong> Figure 4.25 shows the plot of nanomoles vs. the reaction charge<br />

passed following the sample electrolysis. The error bars represent 95%<br />

confidence intervals on the mean of 3 or 4 repetitive measurements.<br />

Conclusions<br />

Pt and PtRu catalyst materials were supported as a thin layer on a Au<br />

electrode and characterized by cyclic voltammetry in 0.1 M H2SO4 and for activity<br />

<strong>to</strong>ward H2CO production from 1.0 M CH3OH. The studies showed good precision<br />

and reproducibility. All catalysts produced high electrolysis currents, but low (<<br />

10%) yields of H2CO, indicating CH3OH and any intermediate H2CO produced is<br />

converted <strong>to</strong> more complete oxidation products <strong>over</strong> the catalyst bed. In this<br />

respect, the catalyst layers are much more reactive <strong>to</strong>ward CH3OH than bulk<br />

electrodes of similar composition. Among the materials studied, the PtRu<br />

catalysts with XRu = 0.5 showed the lowest yield for H2CO. It appears that both<br />

the physical structure of catalyst materials (bulk versus nanometer scale<br />

particles) and composition influence CH3OH oxidation pathways. In this work,<br />

91


the responses observed for bulk and nanometer scale materials are compared<br />

and discussed further in Chapter VI.<br />

92


Rate (nmol/s/cm 2 )<br />

0.045<br />

0.03<br />

0.015<br />

0<br />

0.4 0.5 0.6 0.7 0.8 0.9<br />

E/V vs. RHE<br />

Figure 4.12: Plot of H2CO formation rate averaged <strong>over</strong> 180 s electrolysis periods<br />

vs. potential for the oxidation of 1.0 M methanol in 0.1 M H2SO4 <strong>over</strong> a Pt -Black<br />

catalyst layer adsorbed on a polycrystalline Au electrode.<br />

Rate (nmol/s/cm 2 )<br />

0.03<br />

0.02<br />

0.01<br />

0<br />

0.4 0.5 0.6 0.7 0.8 0.9<br />

E/V vs RHE<br />

Figure 4.13: Plot of H2CO formation rate averaged <strong>over</strong> 180 s electrolysis periods<br />

vs. potential for the oxidation of 1.0 M methanol in 0.1 M H2SO4 <strong>over</strong> a C/Pt, 10<br />

wt% catalyst layer adsorbed on a polycrystalline Au electrode.<br />

93


Rate (nmol/s/cm 2 )<br />

0.035<br />

0.03<br />

0.025<br />

0.02<br />

0.015<br />

0.01<br />

0.005<br />

0<br />

0.45 0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85<br />

E/V vs. RHE<br />

Figure 4.14: Plot of H2CO formation rate averaged <strong>over</strong> 180 s electrolysis periods<br />

vs. potential for the oxidation of 1.0 M methanol in 0.1 M H2SO4 <strong>over</strong> a SC PtRu,<br />

XRu= 0.1 catalyst layer adsorbed on a polycrystalline Au electrode.<br />

Rate (nmol/s/cm 2 )<br />

0.01<br />

0.008<br />

0.006<br />

0.004<br />

0.002<br />

0<br />

0.4 0.45 0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85<br />

E/V vs. RHE<br />

Figure 4.15: Plot of H2CO formation rate averaged <strong>over</strong> 180 s electrolysis periods<br />

vs. potential for the oxidation of 1.0 M methanol in 0.1 M H2SO4 <strong>over</strong> a SC PtRu,<br />

XRu= 0.1 catalyst layer (capped by Nafion) adsorbed on a polycrystalline Au<br />

electrode.<br />

94


Rate (nmol/s/cm 2 )<br />

0.025<br />

0.02<br />

0.015<br />

0.01<br />

0.005<br />

0<br />

0.45 0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85<br />

E/V vs. RHE<br />

Figure 4.16: Plot of H2CO formation rate averaged <strong>over</strong> 180 s electrolysis periods<br />

vs. potential for the oxidation of 1.0 M methanol in 0.1 M H2SO4 <strong>over</strong> a SC PtRu,<br />

XRu= 0.25 catalyst layer adsorbed on a polycrystalline Au electrode.<br />

Rate (nmol/s/cm 2 )<br />

0.0007<br />

0.0006<br />

0.0005<br />

0.0004<br />

0.0003<br />

0.0002<br />

0.0001<br />

0<br />

0.45 0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85<br />

E/V vs. RHE<br />

Figure 4.17: Plot of H2CO formation rate averaged <strong>over</strong> 180 s electrolysis periods<br />

vs. potential for the oxidation of 1.0 M methanol in 0.1 M H2SO4 <strong>over</strong> a SC PtRu,<br />

XRu= 0.50 catalyst layer adsorbed on a polycrystalline Au electrode.<br />

95


Rate (nmol/s/cm 2 )<br />

0.007<br />

0.006<br />

0.005<br />

0.004<br />

0.003<br />

0.002<br />

0.001<br />

0<br />

0.45 0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85 0.9<br />

E/V vs. RHE<br />

Figure 4.18: Plot of H2CO formation rate averaged <strong>over</strong> 180 s electrolysis periods<br />

vs. potential for the oxidation of 1.0 M methanol in 0.1 M H2SO4 <strong>over</strong> a JM PtRu,<br />

XRu= 0.50 catalyst layer adsorbed on a polycrystalline Au electrode.<br />

<strong>Formaldehyde</strong> Quantity (nmol)<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0.6 V RHE<br />

0.5 V RHE<br />

0.7 V RHE<br />

0<br />

0 20 40 60 80 100 120 140<br />

Charge (mC)<br />

96<br />

0.8V RHE<br />

Figure 4.19: Plot of the nanomoles H2CO formed vs. the reaction charge passed<br />

following the oxidation of 1.0 M methanol in 0.1 M H2SO4 <strong>over</strong> a Pt-Black catalyst<br />

layer adsorbed on a polycrystalline Au electrode (Reaction time = 180 s).


<strong>Formaldehyde</strong> Quantity (nmol)<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0.5 V RHE<br />

0.6 V RHE<br />

0.7 V RHE<br />

0 20 40 60 80 100 120 140<br />

Charge (mC)<br />

97<br />

0.8 V RHE<br />

Figure 4.20: Plot of the nanomoles H2CO formed vs. the reaction charge passed<br />

following the oxidation of 1.0 M methanol in 0.1 M H2SO4 <strong>over</strong> a C/Pt, 10 wt%<br />

catalyst layer adsorbed on a polycrystalline Au electrode (Reaction time = 180 s).<br />

<strong>Formaldehyde</strong> Quantity (nmol)<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0.5 V RHE<br />

0.6 V RHE<br />

0.7 V RHE<br />

0 10 20 30 40 50 60 70<br />

Charge (mC)<br />

Figure 4.21: Plot of the nanomoles H2CO formed vs. the reaction charge passed<br />

following the oxidation of 1.0 M methanol in 0.1 M H2SO4 <strong>over</strong> a SC PtRu, XRu=<br />

0.10 catalyst layer adsorbed on a polycrystalline Au electrode (Reaction time =<br />

180 s).<br />

0.8 V RHE


<strong>Formaldehyde</strong> Quantity (nmol)<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0.5 V RHE<br />

0<br />

0 20 40 60 80 100 120 140 160 180<br />

Charge (mC)<br />

98<br />

0.7 V RHE<br />

0.6 V RHE<br />

0.8 V RHE<br />

Figure 4.22: Plot of the nanomoles H2CO formed vs. the reaction charge passed<br />

following the oxidation of 1.0 M methanol in 0.1 M H2SO4 <strong>over</strong> a SC PtRu, XRu=<br />

0.10 catalyst layer (capped by Nafion) adsorbed on a polycrystalline Au electrode<br />

(Reaction time = 180 s).<br />

<strong>Formaldehyde</strong> Quantity (nmol)<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0.5 V RHE<br />

0.6 V RHE<br />

0 5 10 15 20 25 30 35 40 45 50<br />

Charge (mC)<br />

Figure 4.23: Plot of the nanomoles H2CO formed vs. the reaction charge passed<br />

following the oxidation of 1.0 M methanol in 0.1 M H2SO4 <strong>over</strong> a SC PtRu, XRu=<br />

0.25 catalyst layer adsorbed on a polycrystalline Au electrode (Reaction time =<br />

180 s).<br />

0.7 V RHE<br />

0.8 V RHE


<strong>Formaldehyde</strong> Quantity (nmol)<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.5 V RHE<br />

0.6 V RHE<br />

0.7 V RHE<br />

0<br />

12 14 16 18 20 22 24 26 28<br />

Charge (mC)<br />

99<br />

0.8 V RHE<br />

Figure 4.24: Plot of the nanomoles H2CO formed vs. the reaction charge passed<br />

following the oxidation of 1.0 M methanol in 0.1 M H2SO4 <strong>over</strong> a SC PtRu, XRu=<br />

0.50 catalyst layer adsorbed on a polycrystalline Au electrode (Reaction time =<br />

180 s).<br />

<strong>Formaldehyde</strong> Quantity (nmol)<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0.5 V RHE<br />

0.6 V RHE<br />

0<br />

0 5 10 15 20 25 30 35<br />

Charge (mC)<br />

0.7 V RHE<br />

0.8 V RHE<br />

Figure 4.25: Plot of the nanomoles H2CO formed vs. the reaction charge passed<br />

following the oxidation of 1.0 M methanol in 0.1 M H2SO4 <strong>over</strong> a JM PtRu, XRu=<br />

0.10 catalyst layer adsorbed on a polycrystalline Au electrode (Reaction time =<br />

180 s).


References<br />

1. Hogarth, M. P.; Hards, G. A. Platinum Metals Rev. 1996, 40, 150.<br />

2. Hogarth, M. P.; Ralph, T. R. Platinum Metals Rev. 2002, 46, 146.<br />

3. Watanabe, M.; Mo<strong>to</strong>o, S. J. Electroanal. Chem. 1975, 60, 267.<br />

4. Hamnett, A., In Interfacial Electrochemistry. Theory, Experiment, and<br />

Applications; Wieckowski, A., Ed.; Marcel Dekker: New York, 1999, pp<br />

843.<br />

5. El-Sayed, M. A. Accounts of Chemical Research 2001, 34, 257.<br />

6. Mayrhofer, K. J. J.; Blizanac, B. B.; Arenz, M.; Stamenkovic, V. R.; Ross,<br />

P. N.; Markovic, N. M. J. Phys. Chem. B 2005 ,109 , 14433.<br />

7. Stroud, R. M.; Long, J. W.; Swider-Lyons, K. E.; Rolison, D. R. Microsc.<br />

Microanal. 2002, 8, 50.<br />

8. Diaz-Morales, R. R.; Liu, R.; Fachini, E.; Chen, G.; Segre, U. C.; Martinez,<br />

A.; Cabera, C.; Smotkin, E. S. J. Elec<strong>to</strong>chem. Soc. 2004, 151, A1314.<br />

9. Rolison, D. R.; Hagans, P. L.; Swider, K. E.; Long, J. W. Langmuir 1999 ,<br />

15 , 774.<br />

10. Angnes, L.; Richter, E. M.; Augelli, M. A.; Kume, G. H. Anal. Chem. 2000,<br />

2000, 5503.<br />

11. Park, S.; Xie, Y.; Weaver, M. J. Langmuir 2002, 18, 5792.<br />

12. Park, S.; Tong, Y. Y.; Wieckowski, A.; Weaver, M. J. Electrochemistry<br />

Communications 2001, 3, 509.<br />

13. Park, S.; Tong, Y. Y.; Wieckowski, A.; Weaver, M. J. Langmuir 2002 , 18 ,<br />

3233.<br />

14. Day, J. B.; Vuissoz, P.-A.; Oldfield, E.; Wieckowski, A.; Ansermet, J.-P. J.<br />

Am. Chem. Soc. 1996, 118, 13046.<br />

15. Long, J. W.; Rolison, D. R., In New Directions in Electroanalytical<br />

Chemistry II; Leddy, J., Vanysek, P. and Porter, M. D., Ed.;<br />

<strong>Electrochemical</strong> Society: Penning<strong>to</strong>n, NJ, 1999; Vol. PV99-5, pp 125.<br />

16. Long, J. W.; Ayers, K. E.; Rolison, D. R. J. Electroanal. Chem. 2002, 522,<br />

58.<br />

100


17. Park, S.; Wasileski, S. A.; Weaver, M. J. J. Phys. Chem. B 2001 , 105 ,<br />

9719.<br />

18. Gasteiger, H. A.; Markovic, N.; Ross, P. N., Jr.; Cairns, E. J. J. Phys.<br />

Chem. 1993, 97, 12020.<br />

19. Gasteiger, H. A.; Markovic, N.; Ross, P. N., Jr.; Cairns, E. J. J. Phys.<br />

Chem. 1994, 98,617.<br />

20. Jusys, Z.; Kaiser, J.; Behm, R. J. Electrochim. Acta 2002, 47, 3693.<br />

21. Long, J. W.; Swider, K. E.; Merzbacher, C. I.; Rolison, D. R. Langmuir<br />

1999 , 15 , 780.<br />

22. Lee, C. E.; Bergens, S. H. J. Phys. Chem. B 1998, 102, 193.<br />

23. Korzeniewski, C.; Childers, C. L. J. Phys. Chem. B 1998, 102, 489.<br />

24. Childers, C. L.; Huang, H.; Korzeniewski, C. Langmuir 1999 , 15 , 786.<br />

25. Gao, P.; Chang, I. C.; Zhou, Z.; Weaver, M. J. J. Electroanal. Chem.<br />

1989, 272,161.<br />

26. Wasmus, S.; Wang, J.-T.; Savinell, R. F. J. Electrochem. Soc. 1995,142 ,<br />

3825.<br />

27. Lin, W.-F.; Wang, J.-T.; Savinell, R. F. J. Electrochem. Soc. 1997,144 ,<br />

1917.<br />

28. Ralph, T. R.; Hogarth, M. P. Platinum Metals Rev. 2002, 46, 117.<br />

29. Wang, H.; Wingender, C.; Baltruschat, H.; Lopez, M.; Reetz, M. T. J.<br />

Electroanal. Chem. 2001, 509, 163.<br />

30. Jarvi, T. D.; Stuve, E. M., In Electrocatalysis ; Lipkowski, J. and Ross, P.<br />

N., Ed.; Wiley-VCH Publishers: New York, 1998; Vol. Ch. 3, pp 75.<br />

101


Chapter V<br />

A MICRO-VOLUME ELECTROCHEMICAL CELL FOR THE STUDY OF<br />

METHANOL OXIDATION PATHWAYS<br />

Introduction<br />

A definition of thin layer electrochemistry is that area of electrochemical<br />

endeavor in which special advantage is taken of restricting the diffusion field of<br />

electroactive species and products [1]. It is an approach for achieving bulk<br />

electrolysis conditions and a large electrode surface area <strong>to</strong> solution volume ratio<br />

(A/V) without convective mass transfer and involves decreasing the solution<br />

volume (V) such that a very small amount is confined <strong>to</strong> a thin layer adjacent <strong>to</strong><br />

the working electrode surface. The solution thickness (or cell height) should be<br />

comparable <strong>to</strong> or smaller than the diffusion layer of analyte under study [2]. In a<br />

thin-layer flow cell, solution is constrained <strong>to</strong> a film passing <strong>over</strong> a planar<br />

electrode held at a fixed potential. Figure 5.1 illustrates an exploded view of the<br />

thin layer region. An electrochemically active substance passes <strong>over</strong> an<br />

electrode held at a potential sufficiently great (positive or negative) for an<br />

electron transfer (either oxidation or reduction) <strong>to</strong> occur. An amperometric<br />

current is produced that is directly proportional <strong>to</strong> the concentration of the analyte<br />

entering the thin layer cell [1].<br />

There are many electrochemical studies, including quantification of<br />

product and by-product formation during methanol electrochemical reaction,<br />

investigations of adsorption, electrodeposition, complex reaction mechanisms,<br />

102


and n-value determinations where thin layer cells have been widely used. In<br />

addition <strong>to</strong> the benefit of conservation of valuable sample, thin-layer cells are<br />

also very useful for spectroelectrochemical analysis, because the entire solution<br />

in the thin-layer can be rapidly and completely electrolyzed and this makes it<br />

possible <strong>to</strong> obtain spectroscopic information without interference from unreacted<br />

analyte in bulk solution [3].<br />

In the present studies, a thin layer spectrochemical cell was adapted for<br />

electrolysis measurements. The conventional flat window was modified <strong>to</strong><br />

accommodate a 50 µL sample volume. In early work a window was constructed<br />

from Kel-F. However, the permeability of Kel-F <strong>to</strong> atmospheric O2 potentially<br />

limits electrolysis measurements at low (≤ 0.5VRHE) potentials. As part of the<br />

study, a new window materials, MACOR, was tested due <strong>to</strong> its extremely low<br />

permeability <strong>to</strong> gases and easy machinability. A well known redox system, [Ru<br />

(NH3)6] 2+/3+ was employed <strong>to</strong> characterize the cell using cyclic voltammetry. The<br />

cell was used in investigations of methanol electrochemical conversion <strong>to</strong> H2CO<br />

as discussed in Chapters III and IV.<br />

Experimental and Instrumentation<br />

Reagents<br />

Sodium Nitrate (NaNO3, Mallinckrodt Chemicals work, Hazelwood, MO)<br />

and sulfuric acid (H2SO4) solutions were prepared from distilled water that was<br />

further processed with a 4-cartidge Nanopure II system (Barnstead, Dubuque,<br />

IA). The Sodium nitrate was reagent grade and used as received. Hexaammine<br />

ruthenium (III) trichloride (Ru (NH3)6Cl3, Ru 32.6%) was obtained from Alfa Aesar<br />

103


Figure 5.1: Exploded view of electron transfer at an electrode surface in a thinlayer<br />

flow cell [1].<br />

104


in highest purity and used as received. Solutions were freshly prepared prior <strong>to</strong><br />

measurements. The H2SO4 was obtained in 99.999% purity from Aldrich<br />

(Milwaukee, WI). Argon (Trinity Gases, Dallas, TX) was of ultrahigh purity.<br />

Small Volume <strong>Electrochemical</strong> Cell<br />

A diagram showing the small volume region of the electrochemical cell<br />

employed for CH3OH electrolysis was included in Chapter II (Figure 2.2). An<br />

infrared spectroelectrochemical cell was adapted by substituting a conventional<br />

window for one that had been modified <strong>to</strong> accommodate a fixed volume of 50 µL.<br />

The general design of the cell had been reported earlier [7, 8]. As part of this<br />

project, improvements were made in reducing the O2 permeability of the cell.<br />

Figure 5.2 showed the side view and <strong>to</strong>p view of cell window denoting the<br />

detailed dimensions of each section. The working electrode was positioned<br />

adjacent <strong>to</strong> the cavity in a manner that entraps solution. The <strong>to</strong>tal well depth is<br />

2.54 mm. There is a 1.9 mm recess cut in<strong>to</strong> the <strong>to</strong>p edge of the well that helps in<br />

positioning the working electrode reproducibly with the metal disk facing the well<br />

and in contact with the solution. The diameter of the outer well and the inner well<br />

is 12.7 mm and 9.9 mm, respectively. A drilled hole (1.09 mm diameter) through<br />

the disk center allows the transfer of solution through a 23 gauge needle of a gas<br />

tight syringe (Hamil<strong>to</strong>n Company, Reno, Nevada).<br />

Choice of Cell Material and Cell Fabrication<br />

In studying methanol electrooxidation pathways wheather <strong>to</strong> by-products<br />

such as H2CO or the main product CO2, dissolved O2 in the reactant solution can<br />

105


undergo reduction and thereby cause error in the determination of the methanol<br />

oxidation charge. In acidic solution, O2 reduction can progress <strong>to</strong> H2O or H2O2<br />

depending upon the working electrode material and potential, as shown in Eq. 1:<br />

O2 + 2 e - + 2 H + ⎯→ H2O2 [1a]<br />

O2 + 4 e - + 4 H + ⎯→ 2 H2O [1b]<br />

When O2 reduction occurs simultaneously with CH3OH oxidation, the charge<br />

necessary for reaction of O2 off-sets the charge produced in the oxidation<br />

reaction. As a result, the measured charge is lower than the actual charge<br />

passed in the CH3OH oxidation process.<br />

One of the objectives of this project was <strong>to</strong> find a material that is both<br />

impermeable <strong>to</strong> O2 and easily machinable. In the cell used earlier [7, 8], a Kel-F<br />

disk was used as the cell window, because it could be precision machined <strong>to</strong><br />

create the needed 50 µL volume. However, a major limitation of Kel-F in this<br />

application is that it is permeable <strong>to</strong> gases and allows diffusion of O2, in<strong>to</strong> the<br />

cell. The O2 interference becomes problematic as low (below 0.5 VRHE) CH3OH<br />

oxidation potentials are reached. As an alternative <strong>to</strong> Kel-F, we initially explored<br />

the use of glass. Glass windows (25.4 mm diameter by, 6.35 mm thickness) for<br />

use in UV-visiible spectroscopy cells could be easily obtained commercially. The<br />

disks were impermeable <strong>to</strong> O2 and had excellent resistance <strong>to</strong> attack by<br />

chemicals. However, they tended <strong>to</strong> break easily during machining, processes<br />

which involved grinding the glass by pressing the flat end of a high precision cut<br />

metal rod against an abrasive powder on the glass surface. A milling machine<br />

106


well depth =<br />

0.100 "<br />

drilled hole =<br />

0.043 " dia.<br />

drilled hole =<br />

0.043 " dia.<br />

0.5 "<br />

1.0 "<br />

Side View<br />

0.50"<br />

1.0 "<br />

Top View<br />

107<br />

well depth =<br />

0.075 "<br />

0.25 "<br />

0.39"<br />

Figure 5.2: Side view and <strong>to</strong>p view of the electrochemical cell window.


was used <strong>to</strong> carefully align the rod and glass disk during the grinding.<br />

Further searching brought <strong>to</strong> light the glass ceramic MACOR as a<br />

potential material. MACOR is a white, odorless, porcelain-like material<br />

composed of approximately 55% fluorphlogopite mica (KMg3AlSi3O10F2) in a<br />

borosilicate glass matrix. It can be melted and cast in<strong>to</strong> various forms using<br />

conventional glass making techniques. Like Kel-F, MACOR can be easily<br />

machined. It has extremely low O2 permeability and can be purchased in disk<br />

forms the size of glass optical windows. Thus, MACOR was tested as a thin<br />

layer cell material.<br />

Results and Discussion<br />

Cell Characterization in 0.1 M H2SO4<br />

The cell was first tested for cleanliness and response in a conventional<br />

three electrode electrochemical cell configuration. Cyclic voltammograms were<br />

recorded with mirror polished polycrystalline Pt electrode in a conventional<br />

electrochemical cell and then in the electrolysis cell with the electrodes placed at<br />

different positions relative <strong>to</strong> the MACOR window. Cyclic voltammograms were<br />

recorded between -0.2 V and 1.3 V vs. a Ag/AgCl reference electrode and then<br />

the potentials were converted <strong>to</strong> the RHE scale. Figure 5.3 compares<br />

voltammograms obtained at the scan rate of 50 mV/s in a conventional<br />

electrochemical cell [Figure 5.3 (a)] and the electrolysis cell when the electrode<br />

was mounted far from MACOR window [Figure 5.3 (b)]. The two voltammograms<br />

have a similar appearance. The important features are (i) the defined hydrogen<br />

adsorption and hydrogen desorption peaks in the so called hydrogen region<br />

108


etween 0.0-0.3 VRHE, (ii) the flat double layer charging region <strong>over</strong> the potential<br />

range 0.3-0.7 VRHE, (iii) the peaks associated with Pt oxidation which start at ~<br />

0.95 VRHE, and (iv) the Pt reduction peak at ~ 0.65 VRHE. There is a small drop in<br />

current density for the oxide stripping peak (from 280 µA/cm 2 <strong>to</strong> 250 µA/cm 2 ) and<br />

in hydrogen adsorption and desorption regions for the electrode in the<br />

electrolysis cell. Otherwise, the voltammograms exhibit good agreement with the<br />

well-known current-voltage response of clean polycrystalline Pt electrodes [9]. It<br />

can be concluded that when the electrode is positioned in the electrolysis cell far<br />

from the MACOR window, the cell performance is similar <strong>to</strong> that of the<br />

conventional cell.<br />

In Figure 5.4, cyclic voltammograms recorded using the electrolysis cell<br />

and a polycrystalline Pt electrode positioned very near <strong>to</strong> the MACOR window<br />

are shown. The voltammograms were recorded at two different scan rates: 10<br />

mV/s [Figure 5.4(a)] and 50 mV/s [Figure 5.4 (b)]. The peak currents become<br />

lower with decreasing scan rate, as expected, and the position of peaks in the<br />

oxide stripping (~ 0.6 <strong>to</strong> 0.8 VRHE) and hydrogen region (~ 0.0 <strong>to</strong> 0.3 VRHE) does<br />

not change.<br />

Figure 5.5 shows comparison voltammograms for the Pt electrode when it<br />

is positioned outside the MACOR window [Figure 5.5 (a)] and after insertion in<strong>to</strong><br />

the MACOR cavity [Figure 5.5 (b)]. The scan rate was slowed <strong>to</strong> 10 mV/s <strong>to</strong><br />

reduce dis<strong>to</strong>rtions in the voltammogram that would be caused by resistive effects<br />

when the electrode is inside the cavity. Looking at the cyclic voltammogram in<br />

the MACOR window [Figure 5.5 (b)], there is a skewing of the hydrogen<br />

109


adsorption region (~ at 0.03 VRHE, pointed by gray arrow) which is due <strong>to</strong> some<br />

resistance drop. The effect was not observed in the cyclic voltammogram<br />

recorded with the electrode positioned far from the cavity [Figure 5.4 and Figure<br />

5.5 (a)]. In both cases [Figure 5.5 (a) and (b)], oxide reduction peaks were well<br />

resolved and observed at ~ 0.65 VRHE. However, there is a small feature at ~<br />

1.1VRHE that could indicate some contamination arising from the MACOR<br />

material [Figure 5.5 (b)]. This effect was ignored in our study as this<br />

contamination was not seen in the cell characterization part [Figure 5.6 (c) and<br />

(d)] and quantitative analysis through electrolysis measurements (Chapter III and<br />

IV).<br />

Finally, based on the absence of any significant dis<strong>to</strong>rtions, the cyclic<br />

voltammograms presented in Figure 5.3, 5.4 and 5.5 demonstrate that the<br />

electrochemical measurements can be carried out in this thin layer<br />

electrochemical cell configuration constructed with MACOR material without<br />

considerable ohmic drops.<br />

Cell Characterization using a Reversible Redox Active Probe<br />

The response of the thin layer cell used in this work was further<br />

characterized by performing cyclic voltammetry measurements using a soluble,<br />

reversible redox probe, hexamine ruthenium trichloride [Ru(NH3)6Cl3]. The<br />

voltammograms were compared <strong>to</strong> the response of the redox reaction in a<br />

conventional electrochemical cell operating under the same conditions. The<br />

working electrode employed for these measurements was a polycrystalline gold<br />

disk (0.481 cm 2 ) that had been pressure sealed in<strong>to</strong> the end of Kel-F rod<br />

110


Current<br />

Polycrystalline Pt-solid Electrode<br />

in 0.1 M H2SO4<br />

0.1 mA / cm 2<br />

0<br />

0<br />

0.0 0.4 0.8 1.2 1.6<br />

E / V vs. RHE<br />

111<br />

(a) conventional<br />

electrochemical<br />

cell<br />

(b) Electrolysis Cell<br />

Figure 5.3: Cyclic voltammograms of a polycrystalline Pt electrode in 0.1 M<br />

H2SO4 recorded in a conventional electrochemical cell (a) and the electrolysis<br />

cell with the electrode mounted far from MACOR window (b). The scan rate was<br />

50 mV/s.


Current<br />

Polycrystalline Pt-solid Electrode<br />

in 0.1 M H2SO4<br />

0.1 mA / cm 2<br />

0<br />

0<br />

(a) Electrolysis Cell,10 mV/s<br />

(b) Electrolysis Cell, 50 mV/s<br />

0.0 0.4 0.8 1.2 1.6<br />

E / V vs. RHE<br />

112<br />

X 2.5<br />

Figure 5.4: Cyclic voltammograms of a polycrystalline Pt electrode in 0.1 M<br />

H2SO4 recorded in an electrolysis cell with the electrode mounted near the<br />

MACOR window. Scan rate was 10 mV/s (a) and 50 mV/s (b).


Current<br />

Polycrystalline Pt-solid Electrode<br />

in 0.1 M H2SO4<br />

2<br />

0.1 mA / cm<br />

0<br />

0<br />

(a) Electrolysis Cell (Far from<br />

MACOR window),10mV/s<br />

(b) Electrolysis Cell (MACOR<br />

window),10mV/s<br />

0.0 0.4 0.8 1.2 1.6<br />

E / V vs. RHE<br />

Figure 5.5: Cyclic voltammograms of a polycrystalline Pt electrode in 0.1 M<br />

H2SO4 recorded in an electrolysis cell with the electrode mounted near the<br />

MACOR window (a) and mounted inside the MACOR window (Arrow represents<br />

skewing of the hydrogen adsorption region) (b). The scan rate was 10 mV/s.<br />

113<br />

X 3<br />

X 4


(Boedeker Plastics, Shiner, TX). The voltammetric responses were measured by<br />

cycling between -0.3 and 0.4 VAg/AgCl and using [Ru(NH3)6Cl3] solution with 0.1 M<br />

NaNO3 as a supporting electrolyte [Figure 5.6]. Various probe concentrations<br />

and scan rates were tested. The reaction equation can be written as follows:<br />

[Ru(NH3)6] 3+ + e - ↔ [Ru(NH3)6] 2+<br />

The cyclic voltammograms presented in Figure 5.6 exhibit the characteristic<br />

reduction peak at ~0.05 VAg/AgCl and oxidation peak at ~ 0.12 VAg/AgCl. In Figure<br />

5.6 (a) and (b), the voltammograms were recorded in a conventional<br />

electrochemical cell. At the scan rate of 50 mV/s [Figure 5.6 (a)] with 5.0 x 10 -3<br />

M [Ru(NH3)6] 3+ , the response appears reversible and limited by linear diffusion.<br />

Slowing the scan rate <strong>to</strong> 0.2 mV/s, [Figure 5.6 (b)], gives a response closer <strong>to</strong><br />

that of steady-state. A major concern when working with thin layer cells,<br />

especially when nonaqueous solutions or very low supporting electrolyte<br />

concentrations are employed, is the high resistance of solution in the cavity. In<br />

the measurements undertaken, the reference and counter electrodes were<br />

located in the bulk solution outside the thin layer and far from the working<br />

electrode. Thus, the scan rate was slowed <strong>to</strong> compensate for the slower cell<br />

response time due <strong>to</strong> the high resistance. Figure 5.6 (c) shows the<br />

voltammogram recorded at this slower scan rate (0.2 mV/s) with the electrode<br />

positioned in the MACOR window. However, the long time required <strong>to</strong> complete<br />

the scan and limited quantity of reactant in the cavity (2.5 x 10 -9 nmols / µL x 50<br />

µL) resulted in analyte depletion during the scan. Thus, the voltammogram in Fig<br />

5.6c shows the effect of the decreasing [Ru(NH3)6] 3+ .<br />

114


To <strong>over</strong>come the problem of reactant depletion, a higher [Ru(NH3)6] 3+<br />

concentration (2.5 x 10 -2 M) and faster scan rates (5-10 mV/s) were tested (data<br />

not shown). These conditions resulted in a diffusion controlled response<br />

affected by cell resistance. However, effects of reactant depletion were much<br />

less, as completing the forward and reverse scans required less time and thus<br />

less charge was passed. To maintain a balance between cell resistance and<br />

reactant depletion, scan rates slower than 5-10 mV/s were attempted. A<br />

reversible voltammetric response that appeared <strong>to</strong> be limited by linear diffusion<br />

[Figure 5.6 (d)] was finally established by employing the slower scan rate (1<br />

mV/s) with 2.5 x 10 -2 M [Ru(NH3)6] 3+ . The voltammogram showed some<br />

depletion of reactants from the cavity, but very low cell resistance effects.<br />

Peak separation is a unique characteristic of a thin-layer electrochemical<br />

cell response in the cyclic voltammetry of a reversible redox couple [1]. In<br />

theory, the peak potentials for the reduction and oxidation waves should be the<br />

same in an ideal thin layer cell when the cell boundary falls within the analyte<br />

diffusion layer. However, in practice, some peak separation occurs due <strong>to</strong><br />

uncompensated solution resistance [10]. In Figure 5.6 (d), the peak separation is<br />

about 79 mV (and 66 mV in Figure 5.6 (c)), which is close <strong>to</strong> the theoretical peak<br />

separation for a cell operating under linear diffusion control (∆Ep = Epa-Epc = 57/n<br />

mV, where Epa (anodic) and Epc (cathodic) peak potentials) [11]. The result<br />

indicates that while the thin layer cavity has a small volume, the cell does not<br />

display true thin layer behavior, since the boundaries are somewhat large<br />

compared <strong>to</strong> reactant diffusion layer under the conditions of the measurements<br />

115


Current<br />

200 µA/cm 2<br />

0<br />

0<br />

0<br />

0<br />

-0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4<br />

E/V vs. Ag/AgCl<br />

116<br />

(a) 50 mV/s<br />

(b) 0.2 mV/s<br />

x 5<br />

(c) 0.2 mV/s<br />

x 50<br />

(d) 1 mV/s<br />

Figure 5.6: Cyclic voltammograms of [Ru(NH3)6Cl3] in 0.1 M NaNO3 recorded in<br />

a conventional electrochemical cell (a, b) and the MACOR cavity of the thin layer<br />

electrolysis cell shown in Figure 3.2 (c, d). The [Ru(NH3)6] 3+ concentration and<br />

scan rates were : (a) 5 × 10 -3 M and 50 mV/s , (b) 5 × 10 -3 M and 0.2 mV/s (c) 2.5<br />

× 10 -3 M and 0.2 mV/s (d) 2.5 ×1 0 -2 M and 1 mV/s. The scans started from 0.4<br />

VAg/AgCl and swept <strong>to</strong>ward more negative potentials.<br />

x 10


Nevertheless, the voltammograms in Figure 5.6 display equal oxidation and<br />

reduction peak heights, which is a diagnostic for reversibility. Considering peak<br />

separation and peak height in the cyclic voltammograms, both the conventional<br />

and the thin layer cells showed comparable results. Additionally, our thin layer<br />

cell is much better than other recent silicon based designs fabricated through<br />

micromachining technology. In this latter case, the reported peak separation<br />

value is 250 mV [12].<br />

Conclusions<br />

This project advanced the design of a small volume electrolysis cell used<br />

in prior studies by incorporating a machinable MACOR glass ceramic disk<br />

window, which resists oxygen permeation. The cell response was characterized<br />

by observing hydrogen adsorption/desorption and oxide formation waves on Pt in<br />

0.1 M H2SO4 and the reversible, diffusion controlled waves for hexamine<br />

ruthenium trichloride [Ru(NH3)6Cl3] in cyclic voltammetry measurements.<br />

117


References<br />

1. Heineman, W. R.; Kissinger, P. T. Chapter 3, Labora<strong>to</strong>ry Techniques in<br />

Electroanalytical Chemistry; 2nd Edition ed.; MARCEL DEKKER, INC.:<br />

New York, 1996.<br />

2. Bard, A. J.; Faulkner, L. R. <strong>Electrochemical</strong> Methods; Wiley: New York,<br />

1980.<br />

3. Hubbard, A. T.; F.C., A. Electranal. Chem. 1970, 4.<br />

4. Seki, H.; Kunimatsu, K.; Golden, W. G. Appl. Spectrosc. 1985, 39, 437.<br />

5. Bethune, D. S.; Luntz, A. C.; Sass, J. K.; Roe, D. K. Surf. Sci. 1988, 197,<br />

44.<br />

6. Popenoe, D. D.; S<strong>to</strong>le, S. M.; Porter, M. D. Appl. Spectrosc. 1992, 46, 79.<br />

7. Huang, H.; Korzeniewski, C.; Vijayaraghavan, G. Electrochim. Acta 2002,<br />

42, 3675.<br />

8. Gao, L.; Huang, H.; Korzeniewski, C. Electrochim. Acta 2004, 49, 1281.<br />

9. Angerstein-Kozlowska, H.; Conway, B. E.; Sharp, W. B. A. J. Electroanal.<br />

Chem. 1973, 43, 9.<br />

10. Tom, G. M.; Hubbard, A. T. Anal. Chem. 1971, 43, 671.<br />

11. Evans, D. H.; O'Connel, K. M.; Peterson, R. A.; Kelly, M. J. Journal of<br />

Chemical Education 1983, 60, 290.<br />

12. Yun, K.-S.; Joo, S.; H-J, K.; Kwak, J.; Yoon, E. Electroanalysis 2005, 17,<br />

959.<br />

118


CHAPTER VI<br />

SUMMARY<br />

The conversion of methanol <strong>to</strong> H2CO on Pt and PtRu based electrode<br />

materials was investigated with the use of electrochemical methods and<br />

fluorescence spectroscopy. <strong>Methanol</strong> is a promising fuel for PEM type fuel cells<br />

and investigation of its electrochemistry <strong>over</strong> various electrode catalyst materials<br />

is essential <strong>to</strong> establish guidelines for DMFC research. The complete oxidation<br />

of methanol <strong>to</strong> CO2 is a multi-step process (Figure 1.4), and thus, the DMFC<br />

suffers limitations from kinetically slow steps in these pathways. All simple<br />

alcohols considered for fuel cell use (CH3OH, CH3CH2OH, C2H2(OH)2, etc) react<br />

at low <strong>over</strong>potentials <strong>to</strong> form aldehydes. For methanol, the H2CO generated is<br />

challenging <strong>to</strong> detect experimentally, but it is recognized as a significant by-<br />

product [1-4]. As a species formed in the early stages of methanol oxidation,<br />

H2CO accumulation lowers the energy conversion efficiency of the fuel. The<br />

primary in situ techniques used <strong>to</strong> study electrochemical reactions at the<br />

molecular level, infrared spectroscopy and mass spectroscopy, are not at all<br />

sensitive <strong>to</strong> H2CO. Therefore, in this project a sensitive fluorescence assay with<br />

good precision was adapted <strong>to</strong> detect and quantify H2CO.<br />

In the first part of the project, the cell that had been used in prior<br />

investigations of methanol oxidation pathways was improved by making it less<br />

permeable <strong>to</strong> atmospheric O2 and thereby less susceptible <strong>to</strong> interferences from<br />

O2 reduction in measurements at low electrode potentials. A MACOR glass<br />

119


ceramic disk was used in place of a Kel-F window that formed one side of a thin<br />

layer electrolysis cell. The MACOR was easily machined and enabled a fixed<br />

volume (50 µL) solution cavity <strong>to</strong> be created for the placement of sample during<br />

electrolysis. In contrast <strong>to</strong> Teflon or Kel-F, which are also easily machined and<br />

resistant <strong>to</strong> attack by most solvents, MACOR also possessed these properties<br />

and in addition did not allow the transport of atmospheric O2 in<strong>to</strong> the cell.<br />

<strong>Methanol</strong> oxidation studies reported in this thesis were perfomed with the use of<br />

the MACOR cell. Final experiments reported demonstrate quantitative aspects of<br />

cell performance based on simple cyclic voltammetric measurements using<br />

hexaammine ruthenium (III) trichloride [Ru(NH3)6Cl3] as a redox probe.<br />

The cavity was shown <strong>to</strong> be somewhat thicker (~0.5 mm) than that for an Ideal<br />

thin layer cell. However, reversible, diffusion controlled cyclic voltammetry was<br />

achieved at scan rates of about 1 mV/s.<br />

The studies reported focused on applications of the cell in controlled<br />

potential electrolysis measurements of methanol oxidation <strong>over</strong> Pt and PtRu bulk<br />

metal and nanometer scale catalysts materials. Yields of H2CO produced during<br />

the electrolysis of 1.0 M methanol in 0.1 M H2SO4 for periods of 180 s were<br />

determined. In measurements with solid electrodes, it was observed that yields<br />

of H2CO greater than 10 % form at low potentials, especially <strong>over</strong> polycrystalline<br />

Pt, and can account for as much as 78% of the reaction charge. Figure 6.1<br />

compares the percent yield values graphically for H2CO produced during<br />

methanol electrochemical oxidation on all of the Pt and PtRu bulk electrodes<br />

studied. The most notable characteristic is the comparatively lower H2CO %<br />

120


yields (13% for XRu=0.3) for methanol oxidation on PtRu-solid compared <strong>to</strong> Pt<br />

solid electrode. Quantitative detection either for H2CO or CO2 becomes<br />

susceptible at lower electrode potentials on pure Pt catalysts, while on pure PtRu<br />

with optimum Ru composition (XRu=0.3) H2CO or CO2 can be assayed<br />

effortlessly at potentials below 0.6 VRHE. From these results, it can be concluded<br />

that the Ru metal with appropriate composition plays an important role in<br />

facilitating the complete oxidation of methanol. The dissociation of C-H bonds<br />

requires and ensemble of Pt sites. However, Ru sites lowers the potential<br />

(compared <strong>to</strong> Pt) for water activation and facilitates the conversion of methanol<br />

and any H2CO formed <strong>to</strong> more complete oxidation products, such as CO2 and<br />

formic acid (HCOOH) [5, 6]. It is predicted that Ru may break the H-OH bond in<br />

water more effectively than Pt <strong>to</strong> form the surface oxides necessary for the<br />

reaction <strong>to</strong> progress far along the path <strong>to</strong> CO2. In other words, Ru a<strong>to</strong>ms<br />

dissociate water more easily than Pt <strong>to</strong> create oxygen species that lower H2CO<br />

production [6]. With higher concentration of Ru (XRu=0.9) in solid PtRu electrode<br />

the catalytic activity was inferior as it produces higher H2CO yield at all electrode<br />

potential compared <strong>to</strong> PtRu –solid (XRu=0.3) electrode. It is perhaps due <strong>to</strong> the<br />

surface oxides (RuOx) forming on Ru that are inactive for bond cleavage [5-8].<br />

This may hinder the reaction and can cause large capacitance, which we<br />

observe in the cyclic voltammograms (Figure 3.1d, Figure 4.3a and Figure 4.3c)<br />

through out the “double layer charging region”. This is consistent with data<br />

previously reported [6, 8].<br />

121


Following work with bulk electrodes, subsequent experiments extended<br />

the approach <strong>to</strong> investigate methanol electrochemical oxidation <strong>to</strong> H2CO on Pt<br />

and PtRu based nanometer scale catalyst materials. The results for all catalysts<br />

are summarized in Figure 6.2 graphically. Looking at Figure 6.1 and 6.2, it is<br />

noted that the H2CO yields with catalyst materials are considerably lower (below<br />

10% for the conditioned studied) than that of solid metal electrodes, which is<br />

consistent with earlier studies [9]. The CO2 measurements by IR spectrometry<br />

also established that the CO2 yield reached above 90% with rough catalyst<br />

surfaces [10]. This difference in H2CO yield on the bulk electrode and catalysts<br />

is believed <strong>to</strong> result from dissimilarities in the three dimensional structure of the<br />

electrode materials, and this <strong>to</strong>pic will be discussed elaborately in the next<br />

section. It is also important <strong>to</strong> observe that the reaction charges for catalysts are<br />

significantly higher than those of bulk metal electrodes at all potentials (Table 3.1<br />

in Chapter III; Table 4.1 and 4.2 in Chapter IV). For example, at 0.8 VRHE, the<br />

highest reaction charge was 153 mC for nano particles whereas reactions <strong>over</strong><br />

solid electrodes led <strong>to</strong> fairly low charges in the vicinity of 25 mC. The same<br />

pattern was observed for the lowest electrode potential (0.5 VRHE). The catalyst<br />

particles have nanometer size dimensions (~2-6 nm observed by transmission<br />

electron microscopy) with higher surface areas than the bulk electrodes. This<br />

higher surface area is at least partly responsible for the higher reaction charges<br />

recorded. The lower H2CO yields recorded for reactions <strong>over</strong> the catalyst<br />

particles likely result from the three dimensional nature of the catalyst bed, since<br />

partial oxidation products formed at particles deep within the bed can undergo<br />

122


<strong>Formaldehyde</strong> yield%<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

<strong>Formaldehyde</strong> Yields for Oxidation of 1.0 M CH 3OH <strong>over</strong> Solid <strong>Bulk</strong> Electrode Materials<br />

Pt - solid<br />

PtRu - solid (XRu = 0.1)<br />

PtRu - solid (XRu = 0.3)<br />

PtRu - solid (XRu = 0.9)<br />

0.8 0.7 0.6 0.5<br />

E/V (vs. RHE)<br />

Figure 6.1: Bar charts comparing the formaldehyde yields from reaction of 1.0 M<br />

CH3OH in 0.1 M H2SO4 during 180 s electrolysis periods for solid bulk electrodes.<br />

123


<strong>Formaldehyde</strong> Yield%<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

<strong>Formaldehyde</strong> Yields for Oxidation of 1.0 M CH 3OH <strong>over</strong> Nanoscale Catalyst Materials<br />

Pt-black<br />

C/Pt-10 wt%<br />

JM PtRu, XRu = 0.5<br />

SC PtRu, XRu = 0.5<br />

SC PtRu, XRu = 0.25<br />

SC PtRu, XRu = 0.1<br />

SC PtRu, XRu = 0.1,<br />

with nafion<br />

0.8 0.7 0.6 0.5<br />

E/V vs. RHE<br />

Figure 6.2: Bar charts comparing the formaldehyde yields from reaction of 1.0 M<br />

CH3OH in 0.1 M H2SO4 <strong>over</strong> nanoscale catalyst materials during 180 s reaction<br />

periods.<br />

124


further reaction through interactions with other particles encountered along the<br />

diffusion path <strong>to</strong>ward bulk solution. Furthermore, although the effect is small, the<br />

yields of H2CO appear <strong>to</strong> increase <strong>to</strong>ward lower <strong>over</strong>potentials at nanometer<br />

scale PtRu particles, similar <strong>to</strong> the trend displayed for the bulk electrodes.<br />

Comparing Figure 6.1 and Figure 6.2, there is a marked difference<br />

observed in the electrochemistry for methanol on bulk, solid electrodes and<br />

supported nanoparticle catalysts and it represents the more elaborate pictures of<br />

catalytic activity of the materials used for H2CO production. The spatial<br />

arrangement of particles and the diffusion conditions play critical roles in<br />

promoting methanol oxidation. The transport of reactants and products <strong>to</strong> and<br />

away from the electrode surface appears <strong>to</strong> be influenced by the particle layer<br />

properties as a concentration gradient is generated between the electrode<br />

surface and the reaction solution during electrolysis. For bulk metal electrodes,<br />

the transport of reactant occurs via planner diffusion, where reactant can only<br />

arrive following a perpendicular path relative <strong>to</strong> the electrode surface and can<br />

only react with the <strong>to</strong>pmost a<strong>to</strong>ms of the catalyst surface. After completing the<br />

reaction, the product can diffuse away <strong>to</strong> the bulk solution following the same<br />

perpendicular path and will most likely not encounter catalytic sites again.<br />

Chances that such an escape of the by-product without further oxidization will<br />

happen are much higher on smooth than on rough surfaces [11-14]. For<br />

nanoparticle metal cluster catalysts, there is a hemispherical diffusion where<br />

each site can behave similar <strong>to</strong> a small microelectrode, and the number of<br />

electroactive molecules having access <strong>to</strong> this hemispherical catalyst greatly<br />

125


exceeds that of a planner electrode [15]. From hemispherical volumes, the<br />

reaction analyte will be drawn around the electrode surface, and there is the<br />

possibility of more reactant-catalyst encounters as the reactant diffuses <strong>to</strong>ward or<br />

away from the site [3]. The higher chances for multiple encounters move the<br />

methanol oxidation reaction <strong>to</strong> completion and can lead <strong>to</strong> lower yield of H2CO<br />

compared <strong>to</strong> solid metal electrodes. In other words, with multiple catalyst<br />

particles present in multilayer structure, readsorption of H2CO becomes<br />

increasingly more effective and increases the probability for the oxidation of<br />

volatile, oxidizable reaction intermediates <strong>to</strong> major products [2, 16-18]. There is<br />

another considerable hypothesis, which is the effect of the gradient diffusion<br />

layers close <strong>to</strong> and away from the particles surfaces. In the case of<br />

hemispherical diffusion, the gradient lines of the two particles diffusion patterns<br />

<strong>over</strong>lap. So there is a possibility that the reaction product spreads away from the<br />

first particle, it can encounter the other particle’s gradients and be pulled <strong>to</strong>ward<br />

that site <strong>to</strong> react further. A more complete oxidation of the reactant should be<br />

occurred by this effect since byproducts again have more chances <strong>to</strong> encounter<br />

other catalyst sites [2, 3]. Actually this would cause <strong>to</strong> remove more H2CO from<br />

the solution and thus give the very low yields compared <strong>to</strong> that of solid<br />

electrodes.<br />

In the case of catalyst system, SC PtRu, XRu= 0.1 with Nafion, the nano<br />

catalyst structure is confined in a permeable film. The reaction environment is<br />

surrounded by three dimensional arrays of catalyst particles. As reaction<br />

proceeds, there is a spherical diffusion of reactants <strong>to</strong> each catalyst sites <strong>to</strong><br />

126


eact. The partial products like H2CO, entrapping inside the three dimensional<br />

atmosphere, can just diffuse away from the site and feasibly meet another site for<br />

further oxidization <strong>to</strong> CO2. Thus the soluble partial oxidation products experience<br />

multiple encounters with the array of catalyst sites inside the film and the<br />

consequence is the increase of the chances for reactions <strong>to</strong> progress <strong>to</strong><br />

completion with lower H2CO yield.<br />

The results obtained during this research project have opened several<br />

new avenues for research and development. The cell that has been used<br />

throughout these investigations can be further optimized with the modification of<br />

cell dimension which can offer reduced cell resistance and increased thin layer<br />

response. The proposed cell window with a novel MACOR material simplifies the<br />

cell fabrication process and the assembly of the spectrochemical cell. In the<br />

future, an integral piece of MACOR can be explored <strong>to</strong> construct the whole cell<br />

where the oxygen interference problem can be further minimized. The use of<br />

MACOR cavity can also be extended <strong>to</strong> quantify the other oxidation products like<br />

CO2 and HCOOH which can be detected off-line by IR spectrometry and<br />

chroma<strong>to</strong>graphic techniques. It will be very advantageous if the micro dimension<br />

cell can be utilized for studying the oxidation of CO with different electrode<br />

materials and the results can be compared with that of CO stripping process<br />

using conventional electrochemical cell.<br />

127


References<br />

1. Jarvi, T. D.; Stuve, E. M., In Electrocatalysis; Lipkowski, J. and Ross, P.<br />

N., Ed.; Wiley-VCH Publishers: New York, 1998; Vol. Ch. 3, pp 75.<br />

2. Jusys, Z.; Behm, R. J. J. Phys. Chem. B 2001, 105, 10874.<br />

3. Jusys, Z.; Kaiser, J.; Behm, R. J. Electrochim. Acta 2002, 47, 3693.<br />

4. Chen, Y. X.; Miki, A.; Ye, S.; Sakai, H.; Osawa, M. J. Am. Chem. Soc.<br />

2003, 125, 3680.<br />

5. Gasteiger, H. A.; Markovic, N.; Ross, P. N., Jr.; Cairns, E. J. J. Phys.<br />

Chem. 1993, 97, 12020.<br />

6. Rolison, D. R.; Hagans, P. L.; Swider, K. E.; Long, J. W. Langmuir 1999,<br />

15, 774.<br />

7. Long, J. W.; Rolison, D. R., In New Directions in Electroanalytical<br />

Chemistry II; Leddy, J., Vanysek, P. and Porter, M. D., Ed.;<br />

<strong>Electrochemical</strong> Society: Penning<strong>to</strong>n, NJ, 1999; Vol. PV99-5, pp 125.<br />

8. Rolison, D. R.; Long, J. W.; Swider, K. E.; Merzbacher, C. I. Langmuir<br />

1999, 15, 780.<br />

9. Childers, C. L.; Huang, H.; Korzeniewski, C. Langmuir 1999, 15, 786.<br />

10. Gao, L.; Huang, H.; Korzeniewski, C. Electrochim. Acta 2004, 49, 1281.<br />

11. Lu, G.-Q.; Chrzanowski, W.; Wieckowski, A. J. Phys. Chem. B 2000, 104,<br />

5566.<br />

12. Wang, H.; Loffler, T.; Baltruschat, H. J. Appl. Electrochem. 2001, 31, 759.<br />

13. Laborde, H.; Leger, J.-M.; Lamy, C.; Garnier, F.; Yassar, A. J. Appl.<br />

Electrochem. 1990, 20, 520.<br />

14. Lin, W.-F.; Wang, J.-T.; Savinell, R. F. J. Electrochem. Soc. 1997, 144,<br />

1917.<br />

15. Wightman, R.; Wipf, D. Electroanalytical Chemistry 1989, 15, 267.<br />

16. Wang, H.; Wingender, C.; Baltruschat, H.; Lopez, M.; Reetz, M. T. J.<br />

Electroanal. Chem. 2001, 509, 163.<br />

128


17. Jusys, Z.; Kaiser, J.; Behm, R. J. Langmuir 2003, 19, 6759.<br />

18. Wang, H.; jusys, Z.; Behm, R. J. J. Phys. Chem. B 2004, 108, 19413.<br />

129

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!