Download as pdf or txt
Download as pdf or txt
You are on page 1of 204

ARENBERG DOCTORAL SCHOOL

Faculty of Engineering Science

Electroprecipitation of metal
ions from aqueous
electrolytes
Kinetic study and process performance

Sam Eggermont

Supervisors: Dissertation presented in partial


Prof. dr. ir. J. Fransaer fulfillment of the requirements for the
Prof. dr. ir. X. Dominguez Benetton degree of Doctor of Engineering
Science (PhD):
Materials Engineering

May 2020
Electroprecipitation of metal ions from aqueous
electrolytes
Kinetic study and process performance

Sam EGGERMONT

Examination committee: Dissertation presented in partial


Prof. dr. ir. R. Puers, chair fulfillment of the requirements for
Prof. dr. ir. J. Fransaer, supervisor the degree of Doctor of
Prof. dr. ir. X. Dominguez Benetton, Engineering Science (PhD):
co-supervisor Materials Engineering
Prof. dr. K. Binnemans
Prof. dr. ir. S. Kuhn

Prof. dr. M. Koper


(Leiden University)
Prof. dr. ir. J. Ustarroz Troyano
(Université libre de Bruxelles)

May 2020
© 2020 KU Leuven – Faculty of Engineering Science
Uitgegeven in eigen beheer, Sam Eggermont, Kasteelpark Arenberg 44 - box 2450, B-3001 Leuven (Belgium)

Alle rechten voorbehouden. Niets uit deze uitgave mag worden vermenigvuldigd en/of openbaar gemaakt worden
door middel van druk, fotokopie, microfilm, elektronisch of op welke andere wijze ook zonder voorafgaande
schriftelijke toestemming van de uitgever.

All rights reserved. No part of the publication may be reproduced in any form by print, photoprint, microfilm,
electronic or any other means without written permission from the publisher.
Preface

The accomplishments described in this PhD thesis were not mine alone. Here, I
would like to take the opportunity to acknowledge the people that guided and
helped me throughout the last 5 years.
I thank Jan Fransaer and Xochitl Dominguez Benetton, for their financial
support and mentoring. Endless discussions about (electro)chemistry, and
diving deep into the mechanism of GDEx, helped to make this PhD thesis a
success. Rafael Prato, for being my closest colleague, although we were miles
apart thanks to VITO’s extremely strategic location. You were my partner-
in-crime while opening Pandora’s box of the GDEx process. Talking to you
about our common struggles in understanding GDEx always brought us closer
to the answers. Guillermo Pozo and Omar Martinez, for being my other team
members in the GDEx squad. Our monthly meetings would not have been the
same without you.
I would like to thank the members of the examination committee for investing
time and effort to read and comment on my manuscript. I hope you are satisfied
with how I addressed your remarks in this final version.
Special thanks to all of the administrative and technical support personnel at
MTM, in particular, Huberte Cloosen for her endless enthusiasm and patience,
and for being the rock-solid foundation of our department. Tom Van der
Donck, for the many excellent SEM photographs and for training me in the skill
of investigating the microscopic world. Joop van Deursen deserves a special
acknowledgment for assisting me with the many technical practicalities that I
struggled with, as well as for building my electrochemical flow-cell reactor from
scratch using the designs that came from hours of interesting discussions.
I would like to pay tribute to all of my colleagues, friends, and family that were
there when I needed support. Gijs and Nicolò, for being there in the beginning
and for becoming friends rather than colleagues over the years. Isadora, for
being my PhD-mate, going through the same struggles at the same time, and

i
ii PREFACE

helping one another when all hope seemed lost. Of course, I cannot thank
her enough for our coffee moments. Liuba, Basit, Xuan, Kai, Pieter, Palan,
Samy, Joao, among others, for being part of my journey and lighting up my
days. Senne, for cooperating along the way, and especially at the end, when you
were very busy yourself, finishing your PhD thesis. You inspired me to remain
interested in the mathematics that were necessary in the search for answers to
questions of the GDEx process.
MTM would not be the same without its yearly departmental trip. The
Departmental Trip Organizing Committee (DTOC) was by far my most favorite
committee to have been part of during these five years. I thank all of the current
and previous members to help make the departmental trip a yearly success.
I would like to thank my West-Flemish friends that, at some point or another,
co-migrated to Leuven: Mathieu, Karel, Lukas, Thomas, and Xander. It is an
absolute privilege to call them (and their partners) my friends. They made
my years in Leuven both unforgettable and sometimes somewhat forgettable.
Whatever the future brings, I will always be grateful for those moments. I will
not forget.
I thank my mom, my dad, Tom and Annabel, and Lies, for making Annabelle
and me feel at home after our (re)migration to Bruges, as well as for any other
support during the last five years and before.
Most importantly I would like to thank my fiancee Annabelle. She was at
my side during the entire journey. Being young (and somewhat lost) together
has been extremely interesting, educational and inspiring. She helped me to
rediscover the importance of family and taught me so many things about myself.
I love her with all my heart, and I am grateful for everything we have been able
to share.
Thank you.
Sincerely,
Sam
PREFACE iii

Acknowledgment

This research was funded by the Strategic Initiative Materials (SIM) in Flanders,
within the recyclable materials MaRes program, under grant agreement n◦
150626 (Get-A-Met project). Jan Fransaer and Xochitl Dominguez-Benetton
acknowledge funding from the European Union’s Horizon 2020 research and
innovation program under grant agreement n◦ 654100 (CHPM2030 project).
Abstract

The purpose of this work was to discover the fundamental aspects of gas-diffusion
electrocrystallization (GDEx). GDEx is defined as the reactive precipitation
of intermediaries issued from an electrochemical reaction at a gas-diffusion
electrode (GDE), with metal precursors provided in the electrolyte. It can be
used as a synthesis method to prepare (nano)particles or as a metal removal
and recovery process. An exploratory study of the GDEx process was carried
out to demonstrate the similarities and differences of GDEx with alkaline
titration, a well-known precipitation method. The motivation to investigate
GDEx came from preliminary experiments at VITO which demonstrated the
potential of GDEx for producing crystalline metal oxide (nano)particles from
metal-containing waste streams, even for low metal ion concentrations. The
combination of synthesis and metal recovery is desirable because society relies
increasingly on the development of new functional materials preferably from
recyclable sources. In light of this, this work focused on understanding the GDEx
process with regards to how metal ions are recovered, and how and why they
precipitate as nanoparticles. The focus was on air-fed GDEs for the formation
of metal (oxy)(hydr)oxide nanoparticles. In summary, the key contributions of
this PhD were: (1) demonstrating that although it is an electrochemical process
and therefore inherently heterogeneous, GDEx can have the appearance of a
homogeneous process, (2) the comparison of GDEx to alkaline titration, (3)
the discovery of oxidation-assisted alkaline precipitation and its role in GDEx
as a process to synthesize metal (oxy)(hydr)oxide nanoparticles, and (4) the
investigation of the accumulation of species in the porous GDE, a phenomenon
observable during the GDEx process in batch mode. Future work includes the
development of an all-encompassing model to reach the goal of GDEx as a
process for synthesis-by-design.

v
Beknopte samenvatting

Dit werk poogt de fundamentele aspecten van elektroprecipitatie in het algemeen


en meer specifiek van gas-diffusie-elektrokristallisatie (GDEx) te begrijpen.
GDEx is een elektrochemische methode die een gas-gevoede gas-diffusie-elektrode
(GDE) gebruikt om precipitanten te produceren voor de reactie met metaalionen
en de daaropvolgende vorming van metaalhoudende-nanodeeltjes. Om dit doel
te bereiken, werd er een uitgebreide studie gedaan van het GDEx-proces om
overeenkomsten en verschillen met base-titratie, een bekende nanodeeltjes-
synthesemethode, aan te tonen. De motivatie om GDEx te onderzoeken
kwam van verkennende experimenten bij VITO, die aantoonden dat GDEx
in staat was om snel kristallijne metaaloxide-nanodedeeltjes te produceren
uit metaalhoudende afvalstromen, zelfs bij lage metaalionconcentraties. De
combinatie van synthese en metaalterugwinning is wenselijk gezien het feit dat
de samenleving in toenemende mate afhankelijk wordt van de ontwikkeling
van nieuwe functionele materialen die bij voorkeur gewonnen worden uit
recycleerbare bronnen. Om hier aan bij te dragen, lag de focus van dit werk op
het begrijpen van het GDEx-proces met betrekking tot hoe metaalionen worden
teruggewonnen en hoe en waarom ze neerslaan als nanodeeltjes. Bovendien werd
gefocust op de lucht-gevoede GDE voor het vormen van metaal(oxy)(hydr)oxide-
nanodeeltjes. Samengevat waren de belangrijkste bijdragen van dit doctoraat:
(1) het aantonen dat GDEx, als elektrochemisch - en dus inherent heterogeen
proces, ogenschijnlijk uiterlijke kenmerken van een homogeen proces vertoont,
(2) de vergelijking van GDEx met base-titratie, (3) de ontdekking van oxidatie-
geassisteerde base-precipitatie en de rol ervan in GDEx als een proces om
zeer kleine metaal(oxy)(hydr)oxide-nanodeeltjes te synthetiseren, en (4) het
onderzoek naar de accumulatie van ionen in de poriën van de GDE, iets dat
werd waargenomen tijdens het GDEx-proces in batchmodus. Toekomstig werk
omvat de ontwikkeling van een allesomvattend model met als doel van GDEx
een proces voor synthese-naar-ontwerp te maken.

vii
Abbreviations

AM analytical model
AP alkaline precipitation
AT (traditional) alkaline titration
CSTER continuously stirred tank electrochemical reactor
CSTR continuously stirred tank reactor
Cu-ISE Cu2+ -ion selective electrode
CNT classical nucleation theory
DLS dynamic light scattering
GDE gas-diffusion electrode
GDEx gas-diffusion electrocrystallization
Get-A-Met Groundbreaking Extraction Technology for critical Metals
and metalloids from industrial wastewaters
(fundamental research project within the SIM MaRes Programme)
MaRes Materials from solid and liquid industrial process Residues
NM numerical model
NP nanoparticle
ORR oxygen reduction reaction
Ox-AP oxidation-assisted alkaline precipitation
PTC precipitation titration curve
PTFE polytetrafluoroethylene
VITO Vlaamse Instelling voor Technologisch Onderzoek
(Flemish Institute for Technological Research)
SEM scanning electron microscopy
SIM Strategic Initiative Materials
sp ICP-MS single-particle inductively coupled plasma mass spectrometry
VLAIO Vlaams Agentschap Innoveren en Ondernemen
XRD X-ray diffraction

ix
List of symbols

Latin letters

ai [-] activity of species i


Aj [m2 ] surface area with specifier j
cji [mol m−3 ] concentration of species i in
a specific domain j (e.g., diffusion layer or bulk)
Di [m2 s−1 ] diffusion coefficient of species i
Di [m2 s−1 ] effective diffusion coefficient of species i
F 96,485 C mol−1 Faraday’s constant
Iapp [A] applied current
ki [m s−1 ] mass-transfer rate of species i
Ki [molj m−3j ] chemical equilibrium constant of species i
with j dependent on stoichiometry of involved species
n [-] number of electrons transferred in a given
electrochemical reaction
Nji , Nij [mol m−2 s−1 ] flux of species i with specifier j (i.e., in or out)
vector and component
Pi [mol s−1 ] production rate of species i
Qi [molj m−3j ] chemical reaction quotient of species i
with j dependent on stoichiometry of involved species
rV [m3 s−1 ] volumetric addition rate
(titration rate)
Ri [mol s−1 ] reaction rate of species i
Si [-] supersaturation of species i
Shavg [-] average Sherwood number in the flow-cell
Shi [-] Sherwood-like number of species i (kwGDE /Di )
(expresses contribution of convection on
the accumulation of species in the GDE)

xi
xii LIST OF SYMBOLS

t [s] time
Vj [m3 ] volume with specifier j
wGDE [m] penetration depth of the GDE
x [m] x-coordinate

Greek letters
α [-] fitting model correction factor
γi [-] activity coefficient of species i
δ [m] diffusion layer thickness
 [-] hydrophilic porosity of the GDE
ηi [-] production efficiency of species i
ψ [-] porous media coverage of the GDE
τi [s] characteristic diffusion time in the GDE of species i

Subscripts and superscripts

0 initial
app applied
b bulk
dl diffusion layer
f final
GDE gas-diffusion electrode
geo geometric
het heterogeneous
hom homogeneous
i of species i
phil hydrophilic porosity
phob hydrophobic porosity
q with regards to flow rate q
ss steady-state
tot total
Contents

Abstract v

Beknopte samenvatting vii

Abbreviations ix

List of symbols xi

Contents xiii

1 Introduction 1
1 Problem statement . . . . . . . . . . . . . . . . . . . . . . . . . 2
2 Aim . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
3 Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

2 General introduction to the GDEx process 7


1 Electrochemical synthesis of
metal (oxy)(hydr)oxides . . . . . . . . . . . . . . . . . . . . . . 8
1.1 Precipitation of metal (oxy)(hydr)oxides . . . . . . . . . 8
1.2 Electrogeneration of a base . . . . . . . . . . . . . . . . 10
2 Gas-diffusion electrodes and the oxygen reduction reactions . . 11
2.1 Internal structure of a gas-diffusion electrode . . . . . . 11

xiii
xiv CONTENTS

2.2 Phenomena at gas-diffusion electrodes . . . . . . . . . . 12


2.3 Electrochemical reactions with gases . . . . . . . . . . . 13
3 Reactive precipitation: combining metal removal and nanoparti-
cle synthesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

3 Exploratory study of the GDEx process 17


1 GDEx process, AT process and key process parameters . . . . . 17
2 Homogeneous appearance of the GDEx process . . . . . . . . . 23
3 GDEx and the oxygen reduction reaction: production of H2 O2
and high solution redox potentials . . . . . . . . . . . . . . . . 33
4 Universality and selectivity of AT and GDEx processes . . . . . 36
5 Energy considerations of GDEx and effect of supporting electrolyte 40
6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

4 Metal removal from aqueous solutions: insights from modeling


precipitation titration curves 45
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
2 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3 Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.1 Basic equations . . . . . . . . . . . . . . . . . . . . . . . 49
3.2 Analytical model . . . . . . . . . . . . . . . . . . . . . . 53
3.3 Numerical model . . . . . . . . . . . . . . . . . . . . . . 56
4 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . 57
4.1 Taxonomy of the models . . . . . . . . . . . . . . . . . . 57
4.2 Effect of C HCl , C MCln , n, and KM(OH)n . . . . . . . . . 58
4.3 Effect of γi , C NaCl , and multi-species ions . . . . . . . . 60
4.4 Comparison of the numerical model with the experiments 61
5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
CONTENTS xv

5 Oxidation-assisted alkaline precipitation: effect of H2 O2 on the


size of CuO and FeOOH nanoparticles 67
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
2 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
2.1 Electrolyte synthesis . . . . . . . . . . . . . . . . . . . . 72
2.2 Preparation of nanoparticles . . . . . . . . . . . . . . . . 73
2.3 Post-treatment and characterization . . . . . . . . . . . 74
3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
3.1 Visual observations of nanoparticle suspension . . . . . 75
3.2 Size of the nanoparticles . . . . . . . . . . . . . . . . . . 76
3.3 Composition of the nanoparticles . . . . . . . . . . . . . 78
3.4 Morphology of the nanoparticles . . . . . . . . . . . . . 79
4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85

6 Oxidation-assisted alkaline precipitation of nanoparticles using gas-


diffusion electrodes 87
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
2 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
2.1 Electrolyte solutions . . . . . . . . . . . . . . . . . . . . 93
2.2 GDEx flow-cell and GDEx process . . . . . . . . . . . . 93
2.3 AT process . . . . . . . . . . . . . . . . . . . . . . . . . 94
2.4 Analytical measurements . . . . . . . . . . . . . . . . . 95
2.5 Post-process treatment and characterization . . . . . . . 95
3 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . 96
3.1 Colorimetric transitions . . . . . . . . . . . . . . . . . . 96
3.2 Production of H2 O2 . . . . . . . . . . . . . . . . . . . . 97
3.3 Dynamic pH behavior . . . . . . . . . . . . . . . . . . . 98
xvi CONTENTS

3.4 Composition of the precipitates . . . . . . . . . . . . . . 100


3.5 Size and morphology of the precipitates . . . . . . . . . 101
4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104

7 Accumulation of ORR reaction products in porous gas-diffusion


electrodes 105
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
2 Modeling the GDE in one spatial dimension . . . . . . . . . . . 109
2.1 Analytical steady-state model of concentrations in the GDE114
2.2 Analytical time-dependent model of the total accumula-
tion in the GDE . . . . . . . . . . . . . . . . . . . . . . 117
3 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
3.1 Electrolyte solutions . . . . . . . . . . . . . . . . . . . . 119
3.2 Gas-diffusion electrode . . . . . . . . . . . . . . . . . . . 120
3.3 Flow-cell reactor and process conditions . . . . . . . . . 120
3.4 Determining the HO-2 production efficiency . . . . . . . 122
3.5 Determining the electrolyte penetration depth . . . . . . 123
3.6 Determining the value of the mass-transfer coefficients . 123
3.7 Determining the accumulation of OH- and HO-2 in the
GDE and their effective diffusion coefficients . . . . . . 124
4 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . 125
4.1 Hydrogen peroxide efficiency . . . . . . . . . . . . . . . 126
4.2 Electrolyte penetration depth . . . . . . . . . . . . . . . 126
4.3 Mass-transfer coefficient . . . . . . . . . . . . . . . . . . 128
4.4 Accumulation of OH− and HO−
2 in the porous GDE . . 128
5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136

8 General conclusions 137


CONTENTS xvii

A Measuring the copper concentration with a Cu-ISE 143

B PHREEQC input file for modeling PTCs of Zn(OH)2 145

C Sample treatment for and characterization with sp ICP-MS 149

D Calculating supersaturations 151

E Derivation of the analytical time-dependent model of the total


accumulation in the GDE 153

F Long-time experiment of the bulk proton concentration 157

Bibliography 159

Curriculum Vitae 177

List of publications 179


Chapter 1

Introduction

This PhD thesis is the result of a close collaboration between KU Leuven and
VITO within the framework of the Get-A-Met project (funded by SIM, VLAIO,
within the MaRes program). The goal of this PhD was to obtain a fundamental
understanding of Gas-Diffusion Electrocrystallization (GDEx), as a new process
consisting of the electrochemically-driven reactive precipitation of metals or
metalloids. The PhD thesis aims to demonstrate the results of the investigation
of GDEx and to provide the reader with a better understanding of the process
in general, and more specifically of the fundamental principles that guide it.
This introductory chapter defines the research questions and how the problem
statement of this PhD thesis originated (section 1), it states the specific aims of
the PhD within the framework of the Get-A-Met Project (section 2), and the
outline section clarifies the structure of the PhD thesis (section 3).

1
2 INTRODUCTION

1 Problem statement

During preliminary research at VITO, GDEx demonstrated great potential


towards the cost-effective and energy-efficient recovery of metal ions from low
and medium concentrated streams. Additionally, GDEx demonstrated the
synthesis of functional materials (e.g., nanoparticles). One of the first proven
examples was the removal of Ce3+ from a synthetic solution. Simultaneously,
depending on the process conditions, crystalline cerium oxide nanoparticles with
different oxygen stoichiometry were synthesized (i.e., in between Ce2 O3 and
CeO2 ). In this example, GDEx of cerium was carried out in the catholyte side of
a two-compartment electrochemical flow-cell reactor with complete recirculation
to a reservoir.
Several observations sparked the interest for fundamental research of GDEx.
The first observation was the fast removal of metal ions, considered unusual
for the size of the reactor and the surface area of the electrode. The second
observation was the fact that cerium (III) oxidized to cerium (IV), even though
the process was carried out at the catholyte side of the reactor, the side of the
reactor that would imply a reducing, rather than an oxidizing, environment.
The third observation was the fact that suspensions of crystalline metal oxide
nanoparticles of CeO2 could be synthesized whilst recovering more than 99% of
the cerium ions from solution, while instead precipitation at the surface of the
electrode would be expected. The last observation was the fact that by changing
the concentration of the cerium (III) precursor salt, nanoparticles with a large
range of different properties (oxidation state, crystallite size, lattice parameter,
oxygen sorption capacity) were synthesized. These observations implied the
ability to recover the metal ions as functional materials, providing interesting
opportunities for GDEx as a synthesis method.
These observations led to several interesting research questions, some of which
are addressed in this work:

1. What causes the fast removal of metal ions from a solution?


2. How can metal ions be oxidized in the catholyte, a supposedly reducing
environment?
3. What are the conditions that allow the formation of precipitates in general,
and crystalline nanoparticles more specifically?
4. How can the characteristics of these nanoparticles be controlled?
5. How does the power consumption of GDEx compare to other processes?
6. How scalable is the GDEx process?
PROBLEM STATEMENT 3

Besides these questions, the above-mentioned observations arose several


expectations of the GDEx process:

1. as a method with possibly lower power consumption than other synthesis


and metal removal processes;
2. as a method with a high amount of sustainable features such as producing
clean water, removal of polluting metals, operability at room temperature,
among others;

3. as a method with an equal or better performance compared to other


synthesis methods and metal removal processes;
4. as a method with a high potential for industrial upscaling.

Within the framework of the Get-A-Met project, this PhD thesis focused more
on the fundamental aspects of the process. Therefore, the actual essay of these
expectations with the current best-practices was excluded from the scope of
this work.
4 INTRODUCTION

2 Aim

The Get-A-Met project aimed to increase the understanding of the GDEx


process up to the point that its industrial relevance for metal removal from
liquid waste streams and their valorization into functional products could be
confidently demonstrated. This PhD thesis aimed to contribute to the Get-
A-Met project by answering the fundamental questions posed in the problem
statement. The specific objectives of this PhD to answer to these questions
were to:

1. elucidate the nature (homogeneous or heterogeneous) of the different


reactions of the GDEx process;
2. identify the homogeneous and/or heterogeneous reactions of the GDEx
process;

3. discover how selective and how universal the GDEx process is;
4. identify the role of the GDEx process regarding the oxidation of the metal
ions in the catholyte;
5. develop mathematical models that can help to understand specific sub-
processes of the GDEx process or the GDEx process as a whole;

6. map the effect of the different process conditions on the characteristics of


the synthesized nanoparticles.

Answers to all of these questions, and more, were sought during work done for
this PhD. The scope of this PhD was limited to metal removal and nanoparticle
formation in aqueous solutions. Concerning the formation of nanoparticles, the
focus was on metal (oxy)(hydr)oxide nanoparticles. Moreover, this work only
considers the GDEx process conducted with the GDE as an air-fed cathode in
aqueous electrolytes, hereinafter simply referred to as the GDEx process.
OUTLINE 5

3 Outline

This PhD thesis consists of eight chapters. After this first chapter, Chapter 2
contains a literature review relevant to the study of GDEx and situates GDEx
among the relevant metal removal processes and nanoparticle synthesis methods.
Chapter 3 provides an exploratory study of the GDEx process and the different
key parameters. In several instances, GDEx is assessed via a comparison with
traditional alkaline titration (AT), a method known for being used as a metal
removal process and a nanoparticle synthesis method. The oxygen reduction
reaction (ORR) and the production of H2 O2 are discussed and their effect
on the redox potential of the catholyte solution. In addition, it demonstrates
the potential and limitations of the GDEx process in terms of selectivity and
universality. It also elaborates on the power consumption of GDEx.
Chapter 4 contains the comparison of modeled and experimental titration
curves for the traditional alkaline precipitation of Zn(OH)2 . The modeled and
experimental titration curves show good agreement and demonstrate that AT
of ZnCl2 solutions for AP of Zn(OH)2 can serve as a good benchmark for other
metal removal or recovery processes, including GDEx. Specifically in terms of
the pH evolution during the precipitation process.
Chapter 5 contains the experimental results of controlled experiments to
demonstrate the role of H2 O2 in the formation process of metal (oxy)(hydr)oxide
nanoparticles and the effect of H2 O2 on the size of these nanoparticles. It
introduces the concept of oxidation-assisted alkaline precipitation (Ox-AP).
Chapter 6 compares experimental results for the AP process and the GDEx
process of solutions with decreasing oxidation potential (i.e., ZnCl2 , MnCl2 , and
FeCl2 ). It confirms the initial discrepancy of AT and GDEx introduced in the
second chapter. It also confirms AT of ZnCl2 as a good reference process and
how solutions of MnCl2 and FeCl2 deviate from this. In addition, it provides a
practical example of Ox-AP, because GDEx of FeCl2 produces very small Fe3 O4
nanoparticles due to GDEx-elicited oxidation-assisted alkaline precipitation, in
contrast to the much larger FeO nanoparticles produced by AT-elicited alkaline
precipitation.
Chapter 7 provides an answer to why a discrepancy is observed between AT and
GDEx. It contains an analytical model that demonstrates the accumulation
of OH− and HO− 2 within the pores of the GDE and connects these results to
experimental data with the GDE electrochemical flow-cell reactor.
Chapter 8 contains the general conclusion of this work. It also elaborates on
open questions and opportunities for future work.
Chapter 2

General introduction to the


GDEx process

Gas-diffusion electrocrystallization (GDEx) can be defined as the reactive


precipitation of intermediaries issued from an electrochemical reaction at a
gas-diffusion electrode (GDE), with metal precursors present in the electrolyte.
It can be used as a synthesis method to prepare (nano)particles or as a metal
recovery technology. The holy grail for the GDEx process would be to produce
functional materials (e.g., functional nanoparticles) from wastewaters (e.g.,
leachates, extractants, solvents).
GDEx involves different types of reactions: surface reactions (i.e., at the GDE)
and bulk reactions (i.e., in solution). GDEx is assumed to be possible in aqueous
and non-aqueous solutions. The research field of GDEx touches the research
fields of gas-diffusion electrochemistry, electrochemical engineering in flow-cell
reactors, metal removal, and nanoparticle synthesis. In summary, GDEx is a
relatively complex process, combining several scientific research fields, with the
potential to be broadly applicable. Any of the fields connected to GDEx would
deserve a proper review, and even then, the scope would often be too broadly
defined and would need to be specified to remain containable.
On gas-diffusion electrocrystallization, as it is defined in this thesis, very little
literature was found before the Get-A-Met project [1–3]. However, there are
numerous related methods and other subjects that hold relevance for this
topic; hence, this literature review contains an introduction to the more
general concept of electrochemical synthesis of metal (oxy)(hydr)oxides, the
precipitation of metal (oxy)(hydr)oxides, and gas-diffusion electrochemistry. It

7
8 GENERAL INTRODUCTION TO THE GDEX PROCESS

also shortly situates GDEx, as a method that combines metal removal with
nanoparticle synthesis, among other comparable methods. To avoid repetition,
several paragraphs of this chapter will refer to later chapters which discuss the
introduced concepts in more detail, related to the (experimental) research there
discussed.

1 Electrochemical synthesis of
metal (oxy)(hydr)oxides

A U.S. patent from 1904 by Hunt [4] was the oldest record on electrochemical
synthesis of metal (oxy)(hydr)oxides found in this literature investigation. Since
then, a large body of scientific literature on the electrochemical synthesis of
metal (oxy)(hydr)oxides has developed, most of it at least partially rely on the
cathodic electrogeneration of a base [5–10].
This scientific literature study does not pretend to be a complete review of the
topic. As a rule, the cited references in the following paragraphs are the ones
that were decided to be the most relevant or complete sources for the reader
to further investigate the topic of interest. The study is limited to aqueous
solutions, although metal oxides can be produced in non-aqueous solutions as
well [11–17].
In this section, the precipitation mechanisms of metal (oxy)(hydr)oxides are
briefly discussed, followed by a discussion on electrochemical synthesis of metal
(oxy)(hydr)oxides, relying on the cathodic electrogeneration of base.

1.1 Precipitation of metal (oxy)(hydr)oxides

In this work, precipitation is defined as the sedimentation of a solid material (a


precipitate) from a liquid solution in which the material is present in amounts
greater than its solubility in the liquid [18]. According to this definition,
precipitation is, in essence, a separation of phases (i.e., the solid separates
from the liquid). This phase separation happens because the solubility limit of
one phase (e.g., Zn(OH)2 ) in the other (e.g., H2 O) is exceeded. Solubility is
not a material constant and strongly depends on several environmental factors
(e.g., temperature, pressure, solvent, composition of the solution, etc.).
The phase separation or precipitation occurs in four phases: (1) formation of a
zero-charge precursor, (2) creation of nuclei, (3) growth of the nuclei, and (4)
aging of the particles in suspension [19]. These mechanisms are discussed in
ELECTROCHEMICAL SYNTHESIS OF
METAL (OXY)(HYDR)OXIDES 9

more depth in Chapter 5 (p. 81) with appropriate examples. Especially the
book of Jolivet et al. [19] deserves to be mentioned, for discussing the formation
of metal (oxy)(hydr)oxides in depth and width. In addition, Chapter 4 discusses
the literature and results on how to follow the chemical formation of sparingly
soluble Zn(OH)2 (phase 1), by modeling or measuring the pH.
Especially, diagrams representing electrochemical equilibria in aqueous solutions
(e.g., Pourbaix diagrams, E-pH diagrams) are useful tools in aqueous
(electro)chemistry, to determine the thermodynamic feasibility of certain phases
[20]. They indicate, for certain values of the redox potential and the pH, which
are the stable phases, and referring to a 1 M concentration of the targeted
element they are referred to as Pourbaix diagrams [20]; however, they can
be calculated for any pressure and temperature. Figure 2.1 demonstrates
a simplified E-pH diagram for a system with 0.001 M Zn and water, at 1
atmosphere and 25 ◦ C. More sophisticated diagrams can be calculated from
thermodynamic data, containing zones of multi-species ions and taking into the
anion and supporting electrolyte. The dominating phase of a species depends
on the pH and redox potential and can vary by sufficiently changing either
parameter.

-2 0 2 4 6 8 10 12 14 16
2.2 2.2
1.8 1.8
1.4 1.4
1 1
E vs. SHE [V]

0.6 0.6
0.2 Zn2+ 0.2
Zn(OH)2
-0.2 -0.2
-0.6 -0.6
-1 -1
-1.4 Zn -1.4
-1.8 -1.8
-2 0 2 4 6 8 10 12 14 16
pH

Figure 2.1: Simplified Pourbaix diagram of a system with 0.001 M Zn and


water at 1 atmosphere and 25 ◦ C [20].
10 GENERAL INTRODUCTION TO THE GDEX PROCESS

1.2 Electrogeneration of a base

To go from a metal ion in solution to a metal (oxy)(hydr)oxide precipitate,


the metal ion needs to react with species containing oxygen (and hydrogen).
The first species that come to mind are H2 O, OH− , and O− 2 (short lifetime in
aqueous solutions below pH 14 [21]). Except for the anodic formation of metal
oxides, most of the scientific literature on the electrochemical synthesis of metal
(oxy)(hydr)oxides relates, in one way or another, to the electrogeneration of
base (OH− ) at the cathode.
Although several review papers [5–10] try to cover the general topic of
electrochemical synthesis of metal (oxy)(hydr)oxides, none of them cover the
complete scientific literature and it is often more useful to consult review papers
that focus on one specific (set of) metal (oxy)(hydr)oxide(s) of interest, for
example, iron oxide [22] or zinc oxide [23]. The fact that synthesis methods are
often not directly clear from article titles or abstracts, obliging the in-depth
investigation of the materials and methods section and even supplementary
information files of any given publication, might contribute to why such general
review papers are scarce.
Summarizing the investigated literature, the electrochemical synthesis of metal
(oxy)(hydr)oxides can be subdivided in several different ways: carried out in a
one-compartment [2–4, 24–34] or in two-compartment electrochemical cell [1, 35–
42], using metal salts as the necessary metal ion precursor [3,26,28,30–34,36,38–
40] or using anodic dissolution to provide the necessary metal ion precursor [1,2,4,
24,25,27,29,35,37,39,41,42], and at the surface (electrodeposition) [3,4,26,28–34]
or in the bulk solution [1, 2, 24, 25, 27, 35–42]. The majority of the investigated
literature on the electrochemical synthesis of metal (oxy)(hydr)oxides used the
hydrogen evolution or nitrate reduction reactions for the electrogeneration of a
base [5]:

2H2 O + 2e−
H2 + 2OH− (E◦ = −0.828 V) (2.1)

NO− − −
3 + H2 O + 2e
NO2 + 2OH

(E◦ = 0.01 V) (2.2)

Except for one 1978 U.S. patent [1], one 2015 U.S. patent [2] and a recent article
(2017, published after the initial literature investigation for this thesis) [3], no
other literature was discovered where an air-cathode (gas-diffusion electrode) is
suggested to carry out oxygen reduction reactions (see equations (7.3), (7.4),
(7.1) and (7.2) p. 108) for the electrogeneration of a base to synthesize metal
(oxy)(hydr)oxides. Although the methods used in these publications share similar
characteristics with the GDEx process (i.e., use of a gas-diffusion electrode
passing air, precipitating metal (oxy)(hydr)oxides), they are not carried out
GAS-DIFFUSION ELECTRODES AND THE OXYGEN REDUCTION REACTIONS 11

in a two-compartment cell [2, 3], or are carried out at much lower current
densities [3], or are using a sacrificial anode [1, 2]. Moreover, none of them
discuss the cogeneration of hydrogen peroxide during the process. For the cases
using sacrificial anodes [1, 2], the steady-state concentration of the metal ions
in solution was not specifically mentioned. However, this is expected to be
low, because the amount of metal from the anode that is dissolving as metal
ions would be countered with a corresponding amount of hydroxyl ions. The
resulting precipitation reaction would keep the overall metal ion concentration
low. In the other case [3] the concentration was in the 5 mM to 10 mM range.

2 Gas-diffusion electrodes and the oxygen reduc-


tion reactions

In this work, an air-fed gas-diffusion electrode (i.e., air cathode, air-depolarizing


cathode, etc.) was used for the reduction of oxygen, mainly to OH− , HO− 2
or its protonated form H2 O2 . Gas-diffusion electrodes are used for many
applications such as fuel cells, metal-air batteries, and chlor-alkali electrolysis,
among others [43]. Gas-diffusion electrodes are designed to provide a large
interface between solid, liquid, and gas (i.e., the three-phase boundary or
triple-phase boundary) [43]. One of the first applications using a gas-diffusion
electrode was the zinc-air battery, which was first introduced around 1878 [44]
as an alternative to the Leclanché battery [45]. GDEs are also used in fuel cell
technology. Several excellent review papers exist on the current state-of-the-art
gas-diffusion electrodes for AFCs [46] and PEMFC [47–49].

2.1 Internal structure of a gas-diffusion electrode

The gas-diffusion electrodes in the first commercially available zinc-air batteries


of the 1930s were porous carbon air cathodes impregnated with wax to avoid
flooding of the pores [45, 50]. The invention of PTFE in the late 1930s [51] and
its introduction in porous electrodes as hydrophobic agent [52, 53] were the first
steps toward current gas-diffusion electrodes.
The multi-layered cold-rolled PTFE-bound VITO CoRE ®gas-diffusion
electrode originates from a design with fuel cell applications in mind [54].
The VITO CoRE GDE resembles designs of GDEs for alkaline fuel cells (AFCs)
rather than proton-exchange membrane fuel cells (PEMFCs). Different GDE

®
designs are possible and the design depends on how single fuel cells are stacked
together (e.g., monopolar [55, 56] or bipolar [57]). The VITO CoRE GDE
12 GENERAL INTRODUCTION TO THE GDEX PROCESS

design will be discussed and most closely resembles the designs of monopolar
AFC GDEs.
®
In the VITO CoRE GDE, the three-phase boundary between electrode,
electrolyte, and gas is provided using a multi-layer design. The first layer (1) is
a porous hydrophobic layer. This gas-diffusion layer (GDL), forms the barrier
between the gas compartment and the electrolyte compartment. It consists
of porous PTFE and allows gas diffusion but resists liquid water to pass to
the gas compartment. The second layer (2) is partially hydrophobic, partially
hydrophilic. This carbon/PTFE layer (C/PTFE), is a layer made of a mixture
of hydrophilic porous activated carbon particles and hydrophobic PTFE, and a
catalyst (optional). It is a hybrid material consisting of two networks, with a
very large interface: a network of hydrophobic PTFE channels and hydrophilic
porous carbon particles. Alvarez-Gallego et al. determined the overall porosity
of an 80/20 C/PTFE layer to be 67 %, and the hydrophilic and hydrophobic
porosity to be 29 % and 38 %, respectively [58]. The intrinsic BET surface area
of the active carbon is 745 m2 g−1 , however, the overall BET surface area in the
80/20 C/PTFE layer was determined to be 454 m2 g−1 [58]. The porous carbon
network is penetrated with electrolyte. The three-phase boundary emerges
at the interface with the hydrophobic PTFE channels and the porous carbon
network. The carbon network is provided with electrons via a metal-mesh
current collector (3), in this case, 316L stainless steel. Figure 2.2 shows a
schematic of a GDE and the hydrophilic and hydrophobic networks.

2.2 Phenomena at gas-diffusion electrodes

A gas-diffusion electrode is a complex system in which many different phenomena


occur. These are summarized in figure 2.3. The electrolyte will partially wet
the GDE to establish a three-phase boundary. The electrochemical reactions
(and side reactions) in the GDE happen at this three-phase boundary of the
solid phase (e.g., electron-conducting carbon particles), the liquid phase (e.g.,
ion-conducting aqueous NaCl solution), and the gas phase, carrying an oxidizing
gas of interest (e.g., air). In the gas phase, the important transport phenomena
are diffusion and convection, although the gas phase in the pores is subject to
little convection. In the liquid electrolyte phase, migration is added to diffusion
and convection, although the electrolyte phase in the pores is subject to little
convection [59]. Across the boundary of both phases, dissolution (e.g., of O2 in
the electrolyte) and evaporation (e.g., of H2 O into the gas phase) can occur.
GAS-DIFFUSION ELECTRODES AND THE OXYGEN REDUCTION REACTIONS 13

gas
electrolyte
porous carbon particle network
binder network
hydrophobic GDL
active layer
current collector
electrolyte
gas

Figure 2.2: Schematic of GDE with inset showing hydrophilic (electrolyte) and
hydrophobic (gas) channels.

2.3 Electrochemical reactions with gases

2.3.1 Oxygen reduction reactions

Although different gases can be used to feed the GDEx process, in this thesis,
especially the oxygen reduction reactions were of interest. The low availability
of oxygen at the electrode surface, as a result of the low oxygen solubility
in water, is often a limiting factor in the electrochemistry of oxygen [60–62].
However, this limitation is resolved by the design of the internal structure of
the GDE (section 2.1). A large amount of oxygen reduction reactions exist and
are discussed in detail in several reviews and books [63–65]. In general, the
two overall pathways resulting in the most stable products are the four-electron
pathway and the two-electron pathway (written for alkaline media) [63]:

O2 + 2H2 O + 4e−
4OH− (E◦ = 0.401 V) (2.3)

O2 + H2 O + 2e−
HO−
2 + OH

(E◦ = −0.065 V) (2.4)

The oxygen reduction reactions are useful for the electrogeneration of a base.
Moreover, the two-electron pathway yields hydrogen peroxide (deprotonated
14 GENERAL INTRODUCTION TO THE GDEX PROCESS

Figure 2.3: Schematic of possible phenomena in gas-diffusion electrodes [43].


Reprinted with permission of Wiley.

in alkaline media). The key to answer one of the research questions related to
the oxidation of metal ions in the cathode lies in the formation of hydrogen
peroxide in the GDE as a result of the oxygen reduction reactions.

2.3.2 Other gas reactions

Besides the reduction of oxygen from pure oxygen, air or other mixtures, the
reduction of other gasses is widely researched. Especially N2 and CO2 deserve
special attention due to their potential to help reach some of the sustainable
development goals [66]. The electrochemical reduction of N2 to ammonia [67–72],
or the reduction of CO2 to urea in nitrate/nitrite solutions [73–79] could be the
key for sustainable production of artificial fertilizers for food production, while
the reduction of CO2 could help to reduce the concentration of this greenhouse
gas in the atmosphere and provide a solution for the sustainable production of
green carbon-based fuels [80–86].
REACTIVE PRECIPITATION: COMBINING METAL REMOVAL AND NANOPARTICLE SYNTHESIS 15

3 Reactive precipitation: combining metal removal


and nanoparticle synthesis

The preliminary experiments at VITO demonstrated GDEx as an electrochemi-


cal process combining metal removal and nanoparticle synthesis. To the best
of our knowledge, an electrochemical method to combine both metal removal
and nanoparticle synthesis is not discussed in the scientific literature. One
reason for this might be the conflicting interests of both processes. Metal
removal processes are preferably cheap and effective [87], while nanoparticle
synthesis methods aim to produce products, with specific characteristics, of
high value, at acceptable production rates [88]. Related to this, influents of
waste stream processes are often relatively low in metal ion concentration,
compared to influents for nanoparticle synthesis, for which the concentration is
chosen as high as possible (given the product requirements) to have the highest
production rate possible. Nevertheless, increasing awareness of environmental
concerns [89] and the fact that society relies increasingly on the development
of new functional materials [90], might increase the demand for nanoparticles
produced from wastewaters, promoting it in becoming an economically viable
process, regardless of its production rate.
The most current metal removal techniques are chemical precipitation, chemical
coagulation/flocculation, electrochemical methods (electro-Fenton, capacitive
deionization, electrocoagulation, etc.), ion-exchange, membrane filtration, bio-
remediation, and adsorption [91–96]. Of these metal removal processes, chemical
precipitation is the most widely used [97–99].
The most relevant nanoparticle synthesis techniques are physical vapor deposi-
tion, chemical vapor deposition, reactive precipitation, sol-gel, microemulsion,
sonochemical processing, and supercritical chemical processing [100–102].
Because of its convenience in processing, its low cost, and its potential for
mass-production, especially reactive precipitation is of high industrial interest
[100]. Although not among the most used techniques, several electrochemical
nanoparticle synthesis methods are also known. Only recently, a review on
the electrochemical synthesis of nanoparticles was published [103]. In general,
according to Therese et al., electrochemical synthesis can be subdivided into five
different techniques electromigration of reactant species, electrolysis of fused
salts, electrogeneration of base by cathodic reduction, anodic oxidation, and
alternate current synthesis [5]. In this work, GDEx is used with air-fed GDEs
in aqueous electrolytes and falls within the category of electrogeneration of base
by cathodic reduction.
Comparing the metal removal techniques with the nanoparticle synthesis
techniques, and taking into account that chemical precipitation is a form
16 GENERAL INTRODUCTION TO THE GDEX PROCESS

of reactive precipitation, it appears that only reactive precipitation overlaps


between the disciplines of metal removal and nanoparticle synthesis. GDEx, as
earlier defined, falls within the scope of reactive precipitation.
Before this PhD, GDEx had not yet been investigated in detail. Therefore, part
of this work consisted of benchmarking the GDEx process against a well-known
reactive precipitation process, namely alkaline titration (AT). AT was expected
to be a good benchmark because it is a more simple process. Because base is
added via titration, instead of via electrogeneration, a large portion of variables
is eliminated in AT (i.e., the electrolyte flow-rate, the flow cell size and shape,
the electrochemical reaction efficiency, etc.). With such a benchmark, it is easy
to identify deviations, thus identifying interesting starting points for further
investigation. The next chapter is dedicated to the in-depth study of several
aspects of the GDEx process.
Chapter 3

Exploratory study of the


GDEx process

This chapter is an exploratory study of the GDEx process. The most important
process characteristics of the GDEx process are introduced. We highlight the
key process parameters, and by exploring the process, research questions 1, 2,
and 5 posed in Chapter 1 are answered. After this chapter, the reader should
have a clear overview of what the GDEx process is, and why the following
chapters help to answer the other research questions. The first section discusses
the GDEx process, and how it relates to the alkaline titration process (AT)
and the key process parameters. The second section explores the contradictory
homogeneous appearance of the GDEx process. The third section shows how
GDEx, the oxygen reduction reaction, hydrogen peroxide, and the solution
redox potential are connected. The fourth section discusses the universality and
selectivity of GDEx as a metal removal/recovery process. In the last section,
the energy considerations of the GDEx process are investigated.

1 GDEx process, AT process and key process


parameters

For the remainder of this thesis, we redefine GDEx more specifically as the
reactive precipitation of metal ions in the catholyte with OH− , and possibly
H2 O2 or HO− 2 issued from the oxygen reduction reaction (ORR) at an air-fed
GDE. In contrast, AT can be defined as the reactive precipitation of metal

17
18 EXPLORATORY STUDY OF THE GDEX PROCESS

ions in the electrolyte with OH− issued from a titration with NaOH. Figure
3.1 demonstrates schematic drawings of the GDEx and AT processes for the
metal hydroxide precipitation case. It provides a rudimentary summary of the
expected precipitation mechanism for each process.

Gas-diffusion electrocrystallization (GDEx)


gas-diffusion layer M(OH)n*
active carbon layer M(OH)n* H2O
bulk precipitates
e-
H2O2 M(OH)
n*
OH - HO2-
gas channel

e-
OH- OH-
HO2-
HO2 - Mn+
O2 O HO2- H +
Mn+
- OH-
2 -
OH HO M n+
H +
O2 e- OH- OH-
2
O2 H + H+
H2O OH -
Mn+
n*>n
Alkaline Titration (AT)
M(OH)n H2O M(OH)n
titrator tip bulk H2O precipitates
M(OH)n

OH- + OH-
Na Mn+
Na+ H+ Mn+
Na+OH- OH- M H+
n+
OH- + Na+ H+
Na H+
OH- Mn+

Figure 3.1: Schematic drawings of metal hydroxide precipitation with the


GDEx and AT processes.

The GDEx process. The GDEx process could be summarized as follows:


during GDEx, air is fed at the back of a porous GDE; the oxygen reduces at
the triple-phase boundary between air, carbon, and electrolyte; this oxygen
reduction reaction (ORR) results in the production of OH− and HO− 2 ; these
species move into the solution and react with the available metal ions to form
sparingly soluble metal (oxy)(hydr)oxides (or metal hydroxysalts); in some
cases, these sparingly soluble species precipitate as nanoparticles (figure 3.1).
For this PhD, unless stated otherwise, all of the GDEx experiments were carried
out in a two-compartment electrochemical flow-cell reactor, designed in-house.
A schematic of the two-compartment electrochemical flow-cell reactor set-up is
given in figure 3.2a and 3.2b.
GDEX PROCESS, AT PROCESS AND KEY PROCESS PARAMETERS 19

a flow cell reactor

Ar

pH sensor catholyte anolyte


reservoir reservoir

magnetic
stirrer bar

magnetic stirrer

gas
b
anion exchange membrane
Vito Core cathode Pt anode
outlet
gas channel

Front-view cross-section of
catholyte
anolyte

side-view cross-section 10 cm2 gas channel/


3.56 cm anolyte compartment/
catholyte compartment

inlet
1 cm
c Ar
titrator tip
pH sensor
NaOH

magnetic stirrer bar


magnetic stirrer

Figure 3.2: Schematic drawing of the GDEx flow-cell reactor set-up (a) and
cross-sections of the flow-cell reactor (b), and the AT reactor set-up (c).
20 EXPLORATORY STUDY OF THE GDEX PROCESS

The anode was a platinized tantalum plate (with a 10 µm Pt layer plated with

®
the molten salt process by Umicore Galvano), further referred to as the Pt
electrode. In general, the cathode was a VITO CoRE GDE [58], but some
experiments were carried out with a Pt electrode as both the anode and the
cathode. Generally, when the Pt electrode is mentioned, it will be in the capacity
of the cathode, because in this PhD, the anode and anolyte compartment only
served to provide the counter-reactions for the requested currents. Otherwise, a
counter electrode will be simply addressed.
Figure 3.2b shows the electrochemical flow-cell reactor. The two compartments

®
(i.e., the anolyte and catholyte compartments) were separated with an anion-
exchange membrane (FUMASEP FAP-4130-PK). Each compartment had
a volume of 10−5 m3 (10 mL). The compartments had a cylindrical shape,
with a length h of 0.01 m and a base surface area of 0.001 m2 (10 cm2 ). The
cylinder axis was perpendicular to the electrode surfaces, giving the electrodes
(both anode and cathode) a geometrical surface area of 0.001 m2 , equal to
the cylinder base area. Each compartment had radial in- and outflow, against

®
gravity, through in - and outlets with diameters of 0.002 m. A peristaltic pump
(Cole- Palmer MasterflexTM L/s 7551-00 with Easyload II 77200-60 pump
heads) was used to provide the flow rate.
The GDEx process was operated as a batch process. Eventually, if GDEx would
become an industrial process, it is more likely that it would be carried out
as a continuous process rather than a batch process. However, at this stage
research, time-dependent data collection is easier for a batch process. Moreover,
it allows to focus on thermodynamic aspects without these being obscured with
kinetic phenomena or noise related to a fast continuous process. Operating
GDEx as a batch process means that the anolyte and catholyte compartments
of the flow cell reactor connect to the anolyte and catholyte reservoirs outside
the reactor, respectively. The electrolyte was pumped from the reservoir to
the flow cell compartment and back. The four connection tubes (Cole-Palmer
MasterflexTM VersilonTM L/s 06475-16) each had a diameter of 0.0032 m and a
length of 0.9 m. The anolyte and catholyte reservoirs were 250 mL DURAN ®
bottles with three-port screw caps. The bottles were adapted with additional
screw-cap necks in an in-house glass-blowing workshop. The additional necks
were needed to accommodate measuring probes (e.g., pH probe, redox probe)
and for sampling material.
When GDEx is operated as a batch process, the catholyte reservoir can be
considered to be a batch reactor, rather than being the recirculation reservoir
of a flow-cell reactor. This is useful because it makes it easier to benchmark
GDEx against other batch reactor processes (e.g., alkaline titration). To operate
GDEx as a batch reactor, the default flow rate was chosen equal to 3.33 10−5
m3 s−1 (200 mL min−1 ). This flow rate was carefully chosen to be high enough
GDEX PROCESS, AT PROCESS AND KEY PROCESS PARAMETERS 21

so the GDEx reactor could be considered as a batch reactor, while also avoiding
to create turbulence in the connection tubes (Re = 1326). At this flow rate, the
combined residence time in the flow cell compartment (3 s), the inlet tube (1
s), and the outlet tube (1 s) was 5 seconds, which is the time-resolution of the
pH measurements. As such, from the eye of the observer (the pH meter), the
species produced heterogeneously by the electrochemical reaction at the surface
of the cathode might as well have been produced or added homogeneously in
the bulk of the catholyte reservoir.
In this work, the GDEx process was by default carried out galavanostatically.
The default cathodically applied current at the GDE was 100 mA (100 A m−2 )
provided with a potentiostat (Biologic Biostat SP-300). In addition, the focus
was on the GDEx process using an air-fed GDE. The default airflow at the back
of the electrode was chosen to be 100 mL min−1 . A back pressure of 30 mbar
was established by putting the end of the gas outlet tube at the bottom of a 30
cm water column. If the process parameters deviated from these default values,
it is pertinently clarified within this work.
The AT process. The AT process could be summarized as follows: during
AT, a concentrated solution of NaOH is slowly added to the electrolyte; the
OH− ions move into the solution and react with the available metal ions to
form sparingly soluble metal (oxy)(hydr)oxides (or metal hydroxysalts); in some
cases, these sparingly soluble species precipitate as nanoparticles (figure 3.1).
®
The AT process was carried out in one 250 mL DURAN bottle (equivalent

®
to the ones used for the GDEx reservoirs) with a titrator (Methrohm 702SM
Titrino). By default, the titrant was a 0.1 M NaOH solution (Titrisol , Merck).
The default titration rate was 0.62 mL min−1 , the equivalent of 1.036 10−6 mol
NaOH s−1 , which is the theoretical addition rate of OH− in the default GDEx
process at 100 A m−2 using a GDE fed with air.
Default electrolyte solutions. The volume of electrolyte solutions used
during GDEx and AT was 100 mL by default. Two solutions were used most
commonly: (1) a default blank solution without metal ions (except for Na+ ),
hereinafter referred to as solutions without metals, and (2) a default solution
with metal ions (other than Na+ ), further referred to as solutions with metals.
Except if specified otherwise, all solutions were made with a supporting
electrolyte of 0.5 M NaCl and experiment-dependent HCl concentrations (most
often 0.001 M or 0.01 M) to establish the initial pH (e.g., pH 3 or pH 2,
respectively).
The composition of the solutions without metals was a 0.5 M NaCl supporting
electrolyte with HCl. Generally, these solutions were made in 1 L volumetric
flasks. 100 mL of a 5.00 M NaCl stock solution together with the chosen amount
22 EXPLORATORY STUDY OF THE GDEX PROCESS

of HCl stock solution (e.g., 1 mL of 1 M HCl to give 0.001 M in total) were


added to the flask and the rest of the volume was filled with distilled water. The
5.00 M NaCl stock solution was made by dissolving 292.2 g of NaCl crystals in
the required amount of water to complete the volume of a 1 L volumetric flask.
The composition of default solutions with metals was 5.0 mM metal chloride
with a 0.5 M NaCl supporting electrolyte and HCl. To make these solutions,
a 1 L stock solution of a high, but inaccurate, concentration of a given metal
salt was made (e.g., 30 g L−1 ZnCl2 for a solution of approximately 0.2 M
ZnCl2 ). The concentration of this solution was measured with ICP-OES (Varian
720-ES) to determine the real concentration of the solution. From the real
concentration, the dilution factor could be calculated to make solutions with
an accurate composition (e.g., 0.200 M ZnCl2 ) via dilution with distilled water.
This accurate stock solution was used to make the solutions for the experiments
with the required metal ion composition. Thus, in a 1 L volumetric flask, 25
mL of 0.200 M metal chloride stock solution, 100 mL of 5.00 M NaCl stock
solution, and the required amount of HCl to reach the required initial pH, were
mixed with the required amount of distilled water to complete 1 L.
During all of the experiments, argon was bubbled in the electrolyte to deplete
the solution of dissolved gasses (e.g., O2 , CO2 , etc.).
The process parameters are summarized in table 3.1.

Table 3.1: Default process parameters


parameter AT GDEx
initial Volume 100 mL 100 mL
titration rate 0.1 M NaOH 100 A m−2
−1
or current density 0.62 mL min
electrolyte flow rate N/A 200 mL min−1
electrode gas flow rate N/A 100 mL min−1
gas overpressure N/A 30 mbar(g)
(water column) (30 cm)

Investigating the GDEx process in more detail was possible by having established
these default process parameters. Despite its operational simplicity, GDEx is
a complex process and additional parameters (not investigated here) can be
manipulated, such as the gas partial pressure (e.g., amount of O2 in an O2 -N2
mixture) and flow rate, and the size and shape of the flow-reactor, among
others. In the next section AT and the batch GDEx process are demonstrated
to indeed have many similarities, both as metal ion removal/recovery processes
and nanoparticle synthesis processes.
HOMOGENEOUS APPEARANCE OF THE GDEX PROCESS 23

2 Homogeneous appearance of the GDEx process

The first aim of this thesis was to discover the homogeneous and heterogeneous
processes of GDEx. Since GDEx is an electrochemical process, it is inherently
governed by heterogeneous electrochemical reactions at the surface of the GDE.
However, the precipitation reactions can happen at the surface of the electrode
(heterogeneous reaction) or in the bulk electrolyte solution (homogeneous
reaction). Here, model and experiment are compared to investigate if these
precipitation reactions are heterogeneous or homogeneous.
In a homogeneous process, the desirable products result from homogeneous
reactions (i.e., everywhere) in the solution. In a heterogeneous process, the
desirable products result from heterogeneous reactions at a surface (i.e., in this
case, the GDE surface). This section demonstrates a simple time-dependent
model of the concentration in the solution for both a heterogeneous and a
homogeneous process. Then, the data of the time-dependent Cu2+ concentration
and the pH during AT and GDEx processes of a CuSO4 electrolyte are compared,
to investigate the nature of the precipitation reactions in GDEx.
Consider 3 species: Cu2+ , OH− , and Cu(OH)2 . The chemical reaction between
the three species can be written as:

Cu2+ + 2OH−
Cu (OH)2 (3.1)

When all species are in equilibrium, their chemical activity is related to the
equilibrium constant constant KCu(OH)2 [mol3 L−3 ] of Cu(OH)2 :
2 2
KCu(OH)2 = aCu2+ (aOH− ) = γCu2+ cCu2+ (γOH− cOH− ) (3.2)

with aj [-] the activity of species j, cj [-] the concentration of species j, γj [-]
the activity coefficient of species j. For activity coefficients equal to 1, equation
(3.2) simplifies to:
2
KCu(OH)2 = cCu2+ (cOH− ) (3.3)
Assuming equilibrium, every time two OH− ions are added to the solution,
these would quickly react with a Cu2+ and these three ions would be removed
from solution in the form of Cu (OH)2 .
Because electrochemical reactions are inherently heterogeneous, the real question
is whether the precipitation reactions happen at (or in the vicinity of) the surface
of the electrode (heterogeneous process), or in the bulk of the electrolyte solution
(homogeneous process). The Cu2+ concentration in the bulk electrolyte was
determined experimentally and the data was compared to models for each of the
possible scenarios (i.e., homogeneous or heterogeneous precipitation reactions).
24 EXPLORATORY STUDY OF THE GDEX PROCESS

Figure 3.3 demonstrates the boundary conditions for both the homogeneous
and heterogeneous scenarios. It is necessary to look at what happens in the
vicinity of the electrode (model domain 1, figure 3.3), to be able to model the
time-dependent Cu2+ concentration in the bulk electrolyte (model domain 2,
figure 3.3). Considering a well-mixed batch reactor, it is possible to define a
diffusion layer of average thickness δ [m] at the surface of the electrode. At this
distance δ away from the electrode surface, the concentration of the species in
solution could be considered equal to that of the bulk electrolyte. Between the
surface of the electrode and the distance δ, the concentration can be calculated
from the time-dependent constitutive equation for mass-transfer. Assuming that
no reactions occur within the diffusion layer (i.e., either the chemical reaction
happens exactly at the surface, for a heterogeneous process, or the chemical
reaction happens outside of the diffusion layer, for a homogeneous process), the
equation is written as:

∂cdl
j (x, t) ∂ 2 cdl
j (x, t)
− Dj =0 (3.4)
∂t ∂x2
−3
with cdl
j (x, t) [mol m ] the concentration of species j in the diffusion layer, Dj
2 −1
[m s ] the diffusion coefficient of species j, and t [s] and x [m] the variables
of time and space, respectively. As a further simplification, it is possible to
assume that within this layer, a steady-state is quickly arrived at compared to
the time-dependent changes in the bulk:

∂cdl
j (x, t)
≈0 (3.5)
∂t
A time-independent concentration profile is then easily found from equation
(3.4), which is assumed to be valid at any instant of time. Setting the time-
derivative to zero in equation (3.4) then results in:

d2 cdl
j (x)
=0 (3.6)
dx2
with a general solution of the form:

cdl
j (x) = C1 x + C2 (3.7)

with C1 and C2 integration constants that follow from the boundary conditions.
The boundary conditions depend on whether a homogeneous or a heterogeneous
process is considered. This is discussed in the following paragraphs and
summarized in figure 3.3.
HOMOGENEOUS APPEARANCE OF THE GDEX PROCESS 25

GDE diffusion layer bulk


model domain 1 model domain 2
𝑑𝑙 𝑑𝑙 𝑏
𝑐𝐶𝑢 2+ 0 =0 𝑐𝐶𝑢 2+  ~𝑐𝐶𝑢 2+ 𝑡

𝑏
𝑑𝑙
d2 𝑐𝐶𝑢 2+ 𝑥 d𝑐𝐶𝑢 2+ 𝑡 d𝑐 𝑑𝑙 2+ 𝑥
=0 𝑉 = −𝐴𝑔𝑒𝑜 𝐷𝐶𝑢2+ 𝐶𝑢
d𝑥 2 d𝑡 d𝑥 𝑥=

heterogeneous 𝑥 = 0 𝑥=
homogeneous 𝑏 −
d𝑐𝑂𝐻 𝑡 𝑑𝑙 −
d𝑐𝑂𝐻 𝑥
𝑑𝑙 −
d2 𝑐𝑂𝐻 𝑥 𝑉 = 𝐴𝑔𝑒𝑜 𝐷𝑂𝐻− − 2𝑅𝐶𝑢(𝑂𝐻)2
=0 d𝑡 d𝑥 𝑥=
𝑏
d𝑥 2 d𝑐 2+ 𝑡
𝑉 𝐶𝑢 = −𝑅𝐶𝑢(𝑂𝐻)2
𝑥=0 𝑥= d𝑡
𝑑𝑙 −
d𝑐𝑂𝐻 𝑥 𝑂𝐻 − 𝐼𝑎𝑝𝑝
=−
𝑑𝑙 −
𝑐𝑂𝐻  ~𝑐𝑂𝐻
𝑏 −
𝑡
d𝑥 𝐴𝑔𝑒𝑜 𝐷𝑂𝐻 − 𝐹
𝑥=0

Figure 3.3: Schematic of the boundary conditions for modeling a homogeneous


or heterogeneous GDEx process.

The heterogeneous process. First (1) consider the case of a heterogeneous


process. The boundary conditions of the heterogeneous process are Dirichlet
boundary conditions at x = 0 and x = δ. Redefining C1 and C2 in equation
(3.7) as C1het and C2het , these integration constants are easily found to be equal
to:
cdl 2+ (δ) − cdl
Cu2+ (0)
C1het = Cu (3.8)
δ
C2het = cdl
Cu2+ (0) (3.9)
The values of cdl
(0) and
Cu2+ cdl
(δ) are found with the necessary approxima-
Cu2+
tions. In the heterogeneous case, the Cu2+ would be removed at the surface of
the electrode (x = 0) as the result of the chemical reaction in equation (3.1). If
this reaction is assumed instantaneous, cdlCu2+ (0) would be equal to zero:

cdl
Cu2+ (0) = 0 (3.10)

At the end of the diffusion layer, the concentration of copper is considered to


be equal to the bulk concentration:

cdl b
Cu2+ (δ) = cCu2+ (t) (3.11)

The assumption of a well-mixed bulk makes cbCu2+ independent of the spacial


variable x. Since Cu2+ is removed from the bulk, its concentration is time-
dependent. This explicit time-dependency appears to invalidate the previous
assumption of a steady-state condition in equation (3.5). This is called the quasi-
steady state approximation. Nevertheless, if the bulk changes are sufficiently
26 EXPLORATORY STUDY OF THE GDEX PROCESS

small, the value of C1het is not altered significantly. As such, it is possible to


demonstrate that this quasi-steady state approximation remains valid only if
the following inequality is true:

∂cbCu2+ (t) cdl 2+ (δ) − cdl 2+ (0)


<< Cu  2 Cu  (3.12)
∂t δ
DCu2+

This will be the case for thin boundary layers (and for relatively large
bulk volumes, because then the concentration changes more slowly). Using
equations (3.10) and (3.11), the quasi-steady state approximation of the Cu2+
concentration profile at the electrode surface in the heterogeneous case can thus
be written as follows:
x
cdl b
Cu2+ (x) = cCu2+ (t) δ (3.13)

To find the explicit expression of cbCu2+ (t), it is necessary to solve the mass
balance over the bulk solution (model domain 2):

ACCUMULATION = IN − OUT − REACTION


dcb 2+ (t) (3.14)
Vtot Cu = 0 − out
NCu 2+ Ageo − 0
dt
with Ageo [m2 ] the geometric surface area of the electrode, Vtot [m3 ] the volume
out
of the bulk electrolyte, and NCu2+ (δ) [mol m s ] the outgoing flux of Cu2+
−2 −1

at the diffusion layer boundary, equal to:

dcdl cbCu2+ (t)



Cu2+ (x)
out

NCu2+ (δ) = DCu2+ = −D Cu 2+ (3.15)
dx
x=δ δ

Using equation (3.14) and (3.15), the mass balance equation can be written as
a first-order ordinary differential equation:

dcbCu2+ (t) Ageo DCu2+ b


=− cCu2+ (t) (3.16)
dt Vtot δ
whose solution is:
Ageo D
Cu2+
cbCu2+ (t) = cbCu2+ (0) e− Vtot δ t (3.17)

The homogeneous process. Now (2) consider the case of a homogeneous


process. In this case, solving directly for the concentration of Cu2+ is less
desirable. Instead, it is easier to first solve for the OH− concentration and then
solve for the Cu2+ concentration because the boundary conditions for OH− are
HOMOGENEOUS APPEARANCE OF THE GDEX PROCESS 27

more easily defined. The boundary condition at the surface of the electrode
(x = 0) is now a Neumann boundary condition instead of a Dirichlet boundary
condition, while the boundary condition at x = δ remains a Dirichlet boundary
condition.
The boundary condition at the electrode surface is calculated from the
production rate of OH− at the surface of the electrode, POH− [mol s−1 ], given
by:
η − Iapp
POH− = OH (3.18)
F
with ηOH− [-] the efficiency for OH− production, Iapp the applied current [A], and
F = 96, 485 mol C−1 the Faraday constant. The model defines the interaction
between the bulk and the electrode. Because of this only the geometric surface
area of the electrode is of interest, rather than the effective wetted surface
of the electrode. Using the geometric surface area of the electrode Ageo , it is
possible to define the flux of OH− at the surface as R/Ageo and set it equal to
dc − (0)
the flux of OH− caused by the concentration gradient −DOH− OH dx at the
surface. From this follows the Neumann boundary condition at the surface of
the electrode:
dcdl

OH−
(x) POH−
=− (3.19)
dx
x=0 DOH− Ageo

The second boundary condition is equivalent to that of the heterogeneous case.


The concentration of OH− at a distance δ from the electrode cdlOH−
(δ) can be
− b −3
set equal to the bulk concentration of OH , cOH− (t) [mol m ]:

cdl b
OH− (δ) = cOH− (t) (3.20)

Similar to the heterogeneous case, this introduces an explicit time-dependent


term in a steady-state approximation. Again, this approximation will remain
valid under similar conditions as previously described. Redefining C1 and C2 in
equation (3.7) as C1hom and C2hom , these integration constants are easily found
to be equal to:
POH−
C1hom = − (3.21)
DOH− Ageo
POH− δ
C2hom = cdl
OH−
(δ) + (3.22)
DOH− Ageo
Using equations (3.21) and (3.22), the quasi-steady state approximation of the
OH− concentration profile at the electrode surface in the homogeneous can thus
be written as follows:
POH− δ  x
cdl b
OH− (x) = cOH− (t) + 1− (3.23)
DOH− Ageo δ
28 EXPLORATORY STUDY OF THE GDEX PROCESS

Equivalent to the heterogeneous case (equation (3.14)), the explicit expression


of cbOH− (t) can be found by solving the mass balance over the bulk solution
(model domain 2):

ACCUMULATION = IN − OUT − REACTION


dcb − (t) (3.24)
Vtot OH = in
NOH − Ageo − 0 − 2RCu(OH)2 (t)
dt
with RCu(OH)2 (t) [mol s−1 ] the time-dependent formation rate of Cu(OH)2 ,
and NOHin
− (δ) [mol m s ] the in-going flux of OH− at the diffusion layer
−2 −1

boundary, equal to:


dcdl

in OH−
(x) 
POH−

NOH − (δ) = −D OH − = −DOH− − DOH− Ageo
dx
x=δ (3.25)
POH−
= Ageo

Using equation (3.24) and (3.25), the mass balance equation can be written as
an ordinary differential equation:
dcbOH− (t)
Vtot = POH− − 2RCu(OH)2 (3.26)
dt
A similar mass balance can be written for Cu2+ :
dcbCu2+ (t)
Vtot = −RCu(OH)2 (3.27)
dt
dcbCu2+ (t)
From equations (3.26) and (3.27), the following relationship between dt
dcbOH− (t)
and dt is obvious:

dcbCu2+ (t) 1 dcbOH− (t) POH−


= − (3.28)
dt 2 dt 2Vtot
This equation is easily solved with the assumption that all of the OH− ions
dcbOH− (t)
that reach the bulk, immediately react with Cu2+ ions. In this case, dt
is approximately equal to zero:
dcbOH− (t)
≈0 (3.29)
dt

Equation (3.28), using equation (3.18), thus simplifies to a first-order ordinary


differential equation for the concentration of Cu2+ in the bulk electrolyte:
dcbCu2+ (t) POH− η − Iapp
= − = − OH (3.30)
dt 2Vtot 2Vtot F
HOMOGENEOUS APPEARANCE OF THE GDEX PROCESS 29

with its solution given as:


ηOH− Iapp
cbCu2+ (t) = cbCu2+ (0) − t (3.31)
2Vtot F

Comparison of the heterogeneous and homogeneous processes. A


comparison of equations (3.17) and (3.31) shows remarkable differences. For the
heterogeneous process, the concentration of Cu2+ ions is exponential in time,
while that for the heterogeneous process is linear in time. Moreover, for the
heterogeneous process, the inherent properties of the Cu2+ ion are important
(i.e., diffusion coefficient), but they not for the homogeneous process. In addition,
the geometric electrode area Ageo and the diffusion layer thickness δ matter
for a heterogeneous process, while they are unimportant for the homogeneous
process. For the homogeneous process, the only important parameters are
related to the production rate of OH− , POH− (which is related to the applied
electric current Iapp and the current efficiency ηOH− for OH− production).
Figure 3.4 demonstrates the time-dependent Cu2+ concentration in the bulk
electrolyte, cbCu2+ (t) for both the homogeneous and the heterogeneous cases,
using the parameter values from table 3.2. It is clear that for the chosen values
of the model parameters (i.e., they are in line with the experimental process
parameters), the homogeneous process would be far superior in terms of the
removal rate of the Cu2+ ions from solution.

Table 3.2: Parameters for modeling the bulk concentration of Cu2+


parameter value
Ageo 0.001 m2
cbCu2+ (0) 5 mol m−3
DCu2+ 7.14 10−10 m2 s−1 [104]
F 96,485 C mol−1
Iapp 0.1 A
Vtot 0.0001 m3 (100 mL)
δ 4 10−4 m (arbitr.)
ηOH− 1
30 EXPLORATORY STUDY OF THE GDEX PROCESS

5
𝐴𝑔𝑒𝑜 𝐷Cu2+
− 𝑡
4.5 𝑏
𝑐Cu 𝑏
2+ 𝑡 = 𝑐Cu2+ 0 𝑒
𝑉𝑡𝑜𝑡 

Cu2+ concentration [mol m-3]


4
𝑏 𝑏 OH− 𝐼𝑎𝑝𝑝
𝑐Cu2+ 𝑡 = 𝑐Cu2+ 0 − 𝑡
3.5 2𝑉𝑡𝑜𝑡 F

3
2.5
2
1.5
1
0.5 heterogeneous
homogeneous
0
0 500 1000
time t [s]

Figure 3.4: Comparison of the modeled concentration of Cu2+ in the bulk of


the electrolyte for the cases of a heterogeneous process (equation (3.17)) and a
homogeneous process (equation (3.31)). The used parameters are given in table
3.2.

Knowing this, it would be beneficial if the precipitation reactions in the


GDEx process of Cu2+ would be homogeneous in nature, even though the
electrochemical production of OH− is inherently heterogeneous in nature. To
investigate this experimentally, an AT experiment with NaOH and a GDEx
experiment were carried out for a CuSO4 solution. For the GDEx experiments,
the process conditions were chosen so that the parameters in table 3.2 were
valid (i.e., Ageo , cbCu2+ (0), Iapp , Vtot ). The AT process was carried out as a
reference case because (slow) alkaline titration in a well-stirred beaker (i.e., a
batch reactor) can be considered a solely homogeneous process. To be able
to compare both processes, the addition rate of NaOH was matched to the
theoretical electrochemical addition rate of OH− (i.e., Iapp /F = 1.036 10−6
mol s−1 , equivalent to a titration rate of 0.62 mL min−1 of 0.1 M NaOH). The
time-dependent Cu2+ concentration was measured using a Cu2+ ion-selective
electrode (Cu-ISE, Methrohm) and also the pH was measured over time. For AT,
the measurements were done directly in the batch reactor volume. For GDEx,
the measurements were done in the catholyte reservoir. The concentration in
the catholyte reservoir of the two-compartment GDEx electrochemical flow-cell
reactor with complete recirculation behaves like that of a continuously stirred
tank electrochemical reactor (CSTER), which converges to behave like a batch
HOMOGENEOUS APPEARANCE OF THE GDEX PROCESS 31

reactor at high flow rate and long time frames. The method for measuring the
Cu2+ concentration is described in appendix A.
Figure 3.5 shows the experimental results of both experiments. For comparison,
the modeled curves for the homogeneous and heterogeneous reactions are also
provided in the figure. It is immediately obvious that both the AT and GDEx
processes behave in the same way. As expected, the AT process behaves like
a homogeneous process, demonstrating a time-dependent Cu2+ concentration
similar to that of figure 3.4, with similar process times (i.e., approximately
1000 seconds to complete removal). The GDEx process also demonstrates this
behavior. This proves the homogeneous nature of the precipitation reactions in
the GDEx process, under the given process conditions.
Despite the obvious similarities between model and experiment, some differences
should be pointed out. The first observed difference (1) is that the experimental
Cu2+ concentration does not immediately start to decrease. This is because
for the experiments the value of the initial pH was chosen below the value
where the Cu(OH)2 equilibrium is valid. This was done deliberately because it
allows a clear-cut initiation of the precipitation process. Starting the process
at this lower pH delays the initial time where the Cu2+ concentration starts
decreasing. To completely understand the shape of the pH curves, Chapter
4 is dedicated entirely to matching the model and the experiment of the AT
process for a reference case of ZnCl2 . The second observed difference (2) is
that, even after taking into account the delay, the experimental time-frame
is approximately 75% shorter than what was expected theoretically. This is
caused due to a discrepancy between the theoretically predicted precipitate (i.e.,
Cu(OH)2 ) and the experimentally observed precipitate (i.e., Cu4 (OH)6 SO4 ).
The theoretical stoichiometry of Cu(OH)2 is 1 Cu2+ for every 2 OH− , while
the experimental stoichiometry of Cu4 (OH)6 SO4 is 1 Cu2+ for every 1.5 OH− .
Therefore, the value of the theoretical removal rate of Cu should be 75% (i.e.,
1.5 divided by 2) of the value of the theoretical removal rate. In other words,
the experimental removal time should be 75% of the theoretical removal time,
which is confirmed in this case. The third observed difference (3) is the slight
curvature of the relatively linear decrease of the experimental Cu2+ . Given
that the curves of AT and GDEx run rather parallel, this might be the result of
calibration errors, which can be significant when the concentration changes are
small (i.e., concentration changes within the same order of magnitude) because
the measured potential is a logarithmic function of the Cu2+ concentration.
Lastly, one additional observation can be made, which is (4) the discrepancy
between AT and GDEx regarding both the beginning and ending times of the
removal/recovery process, indicated by the orange lines in figure 3.5. Moreover,
the difference increases over time (i.e., the difference between the beginning time
for AT and GDEx is smaller than the difference between the ending time for
32 EXPLORATORY STUDY OF THE GDEX PROCESS

5 12
4.5
10

Cu2+ concentration [mol m-3]


4
3.5
8
3

pH
2.5 6
2
4
1.5
GDEx
1 AT 2
0.5 het.
hom.
0 0
0 500 1000 1500
added NaOH or charge [mol]

Figure 3.5: Experimental time-dependent pH and concentration profiles of AT


with NaOH and the GDEx process for 100 mL solutions containing 0.005 M
CuSO4 and 0.25 M Na2 SO4 with 5 10−4 M H2 SO4 supporting electrolyte. The
yellow lines indicate the discrepancy between the beginning and end times of
the Cu2+ removal. Default process conditions (table 3.1) and other parameter
values (table 3.2).

AT and GDEx). This observation was considered so interesting, that Chapter 7


is completely dedicated to demonstrate the origin of this delay in the GDEx
process.
AT and GDEx were demonstrated to both behave as homogeneous processes.
This is beneficial for their use as metal removal/recovery processes. However,
equally important, it is beneficial in the same way for their use as a nanoparticle
synthesis methods. Since the removal rate of the metal ion is directly related
to the production rate of the precipitate. Since the desired nanoparticle is the
precipitate in a synthesis process, the homogeneous nature of either process
allows for much higher production rates than if they would be heterogeneous.
Although the precipitation reactions for both AT and GDEx were shown to be
homogeneous, important differences between both processes remain. The next
section demonstrates where GDEx differs significantly from AT: the production
of H2 O2 .
GDEX AND THE OXYGEN REDUCTION REACTION: PRODUCTION OF H2 O2 AND HIGH SOLUTION
REDOX POTENTIALS 33

3 GDEx and the oxygen reduction reaction: pro-


duction of H2 O2 and high solution redox poten-
tials

One of the aims of this PhD was to discover what caused oxidation of Ce(III)
to Ce(IV) in the catholyte side during GDEx. The oxidation of metal ions in
the catholyte side is remarkable since at the cathode a reduction reaction is
carried out, which should expectedly yield species that decrease the solution
redox potential and cause reduction reactions, rather than oxidation reactions.
Gas-diffusion electrodes are designed for electrochemical processes wherein at
least one of the reagents is in the gas form (i.e., not dissolved in solution).
They provide large surface areas to increase the triple-phase boundary of
gas/electrolyte/electrode and as such provide an answer to the mass-transfer
limitations posed by the combination of a small electrode surfaces and the
relatively low solubility of most gasses in water [105]. Passing air (or oxygen)
through a GDE is a well-known method to carry out the 4-electron and 2-
electron oxygen reduction reactions (ORRs), among others. The latter reaction
reduces two O(0) in O2 to two O(-I) in H2 O2 (in acid solutions) or HO− 2 (in
alkaline solutions), both well-known oxidants:

acidic : O2 + 2H+ + 2e−


H2 O 2 (E◦ = 0.695 V)
(3.32)
alkaline : O2 + H2 O + 2e−
HO−2 + OH

(E◦ = −0.065 V)

The formation of, what are considered, oxidants with a reduction reaction is
odd. In this case, it results from the fact that many species (e.g., many of the
transition metal ions) have a lower oxidation potential than H2 O2 , even though
it is the product of a reduction reaction of O2 . At the electrode surface, the
potential is chosen sufficiently low for the reduction of O2 to H2 O2 (or HO−
2 in
alkaline media) [20]. However, once in the bulk electrolyte, H2 O2 can oxidize
species with lower oxidation potentials.

®
Figure 3.6 demonstrates the change in pH and H2 O2 concentration as a result

®
of carrying out ORRs at the VITO CoRE GDE [58]. The H2 O2 concentration
was measured with QUANTOFIX measuring strips (details in chapters 6
and 7). Remark that the H2 O2 concentration is measured and not the HO− 2
concentration because, even if HO− 2 would be produced at the electrode, for
the practical pH ranges of the bulk electrolyte, it would be protonated to H2 O2
(pKH2 O2 = 11.62 [20]).
34 EXPLORATORY STUDY OF THE GDEX PROCESS

14
100% efficiency
14
12

concentration H2O2 [10-3 M]


12
10
10
8
8

pH
30% efficiency
6
6

4 4

2 H 2O2
Series2 2
pH
0 0
0 500 1000 1500
time [s]

Figure 3.6: Experimental time-dependent pH and H2 O2 concentration of


the GDEx process of a 0.5 M NaCl solution with 0.05 M HCl. The line of a
theoretical 100% production efficiency of H2 O2 is shown in comparison to the
experimentally observed production efficiency. Exception to default process
conditions (table 3.1): current density 500 A m−2 .

Figure 3.6 shows a relatively linear increase in the concentration of H2 O2 , with


an observed current efficiency of approximately 30% for the production of H2 O2 .
The production efficiency for H2 O2 is further discussed in Chapter 7.
To demonstrate the oxidative effect of the produced H2 O2 in solution,
galvanostatic experiments were carried out while simultaneously measuring
both the pH and redox potential in the catholyte (reservoir). The redox
potential was measured as the potential difference between a platinum wire
and an Ag/AgCl reference electrode (3 M KCl, Metrohm), and recalculated
to the standard hydrogen electrode. Two set-ups were compared using the
same flow-cell reactor: (1) one with a VITO CoRE GDE as the cathode ®
and (2) another one with a platinized tantalum plate as a cathode. Otherwise,
the default process conditions were used (table 3.1). Figure 3.7a shows the
time-dependent measurements, while figure 3.7b shows the same data, linking
the measured redox potential to the measured pH data for every measured time
point. As such, an Eredox -pH diagram can be produced. The grey lines in figure
3.7b show the theoretically expected dependence of the potential and the pH for
the corresponding reactions in ideal standard conditions (25 ◦ C, water without
electrolyte, etc.). The middle grey line was calculated for the experimentally
measured H2 O2 concentration.
GDEX AND THE OXYGEN REDUCTION REACTION: PRODUCTION OF H2 O2 AND HIGH SOLUTION
REDOX POTENTIALS 35

(a) 1 GDE cathode: ORR


12 (b) 1.2 GDE cathode: ORR
Pt cathode: HER 1 Pt cathode: HER

redox potential Eredox vs. SHE [V]


redox potential Eredox vs. SHE [V]

0.8
10
0.8
0.6
0.6
8
0.4 0.4

pH
0.2 6 0.2

0 0
4
-0.2
-0.2
-0.4
2
-0.4 -0.6
-0.6 0 -0.8
0 500 1000 1500 2 3 4 5 6 7 8 9 10 11
time [s] pH

Figure 3.7: (a) Experimental time-dependent pH and redox potential of


the catholyte reservoir in two-compartment electrochemical flow-cell reactor
experiments using either a GDE cathode or a Pt cathode. (b) Eredox -pH
diagrams composed of time-matched data of figure (a) to demonstrate the
Eredox -pH changes in the catholyte during the experiments. The outer grey
lines demonstrate theoretical Eredox -pH lines for standard reactions in water,
the inner grey line is the theoretical ORR line for the H2 O2 concentrations
and pH measured in the GDEx experiment but applied to water in standard
conditions. The initial composition of the catholyte was 100 mL 0.5 M NaCl
with 0.01 M HCl. Default process conditions (table 3.1).

The discrepancy between both experiments is very clear. Using the Pt electrode
creates a reducing environment (potentials below 0 V vs. SHE). In contrast,
using the GDE creates an oxidizing environment (potentials above 0 V vs. SHE).
The theoretical relationship between the potential and the pH (-0.059 V per pH
unit) is relatively well observed in the experiments. The agreement at low pH in
the case of the platinum electrode is especially remarkable. For the Pt electrode,
the redox potential drops rapidly at the beginning of the experiment (figure
3.7a). This is likely caused by dissolved H2 in solution rapidly establishing a
potential close to the theoretical redox potential (figure 3.7b). In the case of
the GDE experiments, the increase is rather limited, but the potential decrease
with increasing pH is barely present. This is likely caused by the increasing
concentration of H2 O2 . The delay in pH increase between the experiment with
the GDE and the Pt electrode is remarkable and shows strong similarity to that
observed in section 2 in figure 3.5. This delay is discussed in Chapter 7.
In addition to the above observations, it is worth noting that the measured H2 O2
36 EXPLORATORY STUDY OF THE GDEX PROCESS

concentration and Eredox -pH values are done in non-equilibrium measurements.


At equilibrium, all of the H2 O2 would eventually self-decompose. Moreover,
whether or not H2 O2 will act as an oxidizing agent or as a reducing agent
depends on the composition of the solution and the other actors in the redox
reactions.
The take-away message is that, although a reduction reaction is carried out
at the GDE, the redox potential can be pushed upwards nonetheless. This is
the result of the relatively high reduction potential of O2 to H2 O2 , and the
dual nature of H2 O2 . The increased redox potential is especially important at
higher pH because most metal ions tend to be more easily oxidized at higher
pH [20]. This is why it is possible to oxidize metal ions in the catholyte side
and why Ce(IV) oxide was observed in the initial experiments at VITO. In this
way, these experiments answered one of the essential research questions of this
PhD. To fundamentally understand the effect of H2 O2 on the characteristics
of selected metal (oxy)(hydr)oxide nanoparticles, a separate investigation was
conducted, of which the results are discussed in Chapter 5. To see this effect in
practice, the nanoparticles produced with AT and GDEx for several metal ion
solutions were compared, and these results are discussed in Chapter 6.

4 Universality and selectivity of AT and GDEx


processes

In section 2, the precipitation reactions in GDEx were demonstrated to be


homogeneous, with a very high degree of similarity to AT with NaOH. For this
PhD, the focus was on using GDEx with an air-fed GDE for the precipitation of
metal (oxy)(hydr)oxides and metal (oxy)(hydr)oxy salts. Given that chemical
reaction precipitation only occurs if a sparingly soluble species is produced, the
universality of the technique is inherently linked to the solubility of these metal
(oxy)(hydr)oxides and metal (oxy)(hydr)oxy salts. Metals that cannot form such
sparingly water-soluble species are per definition excluded from the applicability
of GDEx in aqueous solutions with an air-fed GDE (or AT with NaOH, or any
base for that matter). Remark that using GDEx with other gasses might result
in different precipitants, which might help the removal/recovery of metal ions
in other forms than as the metal (oxy)(hydr)oxide. Moreover, GDEx is not
limited to aqueous media and current investigations include GDEx in organic
media such as ionic liquids, and deep eutectic solvents, among others. Therefore,
GDEx is a very versatile and broadly applicable method.
To further demonstrate the point of universality, AT experiments were performed
for several metal chloride solutions (figure 3.8). Instead of GDEx, AT was chosen
UNIVERSALITY AND SELECTIVITY OF AT AND GDEX PROCESSES 37

14

12

10

8
pH

6
Cu
CuCl2
Zn
ZnCl2
4
Fe
FeCl2
Co
CoCl2
2 Ni
NiCl2
Mn
MnCl2
0
0 0.001 0.002
added NaOH [mol]

Figure 3.8: Experimental time-dependent pH of AT for 100 mL solutions


containing 0.005 M CuCl2 , or ZnCl2 , or FeCl2 , or CoCl2 , or NiCl2 , or MnCl2
and a 0.5 M NaCl with 0.001 M HCl supporting electrolyte. Default process
conditions (table 3.1).

to demonstrate the universality because AT simulates GDEx very well in terms


of the chemical reactions that happen in the bulk. It does so without the
increased complexity of the flow-cell, the electrode processes, the delays of the
inflection points, etc. For reasons of consistency, a large amount of first-row
transition metal ions were chosen, all of the same +II oxidation state. Figure
3.8 demonstrates that all of the metals exhibit similar precipitation behavior.
Except for CuCl2 , all the precipitation plateaus are of the same length. The
discrepancy is caused by the fact that in the case of CuCl2 , the sparingly soluble
tribasic copper chloride (i.e., Cu2 (OH)3 Cl) is precipitated instead of the copper
hydroxide (i.e., Cu(OH)2 ), which needs less hydroxyl ions to form a neutral
species and precipitate (i.e., similar to the CuSO4 case in figure 3.5).
From figure 3.8, and with the default GDEx conditions here investigated, one
can estimate which metals might be removed selectively more easily than others.
Intuitively, species for which the precipitation plateaus occur at substantially
different values of the pH are expected to be more easily separated. Knowing this,
one could hypothesize that dissolved copper and zinc would be easily separated
with AT and GDEx, while separation of nickel and cobalt would be impossible
with these techniques and the default conditions here investigated. This has
38 EXPLORATORY STUDY OF THE GDEX PROCESS

12 12
(a) CuCl2 and ZnCl2 (b) NiCl2 and CoCl2

10 10

8 8

pH
pH

6 6

4 4

2 2

0 0
0 0.0005 0.001 0.0015 0 0.0005 0.001 0.0015
added charge [mol] added charge [mol]

Figure 3.9: Experimental time-dependent pH of GDEx for 100 mL solutions


containing 0.0025 M CuCl2 and 0.0025 M ZnCl2 , or 0.0025 M CoCl2 and 0.0025
M NiCl2 , and a 0.5 M NaCl with 0.001 M HCl supporting electrolyte. Default
process conditions (table 3.1).

been experimentally tested and the results are shown in figure 3.9. Mixtures of
equal amounts of CuCl2 and ZnCl2 , and of NiCl2 and CoCl2 , were subjected
to the GDEx process (see table 3.1 for process conditions). While GDEx of
the copper-zinc mixture demonstrates a clear separation of the precipitation
plateaus, such separation is indistinguishable for GDEx of the nickel-cobalt
mixture. Then again, it has been known for a very long time that nickel-cobalt
mixtures pose difficulties for such type of separation. For the interested reader,
Sole’s recent review is an excellent starting point [106]. Nevertheless, other
processes and mechanisms are available for such cases. Some of which might be
possible using GDEs (e.g., careful choice of cathode potential, use of different
gas, etc.). However, their discussion would lead us too far afield.
In addition, the selectivity of the metal hydroxides is complicated by the fact
that it depends on the individual concentrations of the dissolved metals. The
limitation of metal hydroxide separation can be demonstrated mathematically
as shown in the next paragraphs.
Assume a solution with a sufficiently low pH to have a mixture of two completely
dissolved metal ions: defining metal 1 as the metal having the lowest - and
metal 2 as the metal having the highest metal hydroxide solubility in water. For
infinitely slow (thermodynamic) reactive precipitation it is possible to determine
UNIVERSALITY AND SELECTIVITY OF AT AND GDEX PROCESSES 39

the thermodynamic selectivity limitations. The effective selectivity can be


redefined as the removal efficiency η1 [-], which is the relative amount of metal
1 ions removed during reaction precipitation before metal 2 would start to
precipitate as well (i.e., ending the selectivity):
c1,f
η1 = 1 − (3.33)
c1,0

This removal efficiency can be easily calculated with the necessary assumptions.
To demonstrate the point, the calculations are made assuming that only the
metal hydroxides are precipitating and no other reactions occur (e.g., oxidation,
etc.). Another assumption is to ignore the occurrence of multi-species ions, which
is, in general, a reasonable assumption (see figure 4.2 on page 54 for an example
of the speciation of aqueous zinc). An arbitrary metal hydroxide equilibrium
can be defined using the equilibrium constant of that metal hydroxide in water:
ni
KM(OH)n = a n + (aOH− ) (3.34)
i M i
i

with KM(OH)n [-] the equilibrium constant of metal hydroxide M(OH)ni of


i
metal ion Mni + with valence ni , and aj [-] the activity of ion j. For simplicity,
the value of the activity coefficient γj [-] of species j is assumed to be equal to
1, so that the activity aj is equal to the concentration cj of a given species:

aj = γj cj (3.35)

An expression for the water equilibrium is also needed. For the value of the
activity coefficients of H+ and OH− equal to 1, the water equilibrium constant
KH2 O [mol2 L−2 ] is defined as:

KH2 O = cH+ cOH− (3.36)



The pH is defined as − log cH+ for a proton activity coefficient equal to 1. For
all activity coefficients equal to 1, the pH of a solution containing metal ions
with sparingly soluble metal hydroxides can thus be written as a function of
the concentration of that metal ion using equations (3.34), (3.35), and (3.36):
!
n
1 ci (KH2 O ) i
pHi = − log (3.37)
ni KM(OH)n
i

The removal efficiency can be calculated by setting the initial precipitation pH


of metal 2 (i.e., pH2,0 ) equal to that of the final precipitation pH of metal 1
(i.e., pH1,f ):
40 EXPLORATORY STUDY OF THE GDEX PROCESS

! !
n2 n1
1 c2,0 (KH2 O ) 1 c1,f (KH2 O )
pH2,0 = − log = − log = pH1,f
n2 KM(OH)n n1 KM(OH)n
2 1
(3.38)
Using all of the prior definitions, the removal efficiency can be calculated as:
1 n
(c2,0 ) n2 KM(OH)n
η1 = 1 − 1
 nn1 (3.39)
c1,0 
2
KM(OH)n
2

The valences and the equilibrium constants are fixed for two chosen metals.
When considering two metals with equal valences, for example, η1 is a direct
function of the ratio of the two initial metal ion concentrations. The process
is more selective for one mixture compared to another mixture, if the removal
efficiency is higher for a given ratio of the initial concentrations c2,0 /c1,0 . In
figure 3.10, a higher selectivity for the case of Cu/Zn in comparison to the case
of Ni/Co can thus easily be recognized. This figure also suggests that at a
50/50 ratio, the nickel removal efficiency can still be as high as 45%, contrary to
what figure 3.9 might suggest. Therefore, in theory, stopping the precipitation
process when the pH reaches the initial precipitation pH of cobalt hydroxide,
45% of the nickel would have already been precipitated. In practice, this is
much harder to achieve, because, at the location where the base is introduced, it
would be difficult to control the pH well enough to selectively precipitate nickel
but not cobalt. Although the experimental verification would be interesting, it
was not carried out for this PhD.

5 Energy considerations of GDEx and effect of


supporting electrolyte

In the previous sections, many aspects of the GDEx process have been introduced.
GDEx was demonstrated to be a partly homogeneous process, very similar to
AT. Therefore, the precipitation of metal (oxy)(hydr)oxide or metal hydroxysalt
(nano)particles with GDEx, is essentially caused by the production of OH− ,
which increases the pH. However, in section 3, it was demonstrated that by
using the GDE to carry out oxygen reduction reactions H2 O2 is produced,
which causes a high redox potential. This can cause the oxidation of metal ions
and affect the final characteristics of the produced nanoparticles. The effect of
oxidation is discussed in detail in chapters 5 and 6.
Here, the GDEx process demonstrates another virtue: its potential to work
at lower power consumption than alternative electrochemical OH− producing
ENERGY CONSIDERATIONS OF GDEX AND EFFECT OF SUPPORTING ELECTROLYTE 41

100%
90%

removal efficiency of metal 1


80%
70%
60%
50%
40%
30%
20%
Ni (1), Co (2)
10%
Cu (1), Zn (2)
0%
-4 -3 -2 -1 0 1 2 3 4
initial concentration ratio log(c2,0/c1,0 )[-]

Figure 3.10: Theoretical removal efficiency of metal 1 for two mixtures of


metal ions with valence 2 as a function of the ratio of initial concentrations.
The number in brackets indicates the theoretical sequence of precipitation
for equal concentrations of the metal ions. The orange line indicates the
removal efficiencies for mixtures of 50/50 ratios. Values for the metal hydroxide
equilibrium constants were chosen as KNi(OH)2 = 5.5 10−16 mol3 L−3 [107],
KCo(OH)2 = 1.0 10−15 mol3 L−3 [107],KCu(OH)2 = 2.2 10−20 mol3 L−3 [107],
and KZn(OH)2 = 3.0 10−17 mol3 L−3 [108].

reactions. The power consumption of the galvanostatic experiments with a


GDE and with a Pt electrode, previously discussed in section 3, are compared.
Figure 3.11a demonstrates the time-dependent potential of the experiments
with an air-fed GDE and with a Pt electrode. Figure 3.11b demonstrates the
overall cell potential for those same experiments and the corresponding power
consumption. There is a large difference of 1 V between the work potential of
the GDE and the work potential of the Pt electrode for the same current density
of 100 A m−2 . This is caused by the fact that at the GDE, oxygen reduction
reactions take place (E◦ = 0.401 V vs. SHE for the 4-electron reduction and E◦
= -0.065 V vs. SHE for the 2-electron reduction [63] in alkaline solutions), while
at the Pt electrode, due to the lack of oxygen, the hydrogen evolution reaction
takes place (i.e., E◦ = -0.826 V vs. SHE in alkaline solutions). The theoretical
difference between both electrodes is therefore expected to be 0.76 V or higher.
In the case studied here, the power consumption is up to 40% higher for the
42 EXPLORATORY STUDY OF THE GDEX PROCESS

experiment using the Pt electrode. Because of this large difference, it is clear


that producing OH− using an air-fed GDE is energetically more favorable than
producing OH− with the HER at a Pt electrode. This is exactly the reason
for the current rise in the use of gas-diffusion electrodes for the chlor-alkali
process [109–113].

(a) 0 (b) 0.4 4


GDE
-0.2 Pt 0.35 3.5

cell power consumption [W]


0.3 +40 % 3
potential vs. SHE [V]

total cell voltage [V]


-0.4
0.25 2.5
-0.6
0.2 2
-0.8 1V
0.15 1.5
-1
0.1 1
-1.2 0.05 Pt 0.5
GDE
-1.4 0 0
0 500 1000 1500 2000 0 500 1000 1500 2000
time [s] time [s]

Figure 3.11: Experimental time-dependent (a) working electrode potential


and (b) time-dependent power consumption and total cell voltage for a two-

®
compartment flow cell reactor experiments with a Pt plate electrode compared
to an air-fed VITO CoRE GDE. Default process conditions.

The energy consideration is incomplete if, besides the power consumption,


the treatment time is not taken into account (i.e., energy = power × time).
Solutions that require longer treatment time are more energy-intensive and are
therefore indirectly more costly. In respect to this, retaking the time-dependent
copper concentration (equation (3.31) on page 29), it is possible to calculate
the time to remove all the initial copper:

cbCu2+ (0) 2Vtot F


tremoval = (3.40)
ηOH− Iapp

Therefore, the removal time is directly proportional to the amount of cbCu2+ (0)
that is present, and the total volume Vtot . This is expected since it is obvious
that higher volumes with more species to be removed will take a longer time for
a given rate. The last important proportionality factor is the factor 2, which is
the amount of OH− necessary to remove 1 Cu2+ ion. This factor is equal to
the valence of the metal ion if the precipitate is the metal hydroxide or metal
ENERGY CONSIDERATIONS OF GDEX AND EFFECT OF SUPPORTING ELECTROLYTE 43

12

10

pH 8

2
NaCl
Na2SO4
Na2SO4
0
0 0.0005 0.001
added NaOH [mol]

Figure 3.12: Experimental time-dependent pH for AT of 0.5 M NaCl and 0.25


M Na2 SO4 with an initial pH of 3.

oxide. Therefore, to precipitate all of the metal ions, solutions with metal ions
with a valence of 3 will need 1.5 times more OH− (and thus time, and energy)
than solutions with metal ions with a valence of 2. This factor is different
from the valence if negative anions participate in the precipitated molecule
(e.g., Cu4 (OH)6 SO4 ). This was demonstrated for the case of a 0.005 M CuSO4
solution (figure 3.5 on page 32), where this factor is lower than the valence of
Cu2+ (e.g., 1.5 instead of 2). In contrast, if a cation would be incorporated (e.g.,
Na+ ), this factor would be higher than the valence. Therefore, the available
cations and anions in the supporting electrolyte can be very important in the
consumption of OH− ions in traditional AT as well as in the GDEx process.
Anions can play another important role in the consumption of OH− ions. If
anions are conjugate bases of polyprotic acids they can act as buffers in certain
pH ranges. This is demonstrated in figure 3.12 by comparing AT of two different
supporting electrolytes without metals at the same initial pH of 3: 0.5 M NaCl
and 0.25 M Na2 SO4 . In the Na2 SO4 case, it takes up to 5 times more NaOH to
reach the initiation of the water inflection point than in the NaCl case. This
is caused by the high amount of HSO− 4 in the solution, from which protons
need to be removed before a significant pH change can be obtained. This high
amount of HSO− 4 is caused by the high amount of Na2 SO4 , which disturbs the
sulphate-bisulphate equilibrium. Thus, electrolyte solutions containing buffers
44 EXPLORATORY STUDY OF THE GDEX PROCESS

will be more energy-intensive than unbuffered solutions. This is important in


metal removal/recovery processes but it should also be taken into account if the
aim of the GDEx process is the production of nanoparticles. If working with
buffering anions is not required to obtain a desired precipitate, it is better to
work with unbuffered solutions resulting in the same desired precipitate.

6 Conclusions

This chapter aimed to provide the reader with a comprehensive introduction to


the GDEx process and to spark an interest in the topics that are the focus of the
remainder of this PhD thesis. The apparent homogeneous nature of the batch
GDEx process and its similarities with AT were proven and demonstrated by
pH and Cu2+ concentration measurements. Chapter 4 will elaborate further on
the shape of the pH curves and the parameters that influence the shape. Besides
the obvious similarities the GDEx process was shown to exhibit significant
delays compared to the AT process, in terms of the start and finish of the metal
removal/recovery process. These delays are discussed in detail in Chapter 7.
GDEx using an air-fed GDE (driving the oxygen reduction reactions) to produce
OH− demonstrated to also produce significant amounts of H2 O2 . This H2 O2
increases the redox potential of the bulk solution and allows metal ions to be
oxidized even though they are in the catholyte. This is in stark contrast with an
alternative electrochemical reaction for the production of OH− , the hydrogen
evolution reaction(HER), which results in very low redox potentials, as would
be expected in the catholyte. The effect of H2 O2 on the characteristics of
metal (oxy)(hydr)oxide nanoparticles is elaborated on in Chapter 5 and how
this translates to GDEx as a nanoparticle synthesis process is demonstrated in
Chapter 6.
AT, and by extension GDEx, was demonstrated to be a process that can
be used universally for metal ions that can form sparingly soluble metal
(oxy)(hydr)oxides or metal hydroxysalts. A simple model to estimate selective
removal of metal hydroxides was presented and GDEx was used to demonstrate
two examples, one where selectivity was possible and one where selectivity
proved difficult.
To conclude the introduction, GDEx was discussed in terms of energy
consumption. The oxygen reduction reactions of GDEx using an air-fed GDE
demonstrated to be much less energy-intensive than the hydrogen evolution
reaction at a Pt electrode for similar production rates of OH− . This gives GDEx
an edge over this competitor both as a metal removal/recovery process, and as
a nanoparticle synthesis process.
Chapter 4

Metal removal from aqueous


solutions: insights from
modeling precipitation
titration curves

This chapter aims to demonstrate the important parameters that affect


the shape of the time-dependent pH curves during the GDEx process
(hereinafter referred to as precipitation titration curves or PTCs).
Alkaline titration (AT) was used to investigate these parameters because
it is similar to GDEx. ZnCl2 solutions were chosen because the Zn2+ ion
is not affected by the oxidants produced in the GDEx process. Therefore,
AT and GDEx are practically equivalent processes and by investigating
AT instead of GDEx, many of the complexities related to the flow-cell
design, the porous electrode and the production of oxidants can be avoided.
This allows a focus on the key parameters of alkaline precipitation as a
physicochemical process.

Contributions of the author were hypothesis development, development of the


analytical and numerical models for the time-dependent pH curves, carrying
out all of the experiments, characterizations, and data treatment. R. Prato
helped with the initial stages of hypothesis development and model development,
discussions, and proof-reading. X. Dominguez Benetton and J. Fransaer helped
with funding, discussions, and proof-reading.

45
46 METAL REMOVAL FROM AQUEOUS SOLUTIONS: INSIGHTS FROM MODELING PRECIPITATION
TITRATION CURVES

This chapter has been published in Journal of Environmental Chemical


Engineering as: Eggermont S.G., Prato R., Dominguez Benetton X., Fransaer
J. Metal removal from aqueous solutions: insights from modeling precipitation
titration curves, J. Environ. Chem. Eng (2019). doi: 10.1016/j.jece.2019.103596
The authors thank dr. Arturo de Jesús Garcı́a Mendoza for a useful discussion
and proofreading.
METAL REMOVAL FROM AQUEOUS SOLUTIONS: INSIGHTS FROM MODELING PRECIPITATION
TITRATION CURVES 47

Abstract

To provide a good reference for many metal ion removal processes, we developed
an analytical model (AM) and used the PHREEQC computer program to
numerically model (NM) arbitrary metal hydroxide precipitation titration curves
(PTCs) and compared these with slow titration experiments of ZnCl2 with
NaOH to precipitate Zn(OH)2 for validation. The AM effectively demonstrates
how initial pH, metal ion concentration and - valence and metal hydroxide
equilibrium constant affect the shape of the PTCs. A comparison of AM and
NM exposed AM limitations that include the effect of NaCl concentration,
activity coefficients, and multi-species ions, on the shape of the PTCs. The
experimental results agree well with either model for different concentrations of
ZnCl2 at slow titration rates. However, the models and experiments deviated
at higher titration rates. These deviations were attributed to inhomogeneities
in the solution concentration, which caused precipitation at a bulk pH below
the theoretical precipitation pH (i.e., pre-precipitation).
48 METAL REMOVAL FROM AQUEOUS SOLUTIONS: INSIGHTS FROM MODELING PRECIPITATION
TITRATION CURVES

1 Introduction

Heavy metal contamination of soil, ground - and drinking water, and food
is of major global concern [91, 114, 115]. Chemical precipitation, chemical
coagulation/flocculation, electrochemical methods, membrane filtration, ion-
exchange, bioremediation, and adsorption are the most current techniques for
heavy metal removal [91–96]. Metal hydroxide precipitation is one of the most
frequently used chemical precipitation methods to remove heavy metal ions from
aqueous solutions [98, 116–118]. Most of the existing research on metal removal
with metal hydroxide precipitation focuses on the end-result (i.e., effective metal
removal) but there is a dearth in the literature on the path towards metal
removal using metal hydroxide precipitation. Information on the path would
help in the comparison of different precipitation techniques. To bridge this
knowledge gap, an in-depth investigation of the metal hydroxide precipitation
path imposed itself. Such investigation was the goal of this work and was carried
out through the in-depth analysis of modeled and experimental precipitation
titration curves. The aim was to demonstrate easy-to-use models and to provide
an ideal case for future reference.
Precipitation titrations are traditionally used for the determination of halide
concentrations [119]. However, their usefulness has also been demonstrated for
metal hydroxide precipitation, for example when used for the characterization
of acid mining waters [120, 121].
During the metal hydroxide precipitation titration, the pH is measured and
plotted as a function of the added amount of precipitant (e.g., NaOH). Such a
pH plot can act as the blueprint of the metal removal path because the proton
concentration is indirectly related to the metal ion concentration via chemical
equilibria. A mathematical model of such precipitation titration curve can
demonstrate the effect of different parameters on the precipitation behavior.
Some authors have presented extensive work on the mathematical modeling of
titration curves, but in such cases, precipitation titrations are generally left out
of scope, or it is not the pH that is being modeled [122, 123].
Several parameters were identified as important: (1) the initial pH, (2) the initial
metal ion concentration, (3) the metal ion valence, (4) the metal hydroxide
equilibrium constant, (5) the supporting electrolyte concentration, (6) the
activity coefficients, and (7) the multi-species ions. We developed an analytical
model (AM) because its simple equations reveal the effect of the parameters
that most strongly influence the major features of the titration curve (i.e.,
parameters 1 to 4). We developed a numerical model (NM) with PHREEQC
that was able to include the effect of all parameters (i.e., parameters 1 to 7). A
comparison of AM and the NM demonstrated how parameters 5 to 7 influenced
EXPERIMENTAL 49

the shape of the precipitation titration curves. To demonstrate its usefulness,


the NM was compared to the experimental results of a chosen reference system:
the precipitation of Zn(OH)2 via the titration of ZnCl2 solutions with NaOH.

2 Experimental

Different 100 mL ZnCl2 solutions, supplemented with or without 0.500 M NaCl,


were titrated with a total of 20 mL of a 0.1 M NaOH solution (Titrisol , ®
Merck), at different titration rates (i.e., 2 mL min−1 , 0.2 mL min−1 , and 0.02
mL min−1 ). The slowest rate of 0.02 mL min−1 , corresponding to an addition
rate of 33 10−9 mol NaOH s−1 , resulted in titration experiments of 1000 minutes.
The reactor was a 250 mL DURAN laboratory bottle, with additional flask
openings to allow the introduction of a pH-meter and a titrator tip. The cap was
supplemented with openings for argon purging (figure 3.2c). To approximate
perfect mixing, the solution was stirred with a magnetic stirring bar of 3 cm on
an IKA RCTbasic S1 stirrer at rotation level 7 (approximately 700-800 rpm). A
stock solution of 20 mM ZnCl2 was produced by diluting a 1M ZnCl2 (65.38 g
L−1 anhydrous ZnCl2 , Chem-Lab) with demineralized water. The solutions of
1.0 mM, 2.0 mM, and 5.0 mM ZnCl2 without 0.500 M NaCl were made simply
via dilution of the stock solution with demineralized water. The solutions with
0.500 M NaCl were made by taking the appropriate amount of stock solution,
adding the appropriate amount of NaCl (analytical grade, Fischer-Scientific)
and thereafter adding the appropriate amount of demineralized water. The

®
initial pH of the solutions (i.e., pH = 3) was obtained via small additions (i.e.,
approximately 0.1 mL) of 1 M HCl solution (Titrisol , Merck). All experiments
were carried out at room temperature (i.e., 22 ◦ C ± 2 ◦ C). The pH was measured
every 5 seconds with a Metrohm 781 pH/ion meter equipped with a Metrohm
Unitrode pH electrode. The electrode was calibrated with 4 Merck standard
solutions (pH 4.01, pH 7.00, pH 10.00, and pH 12.00).

3 Models

3.1 Basic equations

The titration curve models in this work aim to simulate a traditional volumetric
titration (figure 3.2c). In volumetric titrations, a reagent is slowly added to the
analyte. This overall process is equivalent to a semi-batch reaction because the
total analyte volume increases (slightly) over time. Therefore, by extension, the
titration model is also a semi-batch reactor model.
50 METAL REMOVAL FROM AQUEOUS SOLUTIONS: INSIGHTS FROM MODELING PRECIPITATION
TITRATION CURVES

Figure 4.1 summarizes the stages of metal hydroxide precipitation. It only shows
the most important species. From the shape of the titration curve, one can
easily distinguish three stages during volumetric titration (figure 4.1b). Stage 1
before precipitation, stage 2 during precipitation (i.e., the so-called precipitation
plateau) and stage 3 after precipitation.
The titration model is calculated in two steps: first, the mass and charge
balances are solved to give time-dependent values of the concentrations and
activity coefficients of the protons and other species, then, the pH is calculated
as a function of the proton activity:

pH = − log (aH+ ) = − log γH+ H+


 
(4.1)
 +
with H [mol L−1 ] the proton concentration, CNaOH [mol L−1 ] the titrant
concentration, Vadded [L] the added titrant volume, rV [L min−1 ] the titration
rate, t the titration time [min], aH+ [mol L−1 ], and γH+ [-] the proton activity
coefficient. Plotting the pH as a function of the added precipitant results in the
titration curve (figure 4.1b).

(a) (b)

Figure 4.1: (a) Simplified scheme of the different stages of the titration process.
The most prominent species at each stage are indicated. Upwards or downwards
arrows respectively indicate a strong concentration increase or decrease. (b)
An arbitrary example of a precipitation titration curve with the corresponding
three stages.

In the PHREEQC manual, Parkhurst et al. nicely elaborate on the basic


equations for chemical equilibria in solution [124]. Here, we demonstrate the
important equations for modeling titration curves, using their notation. We will
describe the model in terms of zinc chloride, but it is easily adapted to other
systems.
MODELS 51

Table 4.1: The different solution species that are considered in the ZnCl2
reference system (minteqv4.dat database [125,126]), and corresponding values for
the equilibrium constants Ki (two or three significant figures are reprinted) [126]
Master H log Ki Zn log Ki Na log Ki Cl log Ki
species
Solution H+ 1.0 Zn2+ 1.0 Na+ 1.0 Cl− 1.6
species OH− 14 ZnCl+ 0.40 ZnCl+ 0.40
H2 O 1.0 ZnCl2 0.60 ZnCl2 0.60
ZnCl−
3 0.50 ZnCl−
3 0.50
ZnCl2−
4 0.20 ZnCl2−
4 0.20
ZnOHCl -7.5 ZnOHCl -7.5
ZnOH+ -9.0
Zn(OH)2 (aq) -17.8
Zn(OH)−3 -28.1
Zn(OH)2−
4 -40.5

In the precipitation reactor, solid and liquid are in (dynamic) equilibrium. The
species in solution are in equilibrium with the species in the solid. To capture the
conservation of mass for each species across the two phases, we define 4 master
species in our system (i.e., H, Zn, Na, and Cl) [124]. All (non-master) solution
species are defined in terms of four solution master species (i.e., respectively
H+ , Zn2+ , Na+ , and Cl− ) [124]. For example, a reaction for ZnCl− 3 would be
defined as Zn2+ + 3 Cl− = ZnCl− −
3 instead of as ZnCl2 + Cl = ZnCl3 . For

our zinc chloride reference system the different solution species for each master
species are given in table 4.1. We also define pure phases (i.e., solids) that are
in equilibrium with the solution (i.e., Zn(OH)2 (s)).
Each solution species i is defined by its activity [124]:

ai = γi mi (4.2)

with ai [mol kg−1 ] the activity of species i, γi [-] the activity coefficient of
species i, and mi [mol kg−1 ] the molality of species i and its total number of
moles in solution [124]:
ni = mi Waq (4.3)
with ni [mol] the total number of moles of species i, and Waq [kg] the total mass
of solvent water in solution [124]. For each aqueous species i, a mass-action
equation is defined [124]:
Maq
Y
Ki = ai a−c
m
m,i
(4.4)
m

with Ki the temperature-dependent equilibrium constant of species i, cm,i the


stoichiometric coefficient of master species m in species i (i.e., equal to zero if
52 METAL REMOVAL FROM AQUEOUS SOLUTIONS: INSIGHTS FROM MODELING PRECIPITATION
TITRATION CURVES

species m is not a part of species i, negative if species m occurs on the right-hand


side of the reaction, and positive if species m occurs on the left-hand side of
the equation), and Maq is the total number of aqueous master species [124]. In
both our AM and the NM, equilibrium is assumed at every time. This requires
that all mass-action equations for the species in solution are satisfied at all
times. The mass-action equation for the reaction of ZnCl− 3 (i.e., Zn
2+
+ 3 Cl−

ZnCl3 ) is for example given by:


aZnCl−
3
KZnCl− = 3 (4.5)
3
aZn2+ (aCl− )

The total number of moles ni of a solution species i is calculated using the


mass-action equation (equation (4.4)) [124]:

Maq
Y
acmm,i
m
ni = mi Waq = Ki Waq (4.6)
γi

The activity ap of a pure phase p is equal to 1, the mass-action equations reduce


to [124]:

Maq
Y
Kp = acmm,p (4.7)
m

with cm,p the stoichiometric coefficient of master species m in the dissolution


reaction (i.e., negative if species m occurs on the left-hand side and positive if
species m occurs on the right-hand side). For Zn(OH)2 (s) (i.e., Zn(OH)2 (s) +
2H+
Zn2+ + 2 H2 O) it is for example given by:
2
aZn2+ (aH2 O )
KZn(OH)2 (s) = 2 (4.8)
(aH+ )

Remark that this definition differs from the definition of the equilibrium product
constant of metal hydroxides given in other parts of this PhD thesis. More
commonly, the equilibrium product constant is defined as KZn(OH)2 (s),common =
2
aZn2+ (aOH− ) . The other definition here is needed because all species are
referenced against H+ in PHREEQC. Of course, the common definition and
the definition of PHREEQC are intricately linked via the water equilibrium:
2
KZn(OH)2 (s),common = KOH − KZn(OH) (s) .
2
MODELS 53

The mass balance equation for each species m is written as [124]:


Naq Np
X X
fm = Tm − bm,i ni − bm,p np (4.9)
i p

with the function fm = 0 when the mole balance is achieved, Tm [mol] the
number of moles of master species m, Naq , and Np the number of aqueous
species and the number of solid phases (i.e., precipitates) respectively, bm the
moles of master species m in one mole of a given entity i or p. bm is usually,
but not always, equal to cm (the coefficient of the master species for m in
the mass-action equations) [124]. The charge balance, in our case, is written
as [124]:
Naq
X
fz = − zi ni (4.10)
i
with the function fz = 0 when the charge balance is achieved, and zi equal to
the charge of species i [124]. With all the above equations, it is possible to find
both an analytical or numerical solution for the proton concentration and to
solve equation (4.1).

3.2 Analytical model

Generally, numerical modeling is the go-to option for solving problems dealing
with metal ions in solution because of the large number of variables and
their mathematical interdependence. The added value of the analytical
model lies in its simplicity. It can easily demonstrate the effect of the most
important parameters on the path towards metal hydroxide precipitation. These
parameters are the initial solution pH, initial metal ion concentration, metal
ion valence, and metal hydroxide equilibrium constant. Six assumptions are
necessary to obtain an analytical solution to the system of mass- and charge
balance equations: (1) consider a limited amount of species, (2) assume constant
activity coefficients, (3) the values of all fi are equal to zero, (4) introduce the
precipitation pH as a new parameter, (5) use the charge balance as the master
equation, (6) define Zn2+ as a piece-wise function in relation to the precipitation
pH.
Assumption 1 states to only consider a few of the species from table 4.1(i.e., five
in total: H+ , OH− , Na+ , Cl− , Zn2+ ) and one pure phase (i.e., Zn(OH)2 (s)). In
essence, the former is equivalent to combining all Zn-containing solution species
into a single species Zn2+ . It should be evaluated separately for every system
(e.g., if other supporting electrolytes would be considered such as NaNO3 and
Na2 SO4 ) if this assumption is reasonable. If the assumption would appear
54 METAL REMOVAL FROM AQUEOUS SOLUTIONS: INSIGHTS FROM MODELING PRECIPITATION
TITRATION CURVES

to be unreasonable, the results should be discussed cautiously. In such a


case, a large deviation is expected between the analytical model and both the
numerical model and the experimental results. Furthermore, in most cases
where anions different from the OH− ion build into the pure phase (e.g., metal
hydroxychloride, metal hydroxysulfate), the model cannot be solved analytically.
It strongly depends on the composition of the hydroxysalt whether or not the
model remains analytically solvable. Figure 4.2 demonstrates a PHREEQC
calculation of the speciation as a function of the pH for a system with an initial
ZnCl2 concentration of 5 mM, with and without 0.5 M NaCl. It confirms the
low degree of complexation in the case without 0.5 M NaCl and the higher
degree of complexation in the case with 0.5 M NaCl.

-2 -2
totZn(aq)
tot Zn(aq) tot
totZn(aq)
Zn(aq)
-3 Zn2+
Zn(2+) -3 Zn(2+)
Zn2+
log(Concentration [mol L-1])

log(Concentration [mol L-1])


Zn(OH)2(s)
Zn(OH)2(s) Zn(OH)2(s)
Zn(OH)2(s)
-4 ZnCl+
ZnCl(+) -4 ZnCl(+)
ZnCl+
ZnCl2
ZnCl2 ZnCl2
ZnCl2
-5 -5
ZnCl3-
ZnCl3(-) ZnCl3-
ZnCl3(-)
-6 ZnCl42-
ZnCl4(2-) -6 ZnCl42-
ZnCl4(2-)
ZnOH+
ZnOH(+) ZnOH+
ZnOH(+)
-7 Zn(OH)2 (aq)
Zn(OH)2 -7 Zn(OH)2
Zn(OH)2 (aq)
Zn(OH)3-
Zn(OH)3(-) Zn(OH)3-
Zn(OH)3(-)
-8 Zn(OH)42- -8 Zn(OH)42-
Zn(OH)4(2-)
Zn(OH)4(2-)
-9 ZnOHCl
ZnOHCl -9 ZnOHCl
ZnOHCl

-10 -10
3 4 5 6 7 8 9 10 11 12 3 4 5 6 7 8 9 10 11 12
pH pH

Figure 4.2: Speciation in function of pH for an aqueous system with Zn-


containing species with an initial concentration of 5 mM ZnCl2 without (left)
and with (right) 0.5 M NaCl.

Assumption 2 states that all the activity coefficients should be constant, instead
of concentration-dependent. Without this assumption, the system of equations
cannot be solved analytically. For very dilute systems, a value of 1 for the activity
coefficients is acceptable. At higher concentrations, the activity coefficients are
often very different from 1 and it is more suitable to use other values than 1.
Assumption 3 states that all the values of fi (equation (4.9)) should be fixed
and equal to zero. These functions only have non-zero values in the numerical
model (i.e., their values are iteratively decreased until a certain precision is
reached).
Assumption 4 introduces a new parameter, the so-called precipitation pH.
The precipitation pH, pHprec , is the lowest pH value at which M(OH)n starts
MODELS 55

precipitating and it depends on the initial concentration of the metal ion.


Exemplified for Zn(OH)2 (s) as:

2
!
γZn2+ Zn2+ (aH2 O )
 
1
pHprec = − log (4.11)
2 KZn(OH)2 (s)
with Zn2+ the concentration of Zn2+ [M].
 

Assumption 5 states the charge balance as the main equation:


Na + H + 2 Zn2+ = OH− + Cl−
 +  +      
(4.12)
If a species can be written as a function of the proton concentration via its mass
action equation (i.e., OH− ), it is substituted as such into the charge balance:

 K − aH2 O KOH− aH2 O


OH− = OH

= (4.13)
γH+ H+ γOH−
 
aH+ γOH−
If a species cannot be written as a function of the proton concentration (i.e.,
Na+ , Cl− ), it is substituted into the charge balance via its mass balance equation
:

 + TNa (CNaCl V0 + CNaOH rV t)


Na = = (4.14)
V0 + rV t V0 + rV t

 − TCl (CNaCl V0 + 2CZnCl2 V0 + CHCl V0 )


Cl = = (4.15)
V0 + rV t V0 + rV t
with V0 [L] the initial volume of the solution, rV [L min−1 ] the titration rate,
t the titration time [min], CNaCl [mol L−1 ] the initial NaCl concentration,
CNaOH [mol L−1 ] the titrant concentration, CHCl [mol L−1 ] the initial HCl
concentration, and CZnCl2 [mol L−1 ] the initial ZnCl2 concentration.
Assumption 6 defines the concentration of Zn2+ as a piece-wise function. For a
pH < pHprec , its value is substituted by its initial value and for a pH ≥ pHprec ,
its value is substituted via its mass action equation:

 2+  TZn CZnCl2 V0
Zn = = (pH < pHprec ) (4.16)
V0 + rV t V0 + rV t

 2
 2+  KZn(OH)2 (s) (aH+ )2 KZn(OH)2 (s) γH+ H+
Zn = 2 = 2 (pH ≥ pHprec )
(aH2 O ) γZn2+ (aH2 O ) γZn2+
(4.17)
56 METAL REMOVAL FROM AQUEOUS SOLUTIONS: INSIGHTS FROM MODELING PRECIPITATION
TITRATION CURVES

The piece-wise definition of Zn2+ , implies the piece-wise definition of the


 

charge balance equation. After substitution of equations (4.13)-(4.17) into


equation (4.12), the system of charge balance equations can be reduced to:

 + 2 CHCl V0 − CNaOH rV t  +  KOH− aH2 O


H − H − =0 (pH < pHprec ) (4.18)
V0 + rV t γH+ γOH−
3
n (γH+ ) KZn(OH)2 (s)  3  2
2 H+ + H+
(aH2 O ) γZn2+

CHCl V0 + 2CZnCl2 V0 − CNaOH rV t  +  (4.19)


− H (pH ≥ pHprec )
V0 + rV t
KOH− aH2 O
− γH+ γOH− =0

This system of equations in its implicit form represents H+ as a function


 

of CNaOH rV t. A general analytical solution to these equations exists because


we used a 2-valence ion (i.e., Zn2+ ). An analytical solution is available only
for metal ions with a valence up to 3 because for higher valences equation
(4.19) would result in a fifth-order polynomial,
  without a general analytical
solution [127]. The analytical solution of H+ as a function of CNaOH rV t is
then used to calculate the pH (equation (4.1)). The initial pH (i.e., the pH at t
= 0) can thus, as an example, be calculated from equation (4.18) as:

q
K a
 2

CHCl + (CHCl ) + 4 γOH+−γ H2−O
H OH
pHi = − log   ≈ − log (CHCl ) (4.20)
2

This demonstrates the straightforward relationship between the initial pH and


the initial concentration of hydrochloric acid. This straightforward relationship
is the main reason to choose a monoprotic acid in our reference case. For
polyprotic acids, this relationship is often less straightforward and complicates
or impedes an analytical solution to our problem.

3.3 Numerical model

The numerical model is made with PHREEQC and the reader is guided to
the manual, where a very detailed description of the implementation can be
found [124].
RESULTS AND DISCUSSION 57

The most important functions fi that are solved by PHREEQC were shown
before in the basic equations section. The numerical model is solvable without
the assumptions made for the analytical model. PHREEQC can take into
account all species in table 4.1 and it can calculate all of the concentration-
dependent activity coefficients. The functions fi are iteratively solved via a
Newton-Raphson method. The PHREEQC algorithms run according to the
instructions of a user-defined input file. The complete input file of the numerical
model implementation with PHREEQC is found in appendix B.
Hereunder follows the modeling sequence in words. A certain solution with
an initial composition is chosen (i.e., an initial volume of water, a given initial
metal ion concentration, a given initial NaCl concentration, and initial acid
concentration). PHREEQC takes the thermodynamic data (e.g., the equilibrium
constant for the mass-action equation or some parameters to help solve the
activity coefficients) of the species in solution (i.e., all the species containing
only those elements provided in the initial composition) from the provided
database. PHREEQC iteratively calculates the initial number of moles of the
different species in solution. This determines the simulated initial experimental
electrolyte composition. From this point on, a chosen titrant solution (i.e., 0.1
M NaOH) is added in a fixed amount of discrete steps. The same iterative
process is used after every addition step. For each step, the electrolyte pH is
plotted as a function of the added amount of NaOH and the corresponding data
points are stored.

4 Results and discussion

4.1 Taxonomy of the models

The analytical model (AM) and the numerical model (NM) were both approached
in two ways. The first version of the analytical model (AM1) assumes all activity
coefficients γi equal to 1. It is the most stripped version of a titration model and
is considered the baseline. The first version of the numerical model (NM1) takes
into account concentration-dependent activity coefficients but does not take into
account multi-species ions. NM1 can also calculate the concentration-dependent
activity coefficients at every titration step. For each species, a time-averaged
activity coefficient can be calculated. The second version of the analytical model
(AM2) uses these average values of the activity coefficients instead of a value
equal to 1. The second version of the numerical model (NM2) takes into account
concentration-dependent activity coefficients as well as multi-species ions.
58 METAL REMOVAL FROM AQUEOUS SOLUTIONS: INSIGHTS FROM MODELING PRECIPITATION
TITRATION CURVES

Analysis of AM1 demonstrates the effect of the most important parameters:


the initial acid concentration, the initial metal ion concentration, the metal ion
valence, and the equilibrium constant on the different stages of the precipitation
titration process. A comparison of AM1 and NM1 demonstrates the effect of the
activity coefficients. A comparison of AM1, AM2, NM1, and NM2 demonstrates
the effect of concentration-dependent activity coefficients and the effect of multi-
species ions and it provides the complete picture, demonstrating the effect of
all parameters.

4.2 Effect of C HCl , C MCln , n, and KM(OH)n

Figure 4.3 shows the results of AM1 and the effect of the initial acid concentration
(i.e., initial pH, equation (4.20)), the metal ion concentration, the metal ion
valence, and the equilibrium constant on the different stages of the precipitation
titration process. In every titration curve, the three stages are apparent. Stage
1 is mathematically described by equation (4.18), while stage 2 and stage 3 are
described by equation (4.19).
Stage 1 is determined by the initial HCl concentration (i.e., initial pH, equation
(4.20)), see figure 4.3a. Stage 1 ends when CHCl V0 − CNaOH rV t reaches a value
of 0 (equation (4.18)). Stage 2 is called the precipitation plateau and is the most
important stage during the precipitation titration process. Stage 2 is determined
by its length and its (initial) height. Its length is influenced by the initial metal
ion concentration and the metal ion valence, see figure 4.3b and figure 4.3c,
respectively. It ends when CHCl V0 + 2CZnCl2 V0 − CNaOH rV t reaches a value of
0. Its initial height is equal to the value of the precipitation pH (see equation
(4.11)). The initial height is influenced by the metal ion valence (i.e., a lower
valence gives a higher plateau height, figure 4.3c), the metal ion concentration
(i.e., a lower concentration gives a higher plateau height, figure 4.3b) and metal
hydroxide equilibrium constant (i.e., a higher value gives a higher plateau height,
figure 4.3d). Because of our assumptions, stage 3 is not significantly influenced
by the chosen parameters. To avoid the redissolution of formed precipitates at
higher pH, most metal removal or metal recovery processes should choose an
endpoint somewhere at the beginning of stage 3. This is also clear from figure
4.2, which demonstrates an increase in the concentration of soluble Zn(OH)− 3
and Zn(OH)2− 4 at high pH.

The process costs of a batch metal removal process is strongly correlated with
the amount of NaOH necessary to reach the second inflection point. Therefore,
because of their effect on the position of the second inflection point, the initial pH,
the metal ion concentration, and the metal ion valence are the most important
parameters concerning cost predictions.
RESULTS AND DISCUSSION 59

14 14
(a) (b)
12 12

10 10

8 8

pH
pH

6 6

4 4

2 pH 2 2 1 mM
pH 3 2 mM
pH 4 5 mM
0 0
0 0.0005 0.001 0.0015 0.002 0 0.0005 0.001 0.0015 0.002
NaOH added/ CNaOH rV t [mol] NaOH added/ CNaOH rV t [mol]
14 14
(c) (d)
12 12

10 10

8 8
pH
pH

6 6

4 4

2 2 2 10
3 11
4 12
0 0
0 0.0005 0.001 0.0015 0.002 0 0.0005 0.001 0.0015 0.002
NaOH added/ CNaOH rV t [mol] NaOH added/ CNaOH rV t [mol]

Figure 4.3: The influence of different parameters on metal hydroxide


precipitation titration curves modeled with analytical model AM1. Unless
otherwise indicated between brackets, parameter values are V0 = 100 mL,
CMCln = 2.0 mM, n = 2, KM(OH)2 = 1012 , KW = 10−14 , pHi = 3, CNaOH =
0.1 M, and all activity coefficients = 1. (a) influence of initial pH (pHi = 2, 3,
4), (b) influence of initial metal ion concentration (CMCln = 1 mM, 2 mM, 5
mM), (c) influence of valence (n = 2, 3, 4), and (d) influence of equilibrium
constant (log KM(OH)2 = 10, 11, 12).
60 METAL REMOVAL FROM AQUEOUS SOLUTIONS: INSIGHTS FROM MODELING PRECIPITATION
TITRATION CURVES

4.3 Effect of γi , C NaCl , and multi-species ions

The effect of activity coefficients γi , the initial NaCl concentration CNaCl , and
the multi-species ions was investigated by comparing the analytical model and
the numerical model. Figure 4.4 compares the four model versions (i.e., AM1,
AM2, NM1, NM2) for specific sets of variables.

14 14 14
AM1 no NaCl AM2 no NaCl AM2 NaCl
12 AM2 no NaCl 12 AM2 NaCl 12 NM1 NaCl
NM1 no NaCl NM2 no NaCl NM2 NaCl
10 NM2 no NaCl 10 NM2 NaCl 10

8 8 8
pH

pH

pH
6 6 6

4 4 4

2 2 2

0 0 0
0 0.0005 0.001 0.0015 0.002 0 0.0005 0.001 0.0015 0.002 0 0.0005 0.001 0.0015 0.002
NaOH added/ CNaOH rV t [mol] NaOH added/ CNaOH rV t [mol] NaOH added/ CNaOH rV t [mol]
(a) (b) (c)

Figure 4.4: Comparison of the analytical model and numerical model. The
parameter values are V0 = 100 mL, CMCln = 5.0 mM, n = 2, KM(OH)2 =
1012 , KW = 10−14 , pHi = 3, CNaOH = 0.1 M. The constant average activity
coefficients for AM2 without NaCl are γH+ = 0.89, γOH− = 0.88 and γZn2+ =
0.63. The constant average activity coefficients for AM2 with NaCl are γH+ =
0.82, γOH− = 0.62 and γZn2+ = 0.23. (a) Influence of the activity coefficients
(equal to 1 for AM1, different from 1 for AM2, NM1, and NM2, see text for
more details) and influence of multi-species ions (species included in NM2, not
in AM1, AM2, and NM1) for electrolytes without NaCl. (b) Influence of NaCl
and its effect on the influence of the activity coefficients. (c) Influence of the
multi-species ions in solutions with NaCl.

In light of the complete precipitation titration curve, the models differ only
slightly. The inserts of figure 4.4 demonstrate the subtle differences. The largest
difference exists between AM1 and the other three models (figure 4.4a) because
AM1 does not take into account activity coefficients. This demonstrates that
the activity coefficients play a certain role, albeit much less pronounced than
the role of the parameters discussed in the previous section. This is especially
clear, taking into account that using average activity coefficients (AM2) almost
completely resolves the difference with the numerical model. A comparison of
NM1 and NM2 demonstrates the rather insignificant effect of multi-species ions
RESULTS AND DISCUSSION 61

in solutions without supporting electrolyte. The comparison of AM2 and NM2


for solutions without NaCl and with 0.5 M NaCl demonstrates the effect of
NaCl on the precipitation titration curves (figure 4.4b). The addition of NaCl
shifts the onset of precipitation to higher amounts of NaOH addition. This is
the result of lower activity coefficient values at higher ion concentrations. This
also shifts the precipitation pH upwards. The addition of NaCl increases the
curvature at the beginning of the precipitation plateau (NM2 no NaCl vs. NM2
NaCl). Because AM2 only takes into account the activity coefficients, but not
the multi-species ions, it overestimates the concentration of Zn2+ . At higher
salt concentration, as discussed before, the concentration of these multi-species
ions is higher. Therefore, the concentration of free Zn2+ is a little bit lower for
NM2 than for AM2. Figure 4.4c demonstrates a slightly higher precipitation
pH for NM2 in comparison to AM2. NM1 and AM2 do not significantly differ
in precipitation pH, which demonstrates the role of the multi-species ions on
the precipitation pH.
The results here demonstrate that the effect of the activity coefficients, the
NaCl concentration, and the multi-species ions on the shape of the precipitation
titration curves is relatively limited in comparison to the effect of the initial
pH, the metal ion concentration, the metal ion valence, and the equilibrium
constant. The exception being the rather significant increase in the value of the
precipitation pH when NaCl is added to the solution compared to it not being
added to the solution, attributed to the significant decrease in activity coefficient
of Zn2+ in the case of high NaCl concentration (i.e., average γZn2+ ,NaCl = 0.23)
compared to the case without NaCl (i.e., average γZn2+ ,no NaCl = 0.63). Within
the limitations posed for the analytical model, we can conclude the analytical
models and numerical models result in quasi identical precipitation titration
curves.

4.4 Comparison of the numerical model with the experiments

NM2 was compared to the experimental results of a reference system. NM2


was chosen because it is the most complete model, but as previously discussed,
the four models result in quasi identical precipitation titration curves. The
reference system of choice was the precipitation titration of ZnCl2 with NaOH
to precipitate Zn(OH)2 . The effect of initial metal ion concentration (figure 4.5),
titration rate (figure 4.6), and NaCl concentration (figure 4.7) are demonstrated.
The deviation of the model and the experiments are discussed in detail.
Good agreement is obtained between NM2 and experiments for 100 mL ZnCl2
solutions of different concentrations (i.e., 1.0 mM, 2.0 mM, and 5.0 mM), see
figure 4.5. To allow for this good agreement, an equilibrium constant of value
62 METAL REMOVAL FROM AQUEOUS SOLUTIONS: INSIGHTS FROM MODELING PRECIPITATION
TITRATION CURVES

1012 is used, rather than the, by the Minteq v4 database, provided value of
1011.334 . This confirms the need, expressed by the authors of the PHREEQC
manual [124], to cautiously use the constants provided in the database and
where necessary verify them experimentally.

14

12

10

8
pH

6
Exp. 1 mM
4 NM2 1 mM
Exp. 2 mM
2 NM2 2 mM
Exp. 5 mM
0 NM2 5 mM
0 0.0005 0.001 0.0015 0.002
NaOH added / CNaOHrVt [mol]

Figure 4.5: Comparison of the numerical model (NM2) to volumetric titration


experiments of electrolytes with different ZnCl2 concentrations. The parameter
values are V0 = 100 mL, rV = 0.02 mL min−1 , n = 2, KZn(OH)2 = 1012 , pHi =
3, CNaOH = 0.1 M. Inset demonstrates the supersaturation bumps.

A local pH maximum, similar to those observed for slow (homogeneous)


precipitation with urea [128, 129], is observable at the onset of precipitation for
each of the different concentrations. Most probably, this behavior originates
from supersaturation conditions. To initiate precipitation, stable nuclei need to
form and this requires an amount greater than the theoretical amount of OH−
needed to initiate precipitation. When the nuclei suddenly form, the excess OH−
ions react away until equilibrium conditions are reestablished. This causes a
local maximum in the pH. Because the difference in free OH− ion concentration
between pH 7 and 8 is very small (i.e., between 10−7 M and 10−8 M), the pH
meter is a very sensitive instrument for the detection of supersaturation of metal
hydroxides.
At higher titration rates, the supersaturation maximum disappears. Model and
experiment increasingly deviate. Figure 4.6 demonstrates this for the cases of
RESULTS AND DISCUSSION 63

5 mM ZnCl2 with titration rates of 0.02 mL min−1 , 0.2 mL min−1 and 2 mL


min−1 . Precipitates are visible in solution before the precipitation plateau is
reached. We define this subsaturated precipitation as pre-precipitation. We
suggest that pre-precipitation is caused by local supersaturation. This can
be explained through consideration of the mixing process. With increasing
titration rates, equal mixing conditions (i.e., the same magnetic stirrer, the
same rotation rate in the same reactor) result in increasingly less homogeneous
solutions. At the titrator tip, where NaOH is fed, a zone of higher pH forms.
If the local pH in such a zone exceeds the precipitation pH, pre-precipitation
will occur and Zn(OH)2 is formed. Ideally, these pre-precipitated Zn(OH)2
should redissolve when reaching subsaturated zones in the solution. However,
redissolution is a heterogeneous process and inherently slow. The precipitates
will therefore not redissolve instantaneously. Moreover, redissolution will be
slower if the value of the bulk pH is close to the precipitation pH because
of a small driving force for redissolution. The amount of OH− that reacts
in the pre-precipitation does not contribute to an increase in pH. At higher
titration rates, it is expected that more Zn(OH)2 pre-precipitates. Therefore, a
slower pH rise is expected for higher titration rates. This is obvious from our
experimental precipitation titration curves (figure 4.6). It is possible to make a
first-order approximation of the amount of metal ions that is pre-precipitated,
by subtracting the theoretical (modeled) charge balance equation from the
experimental charge balance equation (equation (4.12)). With our assumption
that the chlorine and sodium ions do not participate in the precipitation reaction,
and assuming that at pH below 7 the OH− concentration is negligible, the
difference between the theoretical and experimental metal ion concentration is:
 +
H exp − H+ theory
 
 n+   n+ 
= Mn+ pr−pr
 
M theory − M exp = (4.21)
n
with Mn+ pr−pr the amount of metal ions pre-precipitated as M(OH)n . Having
 

both the data of the modeled pH and experimental pH (i.e., the modeled
and experimental proton activities), it is possible to estimate the amount
of pre-precipitated metal ions. From figure 4.6, the estimated amount of
metal precipitated for 2 mL min−1 , 0.2 mL min−1 and 0.02 mL min−1 are
respectively 0.5 10−5 M, 0.5 10−6 M and 0.5 10−7 M. These are back of the
envelope calculations because it is difficult to determine the exact start of
pre-precipitation with experiments conducted for this study. Although the
calculations are only estimates, it is clear that a pH meter is a very sensitive
instrument to measure pre-precipitation, given it can detect pre-precipitated
amounts of 10−6 M and less. We leave to further developments the evaluation
of the time-dependent redissolution behavior.
64 METAL REMOVAL FROM AQUEOUS SOLUTIONS: INSIGHTS FROM MODELING PRECIPITATION
TITRATION CURVES

14

12

10

8
pH

4
NM2
NM2
2 0.02 mL min-1
0.02mL/min
0.2 mL min-1
0.2mL/min
0 mL min-1
22mL/min
0 0.0005 0.001 0.0015 0.002
NaOH added/ CNaOH rV t [mol]

Figure 4.6: Comparison of the numerical model (NM2) to volumetric titration


experiments at different titration rates. The parameter values are V0 = 100
mL, rV = 0.02 mL min−1 , CZnCl2 = 5 mM, n = 2, KZn(OH)2 = 1012 , pHi = 3,
CNaOH = 0.1 M.

At the end of precipitation (i.e., at the second inflection point), the model agrees
more with the experimental titration curves at faster precipitation rates than
with those of the 0.02 mL min−1 titration rate (figure 4.6). This anomaly was
also observed at different ZnCl2 concentrations (not shown here). It has yet
to be confirmed what causes this behavior, but we believe the long time-frame
(1000 min) causes disturbances in the pH measurement.
The influence of the NaCl concentration was also investigated. For industrial
wastewaters of low initial pH, the supporting electrolyte concentration can
become substantial when treated with NaOH. Figure 4.7 compares model NM2
with the experimental titration curves for electrolytes with and without 0.5 M
NaCl. In stage 3, the experimental pH is much lower in the case with NaCl than
in the case without NaCl. Although this is also observed by a comparison of
the models (figure 4.4b), the difference is much more significant experimentally.
This is attributed to the interference of sodium ions with the pH measurement,
the so-called alkali error [130]. At high sodium concentrations, and high pH
(i.e., low proton concentrations), the pH meter has difficulties in distinguishing
sodium ions from protons. This results in an overestimation of the proton
concentration (and thus the proton activity), which corresponds with a lower
RESULTS AND DISCUSSION 65

pH. Another important observation is the fact that in the case of 0.5 M NaCl, the
experimental precipitation plateau is significantly lower than that of the plateau
calculated by the model. It is practically equal to that of the case without 0.5
M NaCl. This might be the result of the choice of Pitzer coefficients for ZnCl2
solutions (see the PHREEQC input file in appendix B), taken from scientific
literature [131]. These coefficients might not be suitable in the concentration
range used, or as indicated before, might be unreliable. Additionally, at initial
precipitation, the supersaturation behavior in the case with and without NaCl
differs significantly. This might suggest the possible precipitation of a different
species, before the actual precipitation of Zn(OH)2 .

14

12

10

8
pH

4
Exp. no NaCl
2 Exp. NaCl
NM2 no NaCl
NM2 NaCl
0
0 0.0005 0.001 0.0015 0.002
NaOH added/ CNaOH rV t [mol]

Figure 4.7: Comparison of the numerical model (NM2) to volumetric titration


experiments for electrolytes with 0.5 M NaCl and without NaCl. The parameter
values are V0 = 100 mL, CZnCl2 = 5 mM, n = 2, KZn(OH)2 = 1012 , pHi = 3,
CNaOH = 0.1 M.
66 METAL REMOVAL FROM AQUEOUS SOLUTIONS: INSIGHTS FROM MODELING PRECIPITATION
TITRATION CURVES

5 Conclusions

Precipitation titration curves (PTCs) of metal hydroxides from chloride


solutions were analyzed in-depth using an analytical model, a numerical
model, and experimental data. A comparison of the analytical model and
the numerical model (made with PHREEQC software) demonstrated the effect
of 7 parameters on the shape of the PTCs: (1) the initial pH, (2) the initial
metal ion concentration, (3) the metal ion valence, (4) the metal hydroxide
equilibrium constant, (5) the supporting electrolyte concentration, (6) the
activity coefficients, and (7) the multi-species ions. The first four parameters
most strongly affect the general shape of the titration curves. The latter three
parameters affect the shape less strongly: they only added subtle changes to the
shape of the PTCs. A comparison of the numerical model with the experimental
results of our reference system (i.e., Zn(OH)2 precipitation from ZnCl2 solutions)
demonstrated good agreement for the major features of the PTCs. The
comparison also revealed the most important limitations of our models: the lack
of kinetic parameters. This was observable in certain phenomena: the occurrence
of supersaturation at low titration rates, and the occurrence of pre-precipitation
at high titration rates. Despite these limitations, the model demonstrated to
serve as a reference system for future investigations of precipitation processes.
Chapter 5

Oxidation-assisted alkaline
precipitation: effect of H2O2
on the size of CuO and
FeOOH nanoparticles

One of the objectives of this PhD was to identify the role of the GDEx
process regarding the oxidation of metal ions in the catholyte. Chapter 3
demonstrated the production of an oxidant, H2 O2 , during GDEx with
an air-fed GDE and the role of H2 O2 on increasing the redox potential.
As such, the oxidant was identified, but not yet its fundamental role in
the precipitation mechanism of the metal (oxy)(hydr)oxide nanoparticles.
This chapter clarifies this. An experimental procedure was specifically
designed to isolate the effect of H2 O2 on the precipitation behavior of
metal (oxy)(hydr)oxides. As such, the mechanism of oxidation-assisted
alkaline precipitation (Ox-AP) was discovered.

Contributions of the author were hypothesis development, literature investi-


gation and carrying out all experiments, some characterizations, and data
treatment. ICP-MS characterization and data treatment were carried out by A.
Rua-Ibarz and K. Tirez. SEM was done by Tom Van Der Donck. X. Dominguez
Benetton and J. Fransaer helped with funding, discussions, and proof-reading.

67
68 OXIDATION-ASSISTED ALKALINE PRECIPITATION: EFFECT OF H2 O2 ON THE SIZE OF CUO AND
FEOOH NANOPARTICLES

This chapter has been published as: Eggermont S.G., Rua-Ibarz A., Tirez K.,
Dominguez Benetton X., Fransaer J. Oxidation-assisted alkaline precipitation:
the effect of H2 O2 on the size of CuO and FeOOH nanoparticles. RSC Advances.
2019;9(51):29902-8. dio: 10.1039/C9RA03086G
A large part of the electronic supporting information of the original publication
was merged into the text, rather than leave it as an appendix, for a more unified
approach in the PhD thesis.
OXIDATION-ASSISTED ALKALINE PRECIPITATION: EFFECT OF H2 O2 ON THE SIZE OF CUO AND
FEOOH NANOPARTICLES 69

Abstract

H2 O2 was demonstrated to narrow the size distribution and decrease the size of
CuO and hydrous FeOOH (2-line ferrihydrite) nanoparticles at conditions of high
supersaturation. We introduce oxidation-assisted alkaline precipitation (Ox-AP)
and compare it to traditional alkaline precipitation (AP). While for AP, a metal
salt solution (e.g., CuCl2 ) is mixed with an alkali (e.g., NaOH), for Ox-AP, the
more reduced form of that metal salt solution (e.g., CuCl) is simultaneously
mixed with that alkali and an oxidant (e.g., H2 O2 ). The resulting precipitates
were characterized by SEM, XRD, DLS and single-particle ICP-MS and showed
to be nanoparticles (NPs). Ox-AP CuO NPs were up to 3 times smaller than AP
NPs. Ox-AP FeOOH NPs were up to 22.5% smaller than AP NPs. We discuss
and propose a possible mechanism of Ox-AP through careful consideration of
the known reaction chemistry of iron and copper. We propose that an increased
monomer formation rate enhances the nucleation rate, which ultimately results
in smaller particles with a more narrow distribution. The more distinct effect
of Ox-AP on copper was attributed to the fast formation of the stable CuO
monomer compared to AP, where the Cu(OH)2 and/or Cu2 (OH)3 Cl monomers
are more likely. Although the exact mechanism of Ox-AP needs experimental
confirmation, our results nicely demonstrate the potential of using Ox-AP to
produce smaller NPs with a more narrow distribution in comparison to using
AP.
70 OXIDATION-ASSISTED ALKALINE PRECIPITATION: EFFECT OF H2 O2 ON THE SIZE OF CUO AND
FEOOH NANOPARTICLES

1 Introduction

We introduce oxidation-assisted alkaline precipitation (Ox-AP) as an alternative


method to traditional alkaline precipitation (AP). With AP, a metal salt (e.g.,
CuCl2 ) is mixed only with an alkali (e.g., NaOH) to achieve the supersaturation
and precipitation of a sparingly soluble metal (oxy)(hydr)oxide (e.g., CuO).
With Ox-AP the metal salt of the more reduced metal ion (e.g., CuCl) is mixed
with both that alkali and an oxidant (e.g., H2 O2 ). A careful comparison is only
possible if the metal ion has at least two relatively stable oxidation states in
solution. For this reason, we chose aqueous copper- and iron chloride solutions.
The comparison of Ox-AP to AP is relevant given that AP is frequently cited
for the production of CuO and FeOOH nanoparticles [132, 133]. Our original
hypothesis was that, for the same supersaturation values, AP and Ox-AP
would yield nanoparticles with the same characteristics (i.e., composition, size,
morphology, etc.).
According to classical nucleation theory (CNT) [134–139], the particle formation
process via hydrolysis proceeds in four steps:(1) formation of zero-charge
monomers in the form of [M(OH)z (H2 O)N −z ]0 , with metal ion charge z and
coordination N , (2) creation of nuclei via the olation and/or oxolation of zero-
charge precursors, (3) growth of the nuclei via the addition of matter via olation
and/or oxolation and (4) aging of the primary particles via Ostwald ripening
and/or aggregation [19]. Recently, steps (3) and (4) have been scrutinized
for several systems, where meta-stable primary particles (i.e., pre-nucleation
clusters, primary nanoparticles, etc.) can also serve as building blocks for the
growth of nanoparticles, a mechanism not included in CNT. These mechanisms
are coined nonclassical and most of the recent findings are summarized in the
review of De Yoreo and coworkers [140].
CuO and FeOOH particles are easily produced via AP. For example, mixing
a CuCl2 or FeCl3 solution with a highly alkaline solution (e.g., with NaOH)
yields Cu(OH)2 or Fe(OH)3 , respectively, which gradually (either fast or slow)
form CuO or FeOOH via oxolation. The overall reaction scheme simplifies to:

Cu2+ + 2OH− → Cu (OH)2 → CuO + H2 O (5.1)


Fe3+ + 3OH− → Fe (OH)3 → FeOOH + H2 O (5.2)
2+ 3+
where the coordinated water ligands of Cu and Fe are omitted for clarity.
The corresponding reaction quotients Q Cu(OH)2 and Q FeOOH are:

QCu(OH)2 = aCu2+ a2OH− = γCu2+ γOH


2
− [Cu
2+
][OH− ]2 (5.3)
INTRODUCTION 71

QFe(OH)3 = aFe3+ a3OH− = γFe3+ γOH


3
− [Fe
3+
][OH− ]3 (5.4)
with ai the activities and γi the activity coefficients of a given species i, and
the concentrations of Cu2+ , Fe3+ , and OH− [mol L−1 ] given in between square
brackets.
CNT couples the chemical reactions of the monomer formation to the physical
process of phase separation (i.e., from liquid to liquid+solid) via the concept of
supersaturation. For reaction precipitation, the supersaturation S [-] is defined
as the ratio between the reaction quotient Q and solubility product Ksp , as
exemplified for Cu(OH)2 :
QCu(OH)2
SCu(OH)2 = (5.5)
Ksp,Cu(OH)2

with Ksp,Cu(OH)2 :
Ksp,Cu(OH)2 = aCu2+ ,eq a2OH− ,eq (5.6)
and ai,eq the equilibrium activities of species i.
If S > 1, the amount of reactants is in excess for equilibrium and precipitation
can occur. Often S needs to be many times higher than 1 before precipitation
occurs, depending on other limiting factors (e.g., surface energy of the precipitate,
temperature, etc.). If S < 1, the amount of reactants is limiting for precipitation
and hence the reactants stay in solution. According to CNT, a higher
supersaturation level increases the nucleation rate which results in smaller
particles [135, 137–139].
The (homogeneous) nucleation rate J [s−1 ] depends on supersaturation via a
power law, which in the simplest form is written as:
J −2
= 10−A[log S] (5.7)
Jmax
with Jmax [s−1 ] the nucleation rate at infinitely high supersaturation and A [-]
a collection of variables [135, 137–139]:

4βa3 γ 3 V 2
A= (5.8)
[27βv2 (kB T ln 10)3 ]

with βa [-] the form factor of the particle surface , γ [J m−2 ] the surface free
energy per unit area, V [m3 ] the molecular volume, βv [-] the form factor of
the particle volume, kB [m2 kg s−2 K−1 ] the Boltzmann constant and T [K] the
temperature.
A more elaborate approach, including for example heterogeneous nucleation
and other important phenomena during precipitation, is found in scientific
72 OXIDATION-ASSISTED ALKALINE PRECIPITATION: EFFECT OF H2 O2 ON THE SIZE OF CUO AND
FEOOH NANOPARTICLES

CNT literature [135, 137–139, 141]. Although CNT demonstrated its usefulness
in the past, recent studies increasingly demonstrate the limitations of CNT.
These studies discovered different nonclassical crystallization routes and the
importance of chemical kinetics [140, 142].
In this study, we refute our original hypothesis, by demonstrating for two cases
that Ox-AP yielded smaller nanoparticles with a more narrow size distribution
than AP. Ox-AP achieves these feats with as little as possible alteration of the
supersaturation. Contrary to recent discoveries of non-classical crystallization
routes, which are related to steps (3) and (4) of the precipitation process, we
believe that the effect of Ox-AP is related to steps (1) and (2) of the precipitation
process. We discuss the possible nature of Ox-AP of CuO and FeOOH through
careful consideration of the known reaction chemistry of iron and copper.

2 Experimental

2.1 Electrolyte synthesis

Four different metal chloride electrolytes, as well as four different NaOH


electrolytes, were prepared for the study. Prior to the experiments, the initial
metal chloride solution was 250 mL of 10 mM metal chloride with 0.5 M NaCl
and 0.1 M HCl. These solutions were made by diluting 50 mL of a stock
solution with 200 mL demineralized water. The stock solutions were all made
similarly. In a 500 mL volumetric flask, 72.5 g of NaCl was dissolved in 400
mL demineralized water. To this NaCl solution, 25 mL of 10 M HCl (made
from 37% HCl Arcos Organics) was added. A certain amount of metal chloride
powder was added to this solution:

ˆ 4.127 g of CuCl2 .2H2 O (Merck)


ˆ 2.475 g of CuCl (Arcos Organics)
ˆ 6.758 g of FeCl3 .6H2 O (Chem-Lab)
ˆ 4.970 g of FeCl2 .4H2 O (Arcos Organics)

When fully dissolved, the solution was diluted with the required amount of
demineralized water to complete a 500 mL volumetric flask. The NaOH
electrolytes were prepared to match the concentration and the valence of the
metal ions in the metal chloride solutions. This was deemed important because
Jolivet and coworkers demonstrated the importance of pH on the size of metal
oxide nanoparticles [143]. The necessary amount of OH− to precipitate all
EXPERIMENTAL 73

of the metal ions as hydroxides were taken as the starting value (n CM Cln ).
Because the metal chloride solutions had 0.1 M HCl, an additional 0.1 M NaOH
was necessary to overcome the water equilibrium. The end pH of the solution
after mixing was chosen to be pH 12 (i.e., 0.01 M OH− ), therefore an additional
0.02 M NaOH (in the 250 mL initial solution to become 0.01 M in 500 mL final
solution) was necessary. In the case of FeCl2 and CuCl, 10 mM H2 O2 in the
form of 0.283 mL of 30% H2 O2 (Sigma-Aldrich), an excess of 100%, was added
to the 250 mL NaOH solution just before mixing (to avoid H2 O2 decomposition
at high pH). The composition of the NaOH solutions was:

ˆ For mixing with CuCl2 solution: 0.14 M (2 x 0.01 M + 0.1 M + 0.02 M)


NaOH
ˆ For mixing with CuCl solution: 0.13 M (1 x 0.01 M + 0.1 M + 0.02 M)
NaOH + 10 mM H2 O2

ˆ For mixing with FeCl3 solution: 0.15 M (3 x 0.01 M + 0.1 M + 0.02 M)


NaOH
ˆ For mixing with FeCl2 solution: 0.14 M (2 x 0.01 M + 0.1 M + 0.02 M)
NaOH + 10 mM H2 O2

2.2 Preparation of nanoparticles

To have a consistent reaction front and consistent mixing, the solutions were
mixed in a Y-junction by pumping them separately at high equal volumetric
rates, after which the mixture was collected in a glass beaker. The experimental
setup used a Watson-Marlow 323 peristaltic pump, three 90 cm Watson Marlow
1.6/3.2 mm Marprene 902 tubes and a Y-junction Hibiki Y-1 3mm. The Y-
junction was connected to the ends of each tube. The two tubes connected to
the upper legs of the Y-junction were each put into a peristaltic pump head
at around the 75 cm mark. One of the inlets went into the NaOH solution,
the other inlet into the metal chloride solution. The outlet went into a beaker
with a WTW pH electrode Sentix 61 sensor (WTW 340i pH meter) calibrated
with a two-point calibration (pH 4 and pH 7), which was used to measure the
final steady-state pH. Both the metal chloride solution and the NaOH solution
were pumped by an equal rate of 100 mL min-1 through the Y-junction into
the beaker. Depending on the composition of the influents, the effluent was
colorless with differently colored precipitates.
A summary of the electrolytes and a schematic drawing of the Y-junction
tubular mixing-reactor are shown in figure 5.1.
74 OXIDATION-ASSISTED ALKALINE PRECIPITATION: EFFECT OF H2 O2 ON THE SIZE OF CUO AND
FEOOH NANOPARTICLES

Solution MCln [M] HCl [M] NaCl [M]


a. CuCl2 0.01 0.1 0.5
b. CuCl 0.01 0.1 0.5
c. FeCl3 0.01 0.1 0.5
d. FeCl2 0.01 0.1 0.5

Solution Precipitate Color pH


100 mL min-1 200 mL min-1 a. CuO/Cu2(OH)3Cl turquoise 11.5
b. CuO brown 11.5

100 mL min-1 c. FeOOH orange 11.7


d. FeOOH orange 11.7

Solution NaOH [M] H2O2 [M] NaCl [M]


a. 0.14 0.00 0.5
b. 0.13 0.01 0.5
c. 0.15 0.00 0.5
d. 0.14 0.01 0.5

Figure 5.1: Schematic of Y-junction tubular mixing-reactor with reagent


characteristics and flow rates. Each metal salt solution is mixed with a suitable
alkaline solution to reach a theoretical pH of 12. Corresponding influents are
numbered equally with a, b, c, and d respectively.

2.3 Post-treatment and characterization

For each of the samples, the post-treatment was equal. 250 mL of the effluent
was kept for storage and possible later use. The other 250 mL was centrifuged
with a Jouan CR422 in 6 centrifuge tubes of 50 mL capacity for 10 min at 3000
RPM. The supernatant of the tubes was removed and the remaining precipitate
was washed with demineralized water and using a vortex (IKA VF2). The
resulting solution was diluted to 50 mL into a 50 mL tube. This tube was
centrifuged for 10 min at 14000 RPM (Sorvall LYNX 6000 Superspeed Centrifuge
- rotor for 12 50 mL tubes), the supernatant removed and subsequently washed
in 50 mL demineralized water. These steps were repeated once more, but adding
only 20 mL demineralized water instead of 50 mL. From this 20 mL, 1 mL was
used for dilutions for the single-particle ICP-MS measurements. 9 mL was used
for dilutions of the DLS measurements. The other 10 mL was dried for dry
analysis.

DLS samples and characterization. From the 9 mL solution, 0.1 mL was


diluted to 1 mL with demineralized water and put into a cuvette (67.754, Sarsted,
polystyrene, 4 optical sides) for DLS characterization (Malvern - Zetasizer
NANO ZSP). Every sample was measured in triplets. Every measurement gave
a log-normal distribution of the precipitate sizes. The average and standard
deviation of the average z-diameter were taken as the values for analysis.
RESULTS 75

Table 5.1: pH resulting after mixing the different solutions (the theoretical pH
value anticipated for all instances is pH 12)
Solution pH measured after mixing
CuCl2 11.67 ± 0.02
CuCl + H2 O2 11.72 ± 0.04
FeCl3 11.49 ± 0.05
FeCl2 + H2 O2 11.54 ± 0.06

sp ICP-MS samples and characterization. These samples required a more


sample-specific approach. The details are given in appendix C.

Dry analysis samples and characterization. 10 mL of the washed solution


was dried in a Heidolph Rotavapor Hei-VAP Precision at 70 ◦ C and a vacuum
of 70 mbar to obtain dried powdered samples for scanning electron microscopy
(SEM) and X-ray powder diffraction (XRD) analysis. SEM was used for the
determination of the size and morphology of the nanoparticles (FEI Nova
Nanosem 450 with CBS detector). XRD was used for the determination of the
composition of the nanoparticles (Bruker D2 phaser, Cu source with Kα,avg =
1.54, line focus with a LynxEye detector).

3 Results

3.1 Visual observations of nanoparticle suspension

Figure 5.2 demonstrates the dispersed precipitates just after mixing, stored in
plastic storage tubes. When H2 O2 was used, some gas evolution was observed,
in the form of bubbles, suggesting the decomposition of excess H2 O2 to oxygen
gas and water (see insets of figures 5.2b and 5.2d). Table 5.1 shows the pH
measured after mixing, which was lower than the expected value of 12. This is
most likely due to the sodium error at 0.5 M NaCl. The pH difference between
the copper case and the iron case was approximately 1 mM OH− , and thus
considered insignificant for this study, because it is less than 1 % of the amount
of the total OH− added. The copper samples had a distinct color difference
(figure 5.2). Initially, the Cu2+ sample was distinctly blue but developed into
a dark brown color by the end of the washing process. The Cu+ (with H2 O2 )
sample was dark greenish-brown for a very short instance, but even before the
washing step started its color changed to dark brown. Both the Fe3+ and the
76 OXIDATION-ASSISTED ALKALINE PRECIPITATION: EFFECT OF H2 O2 ON THE SIZE OF CUO AND
FEOOH NANOPARTICLES

Figure 5.2: Color of the precipitates obtained directly after precipitation. From
left to right: CuCl2 (blue/turquoise), CuCl with H2 O2 (dark green/brown),
FeCl3 (bright orange) and FeCl2 with H2 O2 (bright orange). Insets for better
appreciation of the gas bubbles formed in the cases with H2 O2 .

Fe2+ (with H2 O2 ) samples were bright orange, with little observable change
over time.

3.2 Size of the nanoparticles

Single-particle ICP-MS and DLS were used to determine the size of the
nanoparticles [144]. In all cases, spherical shape was assumed, and thus, the
particle spherical equivalent diameter is reported. For rod-shaped particles
of similar length over width ratio, the comparison of particles with different
sizes remains valid [145]. Figure 5.3 demonstrates the particle size distributions
calculated from the sp ICP-MS measurements for the different cases. In the case
of copper, the size of the particles shows a remarkable difference, contradicting
our initial hypothesis of size consistency between AP and Ox-AP for the same
supersaturation value. In the case of iron, the size difference follows a similar
trend, although much less pronounced. Figures 5.4a and 5.4b respectively
compare the DLS and sp ICP-MS size measurements. Both measurements
demonstrate that the Ox-AP yields smaller particles than the traditional AP.
In the case of copper, the average size of the oxidation-induced precipitates is
3 times smaller (i.e., 30 nm instead of 90 nm) according to sp ICP-MS, and
over 2 times smaller (i.e., 100 nm instead of 250 nm) according to DLS. In the
case of iron, the results are less pronounced, yet inconsistent with our original
hypothesis: the average size of the oxidation-induced precipitates is 12.5%
RESULTS 77

20 20
CuCl2 FeCl3
CuCl + H2O2 FeCl2 + H2O2
relative frequency [%]

relative frequency [%]


15 15

10 10

5 5

0 0
10 30 50 70 90 110 130 150 10 20 30 40 50 60 70 80
equivalent spherical diameter [nm] equivalent spherical diameter [nm]

Figure 5.3: Particle size measurement distributions from single-particle ICP-MS


characterization for the cases of copper (left) and iron (right), respectively.

Table 5.2: Polydispersity index values from single-particle ICP-MS and DLS
measurements of the FeOOH and CuO precipitates formed by Ox-AP and AP
CuCl2 CuCl + H2 O2 FeCl3 FeCl2 + H2 O2
sp ICP-MS 0.32 0.08 0.12 0.07
DLS 0.38 0.28 0.23 0.20

smaller (i.e., 28 nm instead of 32 nm) than with the traditional AP-obtained


precipitates according to sp ICP-MS, and 22.5% smaller (i.e., 117 nm instead
of 151 nm) according to DLS. The difference in the sizes obtained by both
techniques, sp ICP-MS and DLS, can mainly be attributed to two factors. Firstly,
there exists a chance that agglomerates remain in our washed, but unstabilized,
sample, resulting in increased apparent particle diameters [146, 147]. Secondly,
because of anisotropy, the particle size (z-average) of DLS measurements can be
strongly distorted [147]. Table 5.2 shows the polydispersity index (PDI) for both
single-particle ICP-MS and DLS, respectively. A lower PDI value corresponds
to a more narrow distribution. Both techniques demonstrate a more narrow
distribution for Ox-AP than for traditional AP and this for both the copper
case and the iron case.
78 OXIDATION-ASSISTED ALKALINE PRECIPITATION: EFFECT OF H2 O2 ON THE SIZE OF CUO AND
FEOOH NANOPARTICLES

250 100
a 95 b
225
90

sp ICP-MS average ESD [nm]


200
85
175
DLS z-average [nm]

80
150 40
125 35
30
100 25
75 20
50 15
10
25 5
0 0
CuCl2 FeCl3 CuCl2 FeCl3
CuCl + H2O2 FeCl2 + H2O2 CuCl + H2O2 FeCl2 + H2O2

Figure 5.4: Average diameters of the formed nanoparticles for the cases of
copper and iron derived from distributions of (a) z-average diameter from DLS
measurements and (b) average equivalent spherical diameter from single-particle
ICP-MS measurements. Error bars indicate the standard deviation of the 3
measured average particle sizes (i.e., not from the particle distributions).

3.3 Composition of the nanoparticles

The XRD-patterns for the copper and iron cases are shown in figures 5.5a
and 5.5b, respectively. CuO is the main phase of the copper case. The CuCl2
case shows two additional peaks, indicating an additional Cu2 (OH)3 Cl phase.
This indicates that at least part of the precipitated phase is the Cu2 (OH)3 Cl
instead of the Cu(OH)2 and besides the reactions described in equation (5.1),
the following overall reactions are likely:

2Cu2+ + 3OH− + Cl− → Cu2 (OH)3 Cl (5.9)

Cu2 (OH)3 Cl + OH− → 2CuO + 2H2 O + Cl− (5.10)


The mechanism in equation (5.10) would be similar to the conversion mechanism
of Cu(OH)2 to CuO at a higher pH [148, 149]. Both iron cases show highly
amorphous XRD patterns of 2-line ferrihydrite (hydrous FeOOH), although only
the main (110) peak (one of two peaks in 2-line ferrihydrite) is visible [150, 151].
RESULTS 79

■ Cu2(OH)3Cl a • FeOOH b
◻ CuO ◻ •
◻ FeCl2 + H2O2
CuCl + H2O2


◻ ◻ ◻ ◻ ◻◻

a.u. [-]
a.u. [-]

• FeCl3

◻ CuCl2
■ ◻ ■ ◻ ◻
◻ ◻ ◻◻

15 20 25 30 35 40 45 50 55 60 65 70 15 20 25 30 35 40 45 50 55 60 65 70
2θ [�] 2θ [�]

Figure 5.5: XRD-patterns for the precipitates obtained by traditional alkaline


precipitation (bottom) and oxidation-induced alkaline precipitation (top) for
the cases of copper (a) and iron (b). In (b) the grey line is the original
background-subtracted measurement, the black line is the moving average,
for better appreciation. The symbols ,  and • are corresponding to the
diffraction peaks of Cu2 (OH)3 Cl, CuO and FeOOH, respectively.

It is known that, with time ferrihydrite can convert into more crystalline
products such as goethite and hematite [152].

3.4 Morphology of the nanoparticles

Figure 5.6 shows the SEM images for the different precipitates. For the copper
case, separate particles are easily distinguished for AP and Ox-AP. Ellipsoidal
needle-shaped NPs are observable in either case, but the Ox-AP NPs are
smaller than the AP NPs. For the AP case, the existence of a second, less
abundant, phase (i.e., Cu2 (OH)3 Cl) is also clear. For the iron case, even at
a higher magnification than for the copper case, separate particles are hardly
distinguishable for AP and Ox-AP (figures 5.6c and 5.6d). The nanoparticles
form large aggregates of smaller nanoparticles. With digital magnification, it was
possible to identify a few separate nanoparticles (see insets figure5.6). However,
these do not provide conclusive evidence on the size differences between AP
and Ox-AP for the iron case.
80 OXIDATION-ASSISTED ALKALINE PRECIPITATION: EFFECT OF H2 O2 ON THE SIZE OF CUO AND
FEOOH NANOPARTICLES

Figure 5.6: SEM images of the dried precipitates. CuCl2 without H2 O2


(a) and CuCl with H2 O2 (b) and FeCl3 without H2 O2 (c) and FeCl2 with
H2 O2 (d). Insets are digital magnifications to demonstrate very small separate
nanoparticles.
DISCUSSION 81

4 Discussion

The observations can be summarized as follows: (1) the experiments with iron
result in smaller nanoparticles with more narrow size distributions than the
experiments with copper, both for Ox-AP and AP; (2) Ox-AP results in smaller
nanoparticles with more narrow size distributions than AP, both for the copper
and the iron cases; (3) the effect of Ox-AP is much more pronounced in the
case of copper than in the case of iron.
The observations can be discussed in the light of the particle formation process.
As discussed above, the particle formation process via hydrolysis proceeds in
four steps:(1) formation of zero-charge precursors, (2) creation of nuclei via the
olation and/or oxolation of zero-charge precursors, (3) growth of the nuclei via
addition of matter via olation and/or oxolation and (4) aging of the primary
particles via Ostwald Ripening and/or aggregation [19]. Step (1) depends on the
solubility of the species and the concentration of the reagents and the available
ligands (i.e., especially on the pH). Moreover, it is also limited by the lability
of the water ligand in the first coordination shell of the metal ion. Step (2)
depends on the concentration of the zero-charge species and the water lability
as well, because the lability co-dictates the rates at which olation and oxolation
can occur. Step (3) and step (4) are still intensively debated in current literature
and strongly depend on step (1) and step (2), and any of the other process
parameters [140].
In the AP process, OH− ions compete with H2 O and Cl− ions to enter the first
hydration sphere of the metal aquo complex. The strong nucleophilic character
of OH− allows it to form more stable complexes with the metal ion, and at the
right OH− concentration, the most stable form is a sparingly soluble zero-charge
monomer. This monomer can form polymers via olation and oxolation until a
critical size is reached and they are stable in solution. The critical size strongly
depends on the inherent chemical nature of the monomer (i.e., the metal ion,
the ligands, etc.), the olation and oxolation kinetics to which it is subjected,
and also on the level of supersaturation (i.e., the concentration of the precursor
ions).
Supersaturation S is much higher for the iron case than for the copper case
(table 5.3), more detailed calculations are given in appendix D, and in accordance
with CNT, the NPs in the iron case should be smaller than in the copper case.
Since the final pH, supporting electrolyte concentration, metal ion concentration
and process operating conditions were the same for the copper and iron cases,
the inherent differences between aqueous behavior of Cu2+ and Fe3+ (i.e., the
case-specific values of Jmax and A from equation (5.7)) likely cause the observed
size difference between the CuO and hydrous FeOOH nanoparticles. The values
82 OXIDATION-ASSISTED ALKALINE PRECIPITATION: EFFECT OF H2 O2 ON THE SIZE OF CUO AND
FEOOH NANOPARTICLES

Table 5.3: Solubility products and supersaturation values for the hydroxides of
copper and iron acting as a precursor for FeOOH and CuO
Precursor Ksp S for end-pH 12
Cu(OH)2 2.20 10−20 [mol3 L−3 ]( [107]) 9.1 1013
−35 6 −6 ( [153])
Cu2 (OH)3 Cl 2.34 10 [mol L ] 5.2 1024
−39 4 −4 ( [107])
Fe(OH)3 2.79 10 [mol L ] 2.8 1031

of Jmax and A depend on the chemical nature of the monomers and the olation
and oxolation kinetics of those monomers. This is inherently different for copper
and iron. Discussion of these differences would lead us too far afield.
More interesting is to compare the Ox-AP cases of copper and iron with the AP
cases of copper and iron, respectively. Our original hypothesis was that, for the
same supersaturation values, AP and Ox-AP would yield nanoparticles with
the same characteristics (i.e., composition, size, morphology, etc.). Imagine
infinitely fast oxidation of the metal ions in the Ox-AP case. This would cause
complete oxidation prior to any other reaction (i.e., any of the four steps in the
particle formation process), OH− would be formed according to the following
equations:
2Cu+ + H2 O2 + 2OH− → 2Cu2+ + 2OH−
Cu2+
+ 2OH− → 2CuO + 2H2 O (5.11)

2Fe2+ +H2 O2 +4OH− → 2Fe3+ +2OH−


Fe3+
+4OH− → 2FeOOH+2H2 O (5.12)
where the Cu2+ and Fe3+ subscripts stress the fact that these OH− ions are
formed by the redox reaction of the metal ion with H2 O2 . The metal ion
concentration and the amount of OH− would be the same for AP and Ox-AP.
Therefore, the overall supersaturation would be the same and CNT would thus
predict the same precipitates with the same sizes. Interestingly, our experimental
results contradict this: compared to AP, Ox-AP results in smaller nanoparticles
with a more narrow size distribution. This refutes the hypothesis. Therefore
H2 O2 likely interferes with the mechanisms of steps (1) to (4), rather than
cause complete oxidation prior to those steps. The question remains, where
does H2 O2 interfere in the particle formation process?
Because Ox-AP results in smaller particles with more narrow size-distributions,
it is unlikely that in our process H2 O2 would interfere significantly in the aging
process (i.e., step (4)), because although aging tends to narrow the distribution
size, it generally increases the average particle size. This is not in agreement
with our observations. Therefore, prior to aging, the particles should already
have smaller sizes and more narrow distributions.
The later stages of growth and aging are often suppressed using large molecules
that block the surface to avoid coalescence or by using certain ligands that
DISCUSSION 83

suppress the redissolution of smaller particles in favor of the larger particles (i.e.,
Ostwald ripening). It is also unlikely that hydrogen peroxide would interfere in
these stages, especially because most of the hydrogen peroxide reacts away in
an oxidation reaction with the metal, and the excess quickly decomposes (i.e.,
bubbles are visible in the final solution). Therefore the interference of H2 O2
is most likely occurring during step (1) and/or step (2) of the precipitation
process. The olation and oxolation reactions in step (2) depend strongly on the
characteristics of the zero-charge monomers formed in step 1. Therefore, we
argue that H2 O2 takes an important role in the formation of the monomers and
the subsequent olation and/or oxolation reactions.
In Ox-AP, both OH− and HO− 2 (i.e., the deprotonated form of H2 O2 at high pH
values [20]) act as strong nucleophiles (i.e., HO− −
2 even more so than OH [154])

and compete with water and Cl to enter the first hydration sphere of the metal
ion. There are many possible reaction pathways for HO− 2 (i.e., H2 O2 ) with the
metal ion, known in scientific literature as Fenton(-like) reactions [155,156]. The
exact pathways are still subject to much research and heavily system dependent.
Here, we propose, both for copper and iron, only one of the possible pathways.
For clarity, we do not take into account the many possible side-reactions and
intermediaries.
Without hydrogen peroxide, the expected zero-charge monomer of Cu+ is CuOH
at elevated pH, with a subsequent oxolation reaction to Cu2 O:

2Cu+ + 2OH− → 2CuOH → Cu2 O + H2 O (5.13)

With hydrogen peroxide, the CuOOH is another possible monomer:

Cu+ + HO−
2 → CuOOH (5.14)

Subsequently, the oxidative nature of the HO−


2 ligand could result in a Fenton
reaction [156]:
CuOOH → CuO + OH· (5.15)
The proposed reactions require little ligand exchanges and result directly in
the CuO monomer, rather than the Cu(OH)2 monomer or the Cu2 (OH)3 Cl
dimer. The OH· radical is a non-selective oxidant which can oxidize neighboring
species, for example, the Cu+ ion or the CuOH monomer, at high rates [157]:

Cu+ + OH· → CuO + H+ (5.16)

CuOH + OH· → CuO + H2 O (5.17)


In any of the proposed reactions, the CuO is the end product. This might explain
why the conversion to the brown CuO precipitate is much faster for Ox-AP than
for AP. The rapid formation of the stable CuO monomer is a likely explanation
84 OXIDATION-ASSISTED ALKALINE PRECIPITATION: EFFECT OF H2 O2 ON THE SIZE OF CUO AND
FEOOH NANOPARTICLES

of the smaller nanoparticles. The rapid depletion of monomers during the


nucleation phase would then explain the more narrow size distribution. The
lack of available monomers also slows down the subsequent growth. Although
no evidence is currently available, we hope that the exact mechanism can be
determined experimentally in the near future.
In the case of iron, the effect of Ox-AP is less obvious. A possible reason
follows from our proposed reaction mechanism. Without hydrogen peroxide,
the expected zero-charge monomer is Fe(H2 O)3 (OH)3 . The complexed water
results from the fact that Fe2+ and Fe3+ easily form hexaaqua complexes [158].
The Fe(H2 O)3 (OH)3 monomer, subsequently transforms into FeOOH(H2 O)x
and with time (i.e., with aging) crystallizes into FeOOH, with different possible
crystal structures depending on the solution and process characteristics. Aging
was not allowed in our experiments, which limited crystallization. With hydrogen
peroxide, we can draft a reaction similar to equation (5.14), taking into account
that Fe2+ easily forms the hexaaqua complex:
[Fe(H2 O)6 ]2+ + HO− +
2 → [Fe(H2 O)5 OOH] + H2 O (5.18)
[Fe(H2 O)5 OOH]+ → [Fe(H2 O)4 (OH)2 ]+ + OH· + H2 O (5.19)
Unlike with copper, with iron, the reaction does not directly result in a zero-
charge monomer. Moreover, the product itself (i.e., [Fe(H2 O)4 (OH)2 ]+ ) is an
intermediary for ferrihydrite precipitation in AP [159, 160]. The ferrihydrite
results from a second step with OH− :
[Fe(H2 O)4 (OH)2 ]+ + OH− → Fe(OH)3 (H2 O)3 + H2 O (5.20)
Remark that the sequence of these reactions is not fixed. The hydrolysis reaction
with OH− might also occur first in the sequence, after which HO− 2 reacts with
the [Fe(H2 O)5 OH]+ complex to form the zero-charge [Fe(H2 O)4 (OH)OOH]
monomer. This monomer would then further oxidize to form the Fe(OH)3 (H2 O)3
monomer.
In the above reaction mechanisms, the reason might be found for the lesser effect
of Ox-AP in the case of iron. The reaction with HO− 2 does not immediately
give rise to a zero-charge monomer, requiring an extra step. Additionally, the
resulting monomer is the same as the one expected for AP of iron, rather than
directly resulting in a more stable form, as with copper. This is in line with
the visibly observed alikeness of the precipitates (i.e., no transformation is
obvious like with copper). One possible reason that an effect is still observable,
could be that with Fe2+ , the zero-charge monomer forms with only two ligand
substitutions rather than three in the AP case of Fe3+ . Additionally, the water
exchange kinetics of Fe2+ are much faster than those of Fe3+ (4 orders of
magnitude [161]), which might cause faster nucleation rates and the related
observed, but small, effects.
CONCLUSION 85

5 Conclusion

We demonstrated that Ox-AP yields smaller nanoparticles with a more narrow


size distribution than traditional AP for iron- and copper-chloride solutions in
conditions of high supersaturation. For CuO, the differences between traditional
AP and Ox-AP are very clear. For FeOOH, the differences are observable
to a lesser extent. The proposed mechanisms of Ox-AP are the improved
condensation kinetics and the faster formation of a stable zero-charge CuO
monomer. For iron, we believe that the condensation kinetics improve less than
for copper. We propose that the lack of a more stable monomer, in contrast
to the copper case, causes this lesser effect. Although Ox-AP of iron lacks the
more stable monomer, two other factors were proposed as possible causes for
the observed effects. The first factor is the lower amount of necessary ligand
exchanges to form a zero-charge monomer for Fe2+ compared to Fe3+ . The
second factor is the higher water exchange kinetics for Fe2+ than for Fe3+ . Both
factors possibly result in higher condensation kinetics for Ox-AP. In light of
our observations, it is possible to prospect other systems for the production
of small nanoparticles with a narrow size distribution. Herein, a sufficiently
stable precursor metal ion that can be oxidized is detrimental. Our theory is
insightful for further investigations of other systems, but we hope that the exact
mechanism can be confirmed experimentally in the near future to allow for
better synthesis-by-design.
Chapter 6

Oxidation-assisted alkaline
precipitation of nanoparticles
using gas-diffusion electrodes

In Chapter 4, the important parameters that influenced the shape of


precipitation titration curves were discussed. In Chapter 5, the role of
H2 O2 during metal (oxy)(hydr)oxide precipitation and the mechanism of
oxidation-assisted alkaline precipitation (Ox-AP) were discussed. This
chapter compares the GDEx process to alkaline titration (AT) for several
metal-containing solutions (i.e., ZnCl2 , MnCl2 , or FeCl2 ). In this way,
GDEx is demonstrated to elicit alkaline precipitation (AP) in the cases
of Zn and Mn, but to elicit Ox-AP in the case of Fe. This GDEx-induced
oxidation-assisted alkaline precipitation results in smaller nanoparticles
than those nanoparticles obtained by AT-induced alkaline precipitation.
The cause of the smaller sizes obtained is the topic of investigation in
Chapter 5.

Contributions of the author were hypothesis development, carrying out all


experiments, characterizations and data treatment. R. Prato helped with
hypothesis development, discussions, and proof-reading. X. Dominguez Benetton
and J. Fransaer helped with funding, discussions, and proof-reading.

87
88 OXIDATION-ASSISTED ALKALINE PRECIPITATION OF NANOPARTICLES USING GAS-DIFFUSION
ELECTRODES

This chapter is prepared for submission to Reaction Chemistry & Engineering


as: Eggermont S.G., Prato R., Dominguez Benetton X., Fransaer J. Oxidation-
assisted alkaline precipitation of nanoparticles using gas-diffusion electrodes.
OXIDATION-ASSISTED ALKALINE PRECIPITATION OF NANOPARTICLES USING GAS-DIFFUSION
ELECTRODES 89

Abstract

Metal oxide nanoparticles become increasingly important as functional


materials because their diversity in composition and structure allow the
control of their physical properties. This work investigates the synthesis of
metal (oxy)(hydr)oxide nanoparticles (NPs) via oxidation-assisted alkaline
precipitation (Ox-AP) using gas-diffusion electrodes, namely gas-diffusion
electrocrystallization (GDEx). GDEx was benchmarked against alkaline
titration (AT). NPs were synthesized from ZnCl2 , MnCl2 , or FeCl2 precursor
solutions at room temperature. Using AT, Zn(OH)2 , Mn3 O4 , and FeO NPs
were synthesized, respectively. Using GDEx, Zn(OH)2 , Mn3 O4 , and Fe3 O4 NPs
were synthesized, respectively. The AT and GDEx process of the ZnCl2 and
MnCl2 solutions demonstrated very similar pH behavior during precipitation
and the Zn(OH)2 and Mn3 O4 NPs synthesized with either technique were
similar in size, morphology, and composition. For these cases, AT and GDEx
both elicited alkaline precipitation and were considered equivalent processes
for NP synthesis. In contrast, the AT and GDEx process of the FeCl2 solution
demonstrated very different pH behavior during precipitation. Moreover, the
FeO NPs, synthesized with AT, were much larger and of different shape and
composition than the Fe3 O4 NPs, synthesized with GDEx. The smaller sizes
obtained with GDEx were attributed to Ox-AP resulting from the oxidation of
Fe(II) to Fe(III) by H2 O2 or HO−2 during precipitation, improving condensation
kinetics and increasing the supersaturation, both well-known size-determining
factors.
90 OXIDATION-ASSISTED ALKALINE PRECIPITATION OF NANOPARTICLES USING GAS-DIFFUSION
ELECTRODES

1 Introduction

Society is relying increasingly on the development of new functional materials


[90]. Nanoparticles are a branch of functional materials, often praised for their
size-dependent properties, such as electrical, magnetic, mechanical, optical,
and chemical properties, etc. and their wide range of applications in various
fields, such as catalysis, energy storage, medicine, sensing, among others [162].
More specifically, metal (oxy)(hydr)oxides are of interest because of their
diversity and the fact that they can adopt many structural geometries that can
exhibit metallic, semiconductor, or insulator properties [163]. In light of the
increasing interest in these materials, we recently demonstrated the role of the
oxidation-assisted alkaline precipitation (Ox-AP) mechanism in synthesizing
metal (oxy)(hydr)oxide nanoparticles with smaller sizes and more narrow size
distributions than achievable with the traditional alkaline precipitation (AP)
mechanism [164]. While the AP mechanism constitutes the reaction of an alkali
(e.g., OH− ) with a metal ion to form a sparingly soluble metal (oxy)(hydr)oxide
(nano)particle [132, 133, 165–167], Ox-AP constitutes the simultaneous reaction
of an alkali and an oxidant (e.g., H2 O2 ) with a metal ion in solution to form
sparingly soluble metal (oxy)(hydr)oxide (nano)particles of a higher oxidation
state [164].
In general, chemical precipitation consist of three stages: stage 1 before
precipitation, stage 2 during precipitation and stage 3 after precipitation [168].
This is also the case for AP, and it is easily visualized by plotting the pH
as a function of the added amount of alkali (figure 6.1). At the start of any
process driven by the AP mechanism, the pH is sufficiently low to allow all
metal ions to remain dissolved in solution. In stage 1, the H+ ions in solution
react with the added OH− forming water; the pH increases in proportion with
the OH− addition. After the addition of a certain amount of alkali, sufficient
H+ ions are removed and the precipitation pH of a sparingly soluble metal
(oxy)(hydr)oxide is reached. This is the pH value at which the solubility limit
of the metal (oxy)(hydr)oxide is surpassed for a given metal ion concentration.
At this point, stage 1 ends and stage 2 starts. In stage 2, further addition
of alkali hardly increases the pH because the added alkali participates in a
chemical reaction with the metal ion to form the precipitating species. The
chemical process is initiated by the hydroxylation of the metal-aquocomplex,
propagated by subsequent olation or oxolation reactions and completed when a
neutral species is formed, which can nucleate and grow to become a solid metal
(oxy)(hydr)oxide (nano)particle [19, 143]. During this process, a practically
constant pH is observed as a plateau. This plateau is hereinafter referred to as
the precipitation plateau. Stage 2 ends and stage 3 starts when practically all
metal ions have reacted to form precipitates, and the pH increases again upon
INTRODUCTION 91

further addition of alkali. The theoretical amount of OH− ions necessary to


reach the beginning and end of stage 2 can be calculated for an expected initial
pH, initial metal ion concentration, and expected precipitate stoichiometry [168].
Stage 1 is expected to finish after all H+ ions are compensated for. In a 0.1 L
solution at pH 3 (i.e., 10−3 M H+ ions), a total amount of 10−4 mol H+ ions need
to be compensated for. Therefore, prior to precipitation, stage 1 is expected to
end when 10−4 mol OH− ions is added. For metal(II) hydroxide precipitation,
an additional amount of OH− ions needs to be added, equal to two times (i.e.,
the valence of the metal ion) the initial metal ion concentration to reach the end
of stage 2. For a 0.1 L solution with an initial pH of 3 and an initial metal ion
concentration of 5 10−3 M, the end of stage 2 is thus reached after the addition
of 1.1 10−3 mol OH− ions (i.e., 10−4 mol to compensate for the H+ ions and
10−3 mol OH− to precipitate all the metal ions) [168]. Deviations from these
expectations are helpful to identify and determine characteristic phenomena or
singularities.

1 2 3

precipitation plateau
pH

precipitation pH

added NaOH or charge [mol]

Figure 6.1: Arbitrary example of the pH as a function of the added alkali,


indicating the three consecutive stages.

Oxidation-assisted alkaline precipitation (Ox-AP) is similar to AP, except for


the fact that in stage 2, an alkali and an oxidant would react simultaneously
with the metal ion, inducing different olation or oxolation mechanisms and
therefore resulting in other sparingly soluble metal (oxy)(hydr)oxides containing
the metal in a higher oxidation state. It is believed that this change in olation
92 OXIDATION-ASSISTED ALKALINE PRECIPITATION OF NANOPARTICLES USING GAS-DIFFUSION
ELECTRODES

or oxolation mechanisms makes it possible to form smaller nanoparticles with


Ox-AP than with AP under similar average supersaturation conditions [164].
In connection to this, we recently introduced gas-diffusion electrocrystallization
(GDEx) as a new and highly controllable electrochemical synthesis method for
iron oxide nanoparticles [169]. GDEx has the potential to run on renewable
energy and become a viable competitor to other synthesis methods that require
large amounts of chemicals, produced in less-renewable ways (e.g., NaOH).
Otherwise, GDEx would be competitive vs. (hydro)(solvo)thermal synthesis
methods, as it does not operate at high temperatures.
During GDEx with an oxygen- or air-fed gas-diffusion electrode (GDE), an
alkali (i.e., OH− ) and an oxidant (i.e., HO−
2 ) are produced electrochemically
via the oxygen reduction reactions (ORRs):
In the bulk electrolyte, HO−2 will protonate to H2 O2 if the pH is sufficiently low
(pKH2 O2 = 11.62 [20]). The simultaneous production of an alkali and an oxidant
inspired us to investigate the potential of GDEx to elicit Ox-AP, which would
give it the potential to produce small metal (oxy)(hydr)oxide nanoparticles
more sustainably than with traditional methods.
The goal of this work was to benchmark this promising GDEx method against
the alkaline titration (AT) method, a proven AP method for the production of
metal (oxy)(hydr)oxide nanoparticles [132, 133, 165–167]. Our initial hypothesis
was that: (1) if a metal ion can be oxidized by HO− 2 or H2 O2 , GDEx might
elicit Ox-AP of that metal ion and yield nanoparticles with smaller sizes than
those obtained with AT (eliciting AP) under similar process conditions (i.e.,
the same alkali addition rate), and (2) if a metal ion cannot be oxidized, GDEx
would elicit AP instead of Ox-AP, making it effectively equivalent to AT. Our
comparison of AT and GDEx was done for three metal chloride solutions: ZnCl2
(i.e., Zn2+ ), MnCl2 (i.e., Mn2+ ), and FeCl2 (i.e, Fe2+ ). Recently, AT of ZnCl2
was investigated in-depth and was demonstrated as a good reference case to
compare other cases against [168]. In light of this, and because of the inherent
inertness of Zn2+ towards oxidation, it was assumed to be a good reference
case for the comparison of AT and GDEx. Fe2+ was chosen because of previous
success towards the formation of small nanoparticles under conditions that
induce Ox-AP (in a y-junction reactor [164] and with GDEx [169]). Mn2+ was
chosen as a case in between these two cases, since it can be oxidized but only
at higher potentials than Fe2+ [20]. The ability of GDEx to produce only the
necessary reagents, without excess, on-site, and the fact that these reagents
are essentially oxygen (from the air), water and the electron, make GDEx an
inherently green process. Additionally, its use of the oxygen reduction reaction
at the cathode for the production of OH− , instead of the hydrogen evolution
reaction, drastically reduces its power consumption. This reduction is so potent,
EXPERIMENTAL 93

that the use of oxygen-fed gas-diffusion electrodes as cathodes is gaining traction


in the chlor-alkali industry [110, 170–172]. Therefore, the use of GDEx as a
synthesis method makes sense, not only from a synthesis perspective but also
from a sustainability perspective.

2 Experimental

2.1 Electrolyte solutions

A NaCl stock solution of 5.00 M NaCl was made with 292.2 g of anhydrous NaCl
(Merck) in a volumetric flask of 1 L and filled with the appropriate amount
of demineralized water. An acid stock solution of 0.01 M HCl was made by a
two-step 1/10 dilution of a 1 M HCl solution (Chem-Lab) with demineralized
water. Metal chloride stock solutions of 0.500 L with a concentration of 0.200 M
of the individual metal chloride were made, with ZnCl2 (anhydrous, Chem-lab),
MnCl2 ·4H2 O (Sigma Aldrich) or FeCl2 ·4H2 O (Arcos organics). First, 110% of
the necessary mass was weighed and added to the required amount of water to
complete a 0.500 L volumetric flask. The FeCl2 solution was filtered to remove
any oxidized iron-containing particles. Subsequently, ICP-OES (Varian 720-ES)
was used to determine the concentration of Zn, Mn, and Fe. The values of these
concentrations were used to calculate the appropriate dilution factors needed to
establish 1 L metal chloride stock solutions with a final composition of 0.200 M
ZnCl2 , MnCl2 or FeCl2 , respectively. The GDEx catholyte solutions and AT
solutions were made by diluting 0.025 L of the metal chloride stock solution
with 0.100 L of the acid stock solution, 0.100 L of NaCl stock solution, and
0.775 L of demineralized water. Each experiment was carried out with 0.100
L of these solutions, with a composition of 1 mM HCl with 0.50 M NaCl and
with 0.005 M ZnCl2 , MnCl2 or FeCl2 . The GDEx anolyte solution was a 0.100
L 0.50 M NaCl solution.

2.2 GDEx flow-cell and GDEx process

The complete GDEx set-up is shown in figure 3.2a and a schematic of the design
of a representative GDEx flow-cell reactor in figure 3.2b. The volume of the
flow-cell reactor compartments was 10 mL.
®
The cathode was a VITO CoRE gas-diffusion electrode (GDE) [58]. The
anode was a flat platinized tantalum plate. The geometric surface areas of
the anode and cathode were 10 cm2 , respectively. In between the anolyte and
catholyte compartment, there was an anion exchange membrane (FUMASEP ®
94 OXIDATION-ASSISTED ALKALINE PRECIPITATION OF NANOPARTICLES USING GAS-DIFFUSION
ELECTRODES

FAP-4130-PK). This membrane, combined with a sufficiently high chloride


concentration, was chosen to avoid as much as possible the loss of OH− ions

®
through the membrane. The anolyte and catholyte reservoirs were 250 mL
DURAN laboratory bottles, with additional flask openings to allow for a
pH-meter. The cap was supplemented with openings for argon purging. The
solutions in the reservoir bottles were stirred at approximately 700-800 rpm
with a magnetic stirring bar of 3 cm, on an IKA RCTbasic S1 stirrer at rotation
level 7. The electrolytes were circulated with a double-headed peristaltic pump
(Cole-Palmer MasterflexTM L/s 7551-00 with Easyload II 77200-60 pump heads)
at 200 mL min−1 through 90 cm tubes with a diameter of 3.2 mm (Cole-Palmer
MasterflexTM VersilonTM L/s 06475-16). For 200 mL min−1 , the average
residence time is 3 seconds in the flow-cell and 1 second in a connection tube.
At the back of the GDE, an airflow of 50 mL min−1 was provided. The exit tube
of the gas channel was put into a 30 cm high water column to provide a gas back-
pressure of approximately 30 mbar. GDEx was carried out galvanostatically
at 100 A m−2 (Biologic Biostat SP-300) for 1926 seconds (i.e., equivalent to
the total theoretical addition of 0.002 mol of OH− ). At this current density,
the polarization potential is approximately -0.07 V vs. SHE which guarantees
that only oxygen in the air is reduced (and not, for example, CO2 ) and that
the hydrogen evolution reaction (HER) is avoided [20]. A 3 M KCl Ag/AgCl
reference electrode (Radiometer Analytical REF321) was used and in connection
with the working electrode (i.e., GDE) through a Luggin capillary filled with
500 mM NaCl dilution stock solution. All experiments were carried out at room
temperature (i.e., 22 ◦ C ± 2 ◦ C).

2.3 AT process

The design of the AT reactor is schematically demonstrated in figure 3.2c. AT

®
was carried out with a titrator (Methrohm 702 SM Titrino). The batch reactor
was the same DURAN bottle as described before.
®
A total of 20 mL of a 0.1 M NaOH solution (Titrisol , Merck) was added
at a titration rate of 0.62 mL min−1 . This titration volume and titration
rate were chosen to add a total of 0.002 mol OH− at an equivalent rate of
100% efficient OH− production rate using GDEx current density of 100 mA
m−2 (i.e., the chosen galvanostatic current density for the GDEx process).
The relatively low concentration of NaOH in the titrant was chosen to avoid
concentration gradients in the solution as much as possible. This advantage
goes hand in hand with the disadvantage of increasing the total reaction volume
over time. However, this volume increase does not influence the time axis and
only insignificantly influences the pH values compared to the case of infinitely
small additions of highly alkaline NaOH solutions. The pH measurements were
EXPERIMENTAL 95

carried out equivalently to GDEx. All experiments were carried out at room
temperature (i.e., 22 ◦ C ± 2 ◦ C).

2.4 Analytical measurements

Both during AT and GDEx, the pH was measured every 5 seconds with a
Metrohm 781 pH/ion meter equipped with a Metrohm Unitrode pH electrode.
The pH electrode was calibrated with 4 Merck standard solutions (pH 4.01, pH
7.00, pH 10.00, and pH 12.00).
®
®
The hydrogen peroxide concentration was measured every minute (QUANTOFIX
Relax with QUANTOFIX Peroxide 25 and 100 test strips) during the GDEx
experiments. Additionally, the peroxide concentration was also measured for
a GDEx experiment of a blank 0.5 M NaCl with 0.01 M HCl solution as a
benchmark.

2.5 Post-process treatment and characterization

After completion of the GDEx or AT process, the precipitates were washed and
dried. Washing was done in 5 subsequent washing steps. In the first washing
step, the electrolyte was divided into two samples and centrifuged for 5 minutes
at 10,000 rpm (Hettich Rotina 35). The liquid supernatant was decanted and 80
mL of demineralized water was added to each of the two samples to redisperse
and wash the precipitates. The subsequent two washing steps repeated the
centrifuging, decanting and redispersion steps. In the fourth washing step, the
samples were redispersed into 40 mL of water each and joined into one sample
and only then centrifuged. This allowed for a better collection at the end of
the post-processing. In the last washing step, the 80 mL supernatant of the
single sample was removed and 20 mL of ethanol was added to redisperse the
particles. The last centrifugation was completed and the ethanol supernatant
removed. The humid precipitate was collected from the centrifugation container
and put on a weighing boat and left to dry in a desiccator for 12 hours. The
last step in ethanol was carried out to allow for much quicker drying than if
only water was used, avoiding the unnecessary complications of the particles
being in humid conditions for too long.
The composition of the dried precipitates were characterized with XRD (Bruker
D2 phaser, Cu source Kα,avg = 1.54, line focus, LynxEye detector). The size
and morphology of the dried precipitate were characterized by SEM (Philips
XL 30 FEG).
96 OXIDATION-ASSISTED ALKALINE PRECIPITATION OF NANOPARTICLES USING GAS-DIFFUSION
ELECTRODES

3 Results and discussion

3.1 Colorimetric transitions

During AT and GDEx of solutions with ZnCl2 , MnCl2 , or FeCl2 precipitation


was observed visually (figure 6.2). For AT and GDEx of ZnCl2 white precipitates
could be observed when the pH started to rise suddenly and stage 2 initiated
(figure 6.1). Since most zinc-containing inorganic species are white [108], this
provided little indication of the possible composition. AT and GDEx of MnCl2
demonstrate similar behavior: beige precipitates could be observed at the
beginning of stage 2. However, the color changed from beige to light brown
during the process and the subsequent washing steps. In principle, Mn(OH)2 is
white, but since the higher oxides (i.e., Mn2 O3 , Mn3 O4 , MnO2 ) have a dark
brown or black color, small amounts of them are sufficient to change the color
from white to beige [173,174]. The slightly darker tint of the precipitates formed
in GDEx is attributed to the fact that oxygen traces are unavoidable when using
an air-fed GDE. The change from beige to light brown, both for AT and GDEx
precipitates, indicates further oxidation of the Mn(II) to Mn(III) during the
subsequent washing steps [174]. For FeCl2 the precipitates for AT and GDEx
differed significantly. For AT, dark green precipitates were observed, while
for GDEx dark brown precipitates were observed. Dark green precipitates of
Fe2+ solutions are indicative of Fe(OH)2 precipitation [175]. These precipitates
changed color to light brown during the subsequent washing steps, resulting
from oxidation with atmospheric or dissolved oxygen. Dark brown precipitates
are indicative of Fe3 O4 precipitation, which indicates the oxidation of Fe(II) to
Fe(III) already during the precipitation process. The significant color difference
of both precipitates for the case of FeCl2 was the first indication of the effect of
GDEx.
RESULTS AND DISCUSSION 97

Zn Mn Fe

AT GDEx AT GDEx AT GDEx

Figure 6.2: Photographs of dispersed precipitates collected as-is from the AT


and GDEx processes of ZnCl2 , MnCl2 , or FeCl2 .

3.2 Production of H2 O2

Figure 6.3 compares the H2 O2 concentration measurements during GDEx


experiments of a solution without and solutions with ZnCl2 , MnCl2 , or FeCl2 .
The 100% H2 O2 production efficiency line is demonstrated for benchmarking.
The observed H2 O2 production efficiency (i.e., Faradaic efficiency of H2 O2 )
for the blank solution and the ZnCl2 solution is equal to 24%. This indicates
the inertness of Zn2+ and its precipitates with respect to H2 O2 . In contrast,
for solutions with MnCl2 or FeCl2 , the production efficiency for H2 O2 could
not be determined, because the measured H2 O2 was constant and below the
measurement threshold of 1.47 10−5 mol L−1 (0.5 ppm). This indicates the
high amount of reaction of H2 O2 with the manganese - and iron ions or the
manganese- and iron-containing precipitates.
98 OXIDATION-ASSISTED ALKALINE PRECIPITATION OF NANOPARTICLES USING GAS-DIFFUSION
ELECTRODES

7
blank 100% efficiency
Zn

H2O2 concentration [10-3 mol L-1]


6
Mn
5 Fe

2 24% efficiency

0
0 0.0005 0.001 0.0015
added charge [mol]

Figure 6.3: Comparison of H2 O2 concentration measurements during GDEx


experiments of a blank solution and solutions with ZnCl2 , MnCl2 , or FeCl2 .
The H2 O2 production efficiency (24%) for the case of the blank solution and
the ZnCl2 solution are the same (blue interpolation line). The 100% efficiency
line is indicated for comparison.

3.3 Dynamic pH behavior

The expected shape of the dynamic pH curves was briefly discussed in the
introduction (figure 6.1). Figure 6.4 compares the dynamic pH measurements of
the galvanostatic GDEx experiments with the AT experiments for the solutions
with ZnCl2 , MnCl2 , and FeCl2 . Figures 6.4a and 6.4b compare the results of
AT and GDEx respectively for solutions with ZnCl2 , MnCl2 , or FeCl2 . The
shapes of the dynamic pH curves are all in line with the expectations. Figures
6.4c, 6.4d, and 6.4e compare GDEx and AT for solutions with ZnCl2 , MnCl2 ,
or FeCl2 respectively.
For AT, stage 1 and 2 ended as expected upon the addition of respectively
10−4 mol and 1.1 10−3 mol of OH− (figure 6.4a). Also, the theoretical shape
of the curve corresponds to the experimental shape of the curve for the cases
of Zn and Mn. For Fe, a small deviation of this shape is observable: besides
the main precipitation plateau, a short inclination before the main plateau is
visible (figure 6.4a). This is attributed to the fact that small amounts of Fe3+
may have remained or may have formed in the prepared solution. These Fe3+
RESULTS AND DISCUSSION 99

14
a AT b GDEx
12
10

pH 8
6
4
Zn Zn
2 Mn Mn
Fe Fe
0
0 0.001 0.002 0 0.001 0.002
added NaOH [mol] added charge [mol]
14
c Zn d Mn e Fe
12
10
8
pH

6
4
2 AT AT AT
GDEx GDEx GDEx
0
0 0.001 0.002 0 0.001 0.002 0 0.001 0.002
added NaOH or charge [mol] added NaOH or charge [mol] added NaOH or charge [mol]

Figure 6.4: Comparison of the dynamic pH curves for ZnCl2 , MnCl2 and FeCl2
for AT (a) and GDEx (b). Comparison of the dynamic pH curves for AT and
GDEx for ZnCl2 (c), MnCl2 (d) and FeCl2 (e).

ions tend to precipitate as Fe3 O4 which is known to precipitate at lower pH


than Fe(OH)2 [20].
For GDEx, stage 1 and 2 do not end as expected, even though the shape of
the curves are similar to those of AT (figure 6.4b). Comparison of AT and
GDEx is easier for each metal separately (figures 6.4c, 6.4d, and 6.4e), as such
four major differences are observable: (1) the first inflection point (i.e., the
end of stage 1) appears somewhat later for GDEx than for AT (figures 6.4c,
6.4d, and 6.4e), (2) the second inflection point (i.e., the end of stage 2) appears
much later for GDEx than for AT (figures 6.4c, 6.4d, and 6.4e), (3) while the
value of the precipitation pH for Zn and Mn differs little between AT and
GDEx, the value of the precipitation pH for Fe decreases significantly (figure
6.4e), and (4) the delay of the first inflection point of Fe is higher than for
Zn and Mn (figure 6.4b). Observation (1) and (2) can be discussed in light of
the electrochemical production of OH− and HO− 2 . One could argue that the
100 OXIDATION-ASSISTED ALKALINE PRECIPITATION OF NANOPARTICLES USING GAS-DIFFUSION
ELECTRODES

addition-efficiency of OH− might be lower than 100% due to the production


of HO− 2 . This is correct, however, because the bulk pH remains low enough
during the experiment (pKH2 O2 = 11.62 [20]), the produced HO− 2 will remove
protons by protonating to become H2 O2 . Therefore, the theoretical addition of
OH− can be regarded as 100% efficient, as long as the pH is sufficiently low.
Nevertheless, observations (1) and (2) imply an efficiency of less than 100%. A
closer look at the second inflection point teaches us that the delay is independent
of the metal ion: stage 3 starts simultaneously for each case (figure 6.4b). The
individual comparison of GDEx and AT for each metal case demonstrates that
the delay grows over time (figures 6.4c, 6.4d, and 6.4e). In the case of the first
inflection point, AT and GDEx differ significantly less than in the case of the
second inflection point. Moreover, because of this, the precipitation plateau is
longer than is theoretically possible (i.e., 2 times the amount of metal ion in
solution). All these observations point towards the characteristics of the GDEx
process itself, rather than the characteristics of the electrolytes. One possible
explanation is that the delay is caused by the time-dependent accumulation of
OH− and HO− 2 in the pores of the gas-diffusion electrode. This accumulation
would then cause a high pH in the electrode pores [176]. The built-up of these
species slows down the flux of these species into the bulk of the electrolyte,
which is registered as a delay of the measured pH during GDEx compared to
during AT. The topic of accumulation is discussed in detail in Chapter 7.
Observations (3) and (4) are specific to the case of GDEx of FeCl2 . The
decrease in the precipitation pH indicates a change in the chemical reaction.
This change in chemical reaction is caused by the oxidation of Fe2+ to Fe3+ by
H2 O2 [164, 169]. The occurrence of both Fe3+ and Fe2+ is known to result in
the coprecipitation of Fe2+ and Fe3+ as Fe3 O4 [177]. This composition was also
confirmed by XRD (figure 6.5c). The additional OH− needed to precipitate
Fe3+ instead of Fe2+ is produced during the oxidation reaction of Fe2+ with
H2 O2 . For this reason, the length of the plateau remains unchanged compared
to the cases of Zn and Mn (figure 6.4b). The slightly larger shift of the first
inflection point for Fe (figure 6.4e) compared to Mn and Zn (figures 6.4c and
6.4d), is most probably related to pre-precipitation of Fe3 O4 in the flow cell.
Because of a slightly higher pH in the flow cell, precipitation criteria are met
earlier than in the reservoir. The OH− used for pre-precipitation does not
attribute to an increase in the pH, which thus causes an additional delay.

3.4 Composition of the precipitates

The composition of the precipitates was characterized by XRD. Figure 6.5


compares the AT and GDEx precipitates for the cases of Zn, Mn, and Fe. In the
case of zinc, the XRD pattern could be identified as mainly Zn(OH)2 [178, 179]
RESULTS AND DISCUSSION 101

350 10
a Zn(OH)2 Zn b Mn3O4 Mn c Fe3O4 Fe
9
300 □ ZnO
8
250 7
GDEx GDEx GDEx

intensity [a.u.]

intensity [a.u.]
intensity [a.u.]

6
200
5

150 □□ □
□ □ □ 4

100 Zn(OH)2 AT Mn3O4 AT


3
FeO AT
2
50
1

0 0
15 20 25 30 35 40 45 50 55 60 65 70 15 20 25 30 35 40 45 50 55 60 65 70 15 20 25 30 35 40 45 50 55 60 65 70
2 [] 2 [] 2 []

Figure 6.5: XRD patterns for AT and GDEx for solutions with ZnCl2 (a),
MnCl2 (b) or FeCl2 (c). Only for GDEx of ZnCl2 a phase mixture is observed
and the peaks of ZnO are indicated with a -symbol for clarity.

in both the AT and GDEx case (figure 6.5). However, in the GDEx case, some of
the peaks corresponded to the dehydrated ZnO phase [180] (, figure 6.5a). The
XRD pattern for manganese could be identified as Mn3 O4 [166, 181] for both
the AT and GDEx cases (figure 6.5b). The XRD pattern for iron was identified
as that of FeO [182–184] in the case of AT and as that of Fe3 O4 [182, 185] in the
case of GDEx (figure 6.5c). In the case of zinc, the composition was as expected,
with Zn(OH)2 as the main phase with or without some dehydrated ZnO. In
the case of manganese, the observation of Mn3 O4 rather than Mn(OH)2 is in
line with the color observations during the post-treatment washing steps. In
the case of iron, visual observations of the precipitates after processing already
hinted at the occurrence of a different composition. For AT and GDEx, this
deviation from the expected composition of Fe(OH)2 was attributed to different
phenomena. In the case of AT, the dehydrated FeO showed light orange color,
most likely caused by superficial oxidation to Fe2 O3 , which cannot be observed
with XRD. In the case of GDEx, Fe(II) is oxidized by the H2 O2 produced
during the process. This is in agreement with the readily observed dark brown
precipitates and the fact that no H2 O2 could be measured during the process.

3.5 Size and morphology of the precipitates

Scanning Electron Microscopy (SEM) was used to visualize the size and
morphology of the precipitates (Figure 6.6). In all cases, the precipitate was
highly agglomerated. For the cases of Zn and Mn, the sizes and morphology
of the particles processed by GDEx differ little with those processed by AT
(figure 6.6a and 6.6d for Zn, and figure 6.6b and 6.6e for Mn). In the case of
102 OXIDATION-ASSISTED ALKALINE PRECIPITATION OF NANOPARTICLES USING GAS-DIFFUSION
ELECTRODES

Zn, irregular sheetlike particles were identifiable, albeit difficult. In the case of
Mn, spherical nanoparticles were easily observed. These similarities are in line
with the other observations: similar precipitate color, similar composition, and
similar precipitation behavior observed by dynamic pH measurements. Except
for the fact that no H2 O2 could be measured in the case of GDEx of MnCl2 ,
it appears that the GDEx and AT processes are equivalent in the cases of the
synthesis of Zn(OH)2 and Mn3 O4 nanoparticles, within the confines of our
process conditions. Given that AT is used as a benchmark process for alkaline
precipitation and GDEx is equivalent to AT for our solutions of MnCl2 and
ZnCl2 , GDEx is considered only to elicit alkaline precipitation in these cases,
and not oxidation-assisted alkaline precipitation. The fact that no H2 O2 could
be measured in the case of Mn is not entirely clear, but it might be explained
by the effective catalytic decomposition of H2 O2 on Mn(OH)2 [186].
In contrast to the Zn and Mn cases, for the case of Fe, the particles processed
by AT and GDEx differed significantly in size and morphology (figure 6.6c and
6.6f). While the FeO particles synthesized with AT are tile-shaped and with a
long-axis length in the order of 200 nm (figure 6.6c), the Fe3 O4 nanoparticles
synthesized by GDEx are spherical with a diameter in the order of 20 nm or
lower(figure 6.6f). These particles are only just visible by SEM (inset figure 6.6f).
This large size and morphology difference is a strong indication for the fact that
GDEx elicits oxidation-assisted alkaline precipitation, rather than only alkaline
precipitation. This is in line with our expectations, given our experience with
the Ox-AP of Fe2+ in a chemical Y-junction reactor [164] and the controllable
synthesis of iron oxide nanoparticles with GDEx [169]. In our exploratory
study on AP and Ox-AP [164], the size decrease between both processes was
assigned to improved condensation (olation and oxolation) kinetics as the result
of oxidation of the metal ion during the precipitation process. Nevertheless,
such a large size difference between alkaline precipitation and oxidation-assisted
alkaline precipitation was not yet observed. One possible explanation for this
would be the fact that here, Ox-AP and AP result in precipitates of species with
other compositions, while in the previous work [164] the precipitates had the
same composition. Species with different compositions have different solubility
products. Because Fe3 O4 (Ksp,Fe3 O4 = 10−108.6 [187]) has a much lower solubility
than FeO (Ksp,FeO = 10−14.5 [188]), the supersaturation in the GDEx process
will be significantly higher than in the AT process, as a result of the oxidation
of Fe(II) to Fe(III). It has long been established from classic nucleation theory
that higher supersaturation results in smaller (nano)particles [135, 137, 139].
Therefore, the large size difference, between the nanoparticles synthesized by AT
(eliciting AP) and the nanoparticles synthesized with GDEx (eliciting Ox-AP),
could be explained by the combined effect of the improved condensation kinetics
and the increased supersaturation in the GDEx process.
RESULTS AND DISCUSSION 103

a Zn b Mn c Fe

200 nm AT 200 nm AT 200 nm AT


d GDEx e GDEx f GDEx

50 nm

200 nm Zn 200 nm Mn 200 nm Fe

Figure 6.6: SEM images of the precipitates synthesized with AT (a, b, c) and
GDEx (d, e, f) for the cases of ZnCl2 (a, d), MnCl2 (b, e), and FeCl2 (c, f).
104 OXIDATION-ASSISTED ALKALINE PRECIPITATION OF NANOPARTICLES USING GAS-DIFFUSION
ELECTRODES

4 Conclusions

Gas-diffusion electrocrystallization (GDEx) was benchmarked against alkaline


titration (AT) to demonstrate the potential of GDEx to elicit oxidation-assisted
alkaline precipitation (Ox-AP) of metal (oxy)(hydr)oxide nanoparticles. The
comparison of GDEx with AT for solutions of ZnCl2 , MnCl2 , or FeCl2 showed
that GDEx could only elicit Ox-AP in the case of FeCl2 . In this case,
GDEx synthesized Fe3 O4 nanoparticles with a much smaller size than the
FeO nanoparticles synthesized with AT. The smaller sizes obtained with GDEx
were attributed to the oxidation of Fe(II) to Fe(III) by H2 O2 or HO− 2 , which
had the combined effect of improving condensation kinetics and drastically
increasing the supersaturation, both known as size-determining parameters. In
the cases of ZnCl2 and MnCl2 , AT and GDEx synthesized similar nanoparticles.
Therefore, in these cases, AT and GDEx were considered equivalent processes
and GDEx only elicited alkaline precipitation rather than oxidation-assisted
alkaline precipitation. Our group is currently exploring GDEx into more detail to
further unravel its potential and to expand the library of metal (oxy)(hydr)oxide
nanoparticles that might benefit a size reduction from being produced with
GDEx rather than AT.
Chapter 7

Accumulation of ORR
reaction products in porous
gas-diffusion electrodes

This chapter investigates the delays of the time-dependent pH in the


GDEx process of metal-containing solutions, observed in Chapter 6. These
delays are caused by the accumulation of OH− and HO− 2 in the pores
of the GDE. Because of this, a lower than expected production rate of
OH− causes observable delays in the time-dependent pH measurements
during the GDEx process. This chapter goes into detail about this
phenomenon. A steady-state analytical model of the concentration profile
of OH− and HO− 2 in the GDE was made. Also, a time-dependent
analytical model of the total accumulation of these species in the
GDE was made. Using different experimental procedures, key model
parameters concerning the accumulation of OH− and HO− 2 in the GDE
were determined. Additionally, fitting the model to experimental data
of the observed accumulation made it possible to estimate the effective
diffusion coefficients of OH− and HO−2 in the pores of the GDE.

Contributions of the author were hypothesis development, development of the


steady-state analytical model for the concentration profile of the species in the
GDE, developing the time-dependent analytical model with S. Fransen, carrying
out all experiments, characterizations and data treatment. S. Fransen helped
with the development of the time-dependent analytical model, data treatment

105
106 ACCUMULATION OF ORR REACTION PRODUCTS IN POROUS GAS-DIFFUSION ELECTRODES

®
and everything related to MATLAB . X. Dominguez Benetton and J. Fransaer
helped with funding, discussions, and proof-reading.
This chapter is prepared for submission to Journal of Electrochemical Society as:
Eggermont S.G., Fransen S., Dominguez Benetton X., Fransaer J. Accumulation
of ORR reaction products in porous gas-diffusion electrodes.
ACCUMULATION OF ORR REACTION PRODUCTS IN POROUS GAS-DIFFUSION ELECTRODES 107

Abstract

In this work, the accumulation of OH− and HO− 2 in a porous air-fed gas-diffusion
electrode was investigated. An analytical steady-state model was derived for the
concentration profiles in the GDE and an analytical time-dependent model was
derived for the total accumulation of OH− and HO− 2 in the GDE. Measuring the
bulk electrolyte pH in a one-compartment electrochemical flow-cell reactor, it
was possible to experimentally determine the accumulation of these species in the
GDE. Other experiments were carried out to determine the penetration depth of
the electrolyte in the GDE, the efficiency of hydrogen peroxide production, and
the mass-transfer coefficients in the flow-cell reactor. Using these parameters,
the model and the experimental data of the pH were fitted to obtain a value for
the effective diffusion coefficient of OH− and HO− 2 in the GDE. The analytical
model is limited by its assumption that no migration of OH− and HO− 2 occurs.
This fact, and the occurrence of an undetermined parasitic side-reaction in the
GDE, were considered to be the two main contributing factors for mismatch
between the model and the experiments. Overall, this work succeeded in
demonstrating the existence of and quantifying the accumulation of species
in porous air-fed gas-diffusion electrodes, as well as give an estimate of the
concentration profile of OH− and HO− 2 in such porous electrodes.
108 ACCUMULATION OF ORR REACTION PRODUCTS IN POROUS GAS-DIFFUSION ELECTRODES

1 Introduction

Recently, gas-diffusion electrocrystallization (GDEx) was demonstrated as a


new electrochemical nanoparticle synthesis method [169]. There, GDEx was
carried out with an air-fed porous gas-diffusion electrode (GDE). Such an
electrode can be used to carry out oxygen reduction reactions (ORRs) for the
production of OH− , H2 O2 , and HO− 2 . Metal ions can react with these species in
solution in an (oxidation-assisted) alkaline precipitation process to produce metal
(oxy)(hydr)oxide nanoparticles with controllable size and composition [164, 169].
The two most important ORRs are the 4-electron ORRs and the 2-electron
ORRs. The 4-electron ORRs in acidic and alkaline media are [63]:
acidic : O2 + 4H+ + 4e−
2H2 O (E◦ = 1.223 V) (7.1)

alkaline : O2 + 2H2 O + 4e−


4OH− (E◦ = 0.401 V) (7.2)
The 2-electron ORRs in acidic and alkaline media are [63]:
acidic : O2 + 2H+
H2 O2 (E◦ = 0.695 V) (7.3)

alkaline : O2 + H2 O + 2e−
HO−
2 + OH

(E◦ = −0.065 V) (7.4)

Several other ORRs exist [63], for example, the further reduction of HO− 2 to
OH− , but in general, the most stable generated species are OH− , H2 O2 , and
HO−2 . These species are the focus of this work.

The electrochemical reactions in acidic media happen when the concentration


of H+ is high at the surface of the electrode, while those in alkaline media
happen when the concentration of H+ is low at the surface of the electrode.
The electrode surface is emphasized here because the reactions themselves
change the acidity of the medium at the electrode surface. Take for instance
the 4-electron ORR in acids, which removes 1 H+ ion for every electron put
into the system. Removing H+ decreases its concentration, which in turn makes
the medium close to the electrode more alkaline. When applying very high
current densities in electrolytes with insufficiently high proton concentration,
mass-transfer limitations of the H+ ions quickly result in a shift from the acidic
to the alkaline electrochemical reaction, even though, the pH in the bulk is still
acidic.
In addition to the ORRs, H+ , OH− , H2 O2 , and HO−
2 are involved in two
chemical equilibrium reactions:
H+ + OH−
H2 O (7.5)

H+ + HO−
2
H2 O 2 (7.6)
MODELING THE GDE IN ONE SPATIAL DIMENSION 109

In practice, far from the equilibrium, only one of the species is dominating.
In the case of H2 O, below pH 7, the dominant species is the H+ ion, while
above pH 7, the dominant species is the OH− ion. In the case of H2 O2 , the
dominant species is H2 O2 , below pH 11.62, and the HO− 2 ion, above pH 11.62
(pKH2 O2 = 11.62 [20]). This has interesting consequences for the concentrations
of these species in solution. For example, in a solution of low pH, when OH−
is added, the concentration of OH− will barely increase (in absolute terms),
and instead, the H+ concentration will decrease with an amount (practically)
equivalent to the added amount of OH− ions. Similarly, adding HO− 2 to a
solution of low pH will hardly change the HO− 2 concentration and will instead
increase the H2 O2 concentration, while decreasing the H+ concentration with
amounts (practically) equivalent to the added amount of HO− 2 ions.

Carrying out GDEx, we discovered that the production rate of OH− (i.e., the
removal rate of H+ ) was below the predicted rate (Chapter 3). This is odd
because any electron should lead to an equivalent decrease in the amount of
H+ (due to the water equilibrium). One could argue that part of the electrons
would produce the negative HO− 2 species, which is correct, but as discussed
before, at sufficiently low pH, the HO−2 completely protonates, removing H
+

nonetheless.
We hypothesize that this apparently-lower production rate is caused by the initial
accumulation of OH− in the pores of the GDE. Initially, this work aimed to
model this accumulation and verify the model with experimental data. However,
obtaining realistic values for all of the model parameters proved difficult. Instead,
using the model and empirical data of all but one modeling parameter, the last
modeling parameter could be estimated: the effective diffusion coefficient of the
accumulating species. Therefore, in addition to our overarching work, this work
demonstrates a new approach for estimating the effective diffusion coefficient of
OH− and HO− 2 in air-fed porous gas-diffusion electrodes. Adaptations of the
method are believed to be useful for obtaining effective diffusion coefficients of
other species in other porous electrodes and reactor designs.

2 Modeling the GDE in one spatial dimension

The focus of this work is discovering accumulation phenomena in the porous


GDE by using observations of the concentration of accumulating chemical species
in the bulk. The accumulation of chemical species in the GDE is equal to the
integration of the concentration over the hydrophilic pore volume. Because
of this integration, modeling the average concentration along the thickness
in 1 dimension, and considering homogeneity in the other dimensions suffices
110 ACCUMULATION OF ORR REACTION PRODUCTS IN POROUS GAS-DIFFUSION ELECTRODES

(i.e., an exact 3-dimensional model is unnecessary). The same holds for the
diffusion layer at the surface of the GDE. Figure 7.1 demonstrates a schematic
representation of the GDE and the diffusion layer.

𝜕𝑐𝑖 𝑤𝐺𝐷𝐸 ,𝑡
𝜕𝑐𝑖 0, 𝑡 𝜖𝜓𝒟𝑖 = 𝑘𝑖 𝑐𝑖 𝑤𝐺𝐷𝐸 , 𝑡 − 𝑐𝑖𝑏 𝑡
=0 𝜕𝑥
𝜕𝑥
wetted active carbon layer
gas−diffusion layer

diffusion layer
of GDE

𝜕𝑐𝑖 𝑥, 𝑡 𝜕 2 𝑐𝑖 𝑥, 𝑡
− 𝒟𝑖 = 𝑃𝑖 − ෍ 𝑅𝑖,𝑗
𝜕𝑡 𝜕𝑥 2
𝑗

𝑥=0 𝑥 = 𝑤𝐺𝐷𝐸

Figure 7.1: Schematic figure of the model geometry indicating the governing
equation and the boundary conditions in the porous active carbon layer of the
GDE. The curve demonstrates the shape of the expected profile of the OH−
concentration.

The GDE used in the experiments is a commercially available VITO CoRE ®


GDE [58]. This GDE consists of a hydrophobic PTFE layer and an active
carbon layer with a stainless steel current collector. The necessary details of
the electrode are discussed in the experimental section. In this work, the model
is limited to the flooded part of the active carbon layer (i.e., the hydrophilic
porosity), because the interest is in the accumulation of the chemical species in
the liquid phase (figure 7.3), derived from electrochemical reactions.
The model requires the definition of some terms regarding the volume and
surface of the electrode. The total electrode volume within the wetted perimeter
Vtot [m3 ] is defined as:

Vtot = Vmetal + Vcarbon + Vphob + Vphil = wGDE Ageo (7.7)

with Vmetal [m3 ] the volume of the stainless steel mesh, Vcarbon [m3 ] the volume
of the carbon and PTFE in the porous media, Vphob [m3 ] the volume of the
hydrophobic pores, Vphil [m3 ] the volume of the hydrophilic pores filled with
liquid electrolyte, wGDE [m] the electrolyte penetration depth in the GDE,
and Ageo [m2 ] the total geometric electrode surface area. Because only the x
dimension is considered in the model, it is necessary to define an average surface
MODELING THE GDE IN ONE SPATIAL DIMENSION 111

through which species in solution can move. This is the average surface area of
the hydrophilic pores Aphil , projected on the yz-plane. Aphil can be related to
the total geometric electrode surface area Ageo [m2 ], which is defined as:

Ageo = Ametal + Acarbon + Aphob + Aphil (7.8)

with Ametal [m2 ] the average projected surface area of the stainless steel mesh,
Acarbon [m3 ] the average projected surface area of the carbon and PTFE in the
porous media, Aphob [m3 ] the average projected surface area of the hydrophobic
pores, Vphil [m3 ] the average projected surface area of the hydrophilic pores
filled with liquid electrolyte. The hydrophilic porosity of the GDE is defined as:
Vphil
= (7.9)
Vphob + Vphil + Vcarbon

The portion of the electrode covered by porous carbon/PTFE (i.e., the total
volume of the electrode minus the mesh), the porous media coverage, is defined
as:
Vphob + Vphil + Vcarbon
ψ= (7.10)
Vtot
The hydrophilic pore volume is thus related to the total volume via the following
relationship:
Vphil = ψVtot (7.11)
A similar calculation could be made for the surface area, however, because of
the reduction to 1 dimension, the equivalent for the surface area can be written
as:
Aphil = ψAgeo (7.12)

The model assumes the electrolyte in the active layer to be stagnant, without
convective mass transport. In addition, the model assumes that the electrolyte
is well-supported with ions that do not participate in the (electro)chemical
reactions, as such the current is considered to be carried only by the ions in the
supporting electrolyte. In a well-supported electrolyte, the potential gradient
in the electrolyte is small, electromigration can be neglected for species with a
low concentration compared to the concentration of the ions of the supporting
electrolyte and the Nernst-Planck and Poisson equations can be simplified,
reducing the problem to that of diffusion-only [189]. The model considers 4
species: H+ , OH− , HO− 2 , and H2 O2 . The general 1D mass-balance equation for
a species i in the porous GDE can be stated as:
P
Ri,j
∂ci ∂ 2 ci Pi j
− Di 2 = − (7.13)
∂t ∂x Vphil Vphil
112 ACCUMULATION OF ORR REACTION PRODUCTS IN POROUS GAS-DIFFUSION ELECTRODES

with ci [mol m−3 ] the concentration of species i and i equal to H+ , OH− , HO− 2,
or H2 O2 , Di [m2 s−1 ] the effective diffusion coefficient of species i, P i [mol
s−1 ] the production rate of species i and
P
Ri,j the sum of the individual
j
homogeneous chemical reaction rates Ri,j of species i with species j. The
species-specific production
Pterm Pi can be written as a function of the total
production term Ptot = Pi [mol s−1 ], which is a function of the applied
i
current Iapp [A]:
ηi Iapp
Pi = ηi Ptot = (7.14)
F
with F Faraday’s constant and ηi [-] the efficiency factor that expresses the
percentage of the current that is used to produce species i. Looking more closely
to the reactions in equations (7.1), (7.2), (7.3), and (7.4), the maximum value
of ηOH− is 1, because if only the 4-electron ORR occurs, each electron is used to
produce OH− . The maximum value of ηH2 O2 or ηHO− is 0.5 because, if only the
2
acidic or the alkaline 2-electron ORR occur, each electron only yields 0.5 H2 O2
or HO− 2 . Given that in reality a linear combination of the 4-electron ORR, the
2-electron ORR, and possibly some other reactions takes place, the values of ηi
will need experimental validation.
In light of this, the model assumes that the current produces only OH− and
HO− 2 with a certain efficiency and that H
+
and H2 O2 do not participate in
the electrochemical reaction and therefore have no production term. This is a
reasonable assumption if the concentration of H+ is low. However, two chemical
equilibrium reactions are allowed to occur (equation (7.5) and (7.6)), which
couple the concentrations of H+ , OH− , HO− 2 , and H2 O2 . These equations
imply
the
existence
of two homogeneous reaction terms Ri,j : RH+ ,OH− = RH2 O and
RH+ ,HO− = RH2 O2 . Equation (7.13) thus specifies a system of four equations

2

with 6 unknowns: [H+ ], [OH− ], [HO−


2 ], [H2 O2 ], and RH2 O and RH2 O2 . To solve
the system, two remaining equations that define the chemical equilibria are
added (equations (7.16) and (7.17)). The system of differential equations is
MODELING THE GDE IN ONE SPATIAL DIMENSION 113

thus written as follows:


∂ H+ ∂ 2 H+
   
− DH+ = −RH2 O − RH2 O2
∂t ∂x2  

∂ OH− 1 − ηHO− I (x)
  2
 
∂ OH
− DOH− = 2
− RH2 O
∂t ∂x2 ψAgeo wGDE F

∂ HO−
 
∂ 2 HO−
 
ηHO− I (x) (7.15)
2 2
− DHO− = 2
− RH2 O2
∂t 2 ∂x2 ψAgeo wGDE F

∂ [H2 O2 ] ∂ 2 [H2 O2 ]
− DH2 O2 = RH2 O2
∂t ∂x2

and supplemented with two equilibrium equations:

KH2 O = H+ OH−
  
(7.16)

HO−
 +  
H 2
KH2 O2 = (7.17)
[H2 O2 ]
In addition to the equilibrium equations, solving the differential equations
requires an initial condition and two boundary conditions for each species.
Figure 7.1 shows a scheme with the model and the boundary conditions. The
initial condition is:
ci (x, 0) = ci,0 (7.18)
with ci,0 [mol m−3 ] the initial concentration of species i everywhere in the liquid.
The first boundary condition is a zero-flux boundary condition. It is defined
at the boundary between the liquid phase and the gas phase (i.e., the internal
wetted boundary in the GDE) because the chemical species are assumed not to
move into the gas phase. The first boundary condition (at x = 0, figure 7.1) is
therefore defined as:
∂ci (0, t)
− Di =0 (7.19)
∂x
The second boundary condition is defined at the surface of the electrode facing
the bulk of the flow-cell reactor (at x = wGDE , figure 7.1). At this boundary,
the number of species coming from the pores needs to be equal to the amount
of species going into the bulk solution:

Ap Npi (wGDE , t) = Ageo Ndl


i (wGDE , t) (7.20)

with Npi [mol m−2 s−1 ] the flux with respect to the pore surface and Ndl i [mol
m−2 s−1 ] the flux term with respect to the total surface contacting the diffusion
114 ACCUMULATION OF ORR REACTION PRODUCTS IN POROUS GAS-DIFFUSION ELECTRODES

layer. Equation (7.20) can be rewritten, using the following definitions:

∂ci (wGDE , t)
Nip (wGDE , t) = −Di (7.21)
∂x
Nidl (wGDE , t) = ki ci (wGDE , t) − cbi (t)

(7.22)

with ki [m s−1 ] the mass transfer coefficient of species i, and cbi [mol m−3 ]
the concentration in the bulk solution. Remark that, although no convective
mass transfer is occurring in the porous electrode, equation (7.22) introduces a
dependency on the convective mass transport in the flow-cell reactor via the
mass transfer coefficient. Using previous definitions (equations (7.12), (7.20),
(7.21), (7.22)), the second boundary condition simplifies to:

∂ci (wGDE , t) ki 
ci (wGDE , t) − cbi (t)

=− (7.23)
∂x ψDi
With the initial condition and boundary conditions defined, the system
of equations (7.15) could be solved numerically but it needs additional
simplification to be solved analytically. An analytical model is preferred
over a numerical model in this work because the focus is on determining the
contributions of the different key parameters to the accumulation of chemical
species in the GDE. This is more easily achieved with an analytical model than
with a numerical model.

2.1 Analytical steady-state model of concentrations in the


GDE

To solve the system of equations (7.15), (7.16), and (7.17), analytically, it is


helpful to make the problem time-independent by solving for the steady-state
instead. This removes the first term in the four differential equations. It also
requires the bulk concentration to be time-independent

cbi (t) = ci,0 (7.24)

with ci,0 [mol m−3 ] the initial bulk concentration as well as at any position in
the GDE. Moreover, the chemical equilibria introduce a large deal of complexity.
Assuming working conditions away from the water equilibrium (i.e., below H+
concentrations of 10−4 mol m−3 or thus above pH = 7), the water reaction
would have a negligible effect on the values of the OH− concentration and
the reaction terms could be neglected. The H+ concentration could then be
calculated from the H2 O equilibrium alone. As a consequence of working at low
initial H+ concentration, and assuming no H2 O2 to be present initially, also
MODELING THE GDE IN ONE SPATIAL DIMENSION 115

the reaction term of H2 O2 could be neglected. Equivalent to the water case,


the H2 O2 concentration could be calculated from the H2 O and H2 O2 equilibria.
Four equations remain in the system:
 
2 − 1 − η Iapp
 
d OH −
HO2
−DOH− =
dx2 ψAgeo wGDE F

d2 HO− ηHO− Iapp


 
2
−DHO− 2 = 2
2 d x ψAgeo wGDE F
(7.25)
K
 H2−O
 +
H =
OH

KH2 O HO−
 
2
[H2 O2 ] =
KH2 O2 OH−
 

Using equations (7.25) and the boundary conditions equations, (7.19) and (7.23),
the general analytical solution for the steady-state concentration of i equal to
OH− or HO− 2 is found to be given by:
" 2 #!
2

Pi wGDE ψ 1 wGDE x
ci,ss (x) = ci,0 + + 1− (7.26)
Vphil ki 2 Di wGDE

The concentration of species in the GDE is the sum of an initial term, a reactor
design term (i.e., defined by the mass-transfer coefficient ki ) and a GDE design
term (see figure 7.2). The initial term defines the initial concentration in the
pores. The reactor design term defines the concentration that can be reached
at the surface of the electrode given a certain mass-transfer coefficient. The
GDE design term determines how high the concentration can increase above
this surface concentration value, given the electrolyte penetration depth of the
GDE and the effective diffusion coefficient. In addition, both the reactor design
term and the GDE design term depend linearly on the Faraday-compensated
Pi η Iapp
current in the pores Vphil = Viphil F in front of the brackets. The concentration
in the GDE is a quadratic function of the spatial variable x with a maximum
value at the back of the wetted part of the active carbon layer.
116 ACCUMULATION OF ORR REACTION PRODUCTS IN POROUS GAS-DIFFUSION ELECTRODES

2 2
𝑃𝑖 𝑤𝐺𝐷𝐸 𝜓𝜖 1 𝑃𝑖 𝑤𝐺𝐷𝐸 𝑤𝐺𝐷𝐸
𝑐𝑖 𝑥 = 𝑐𝑖,0 + + 1−
𝑉𝑝ℎ𝑖𝑙 𝑘𝑖 2 𝑉𝑝ℎ𝑖𝑙 𝒟𝑖 𝑥
wetted active carbon layer
gas−diffusion layer

diffusion layer
2
1 𝑃𝑖 𝑤𝐺𝐷𝐸
of GDE

2 𝑉𝑝ℎ𝑖𝑙 𝒟𝑖

𝑃𝑖 𝑤𝐺𝐷𝐸 𝜓𝜖
𝑐𝑖,0 𝑉𝑝ℎ𝑖𝑙 𝑘𝑖
𝑥=0 𝑥 = 𝑤𝐺𝐷𝐸

Figure 7.2: Schematic figure of the model geometry indicating the different
contributions to the steady-state concentration of a produced species in the
GDE

Although this notation allows easy definition of the different contributions, a


more general notation is possible:
"  2 #!
Pi 1 1 1 x
ci,ss (x) = ci,0 + τi + − (7.27)
Vphil Shi 2 2 wGDE

with τi [s] the characteristic diffusion time in the GDE and Shi [-] a Sherwood-like
number equal to:
w2
τi = GDE (7.28)
Di
ki wGDE
Shi = (7.29)
ψDi
The characteristic diffusion time τi of chemical species i is an indicative measure
of the time necessary to diffuse over a distance equal to the penetration depth.
The dimensionless number Shi has the appearance of a Sherwood number and
is equal to the ratio of the convective mass transfer in the bulk and the diffusive
mass transfer term in the GDE. If it is large, it indicates that convective mass
transfer in the bulk is much faster than diffusive mass transfer in the GDE.
Because the convective mass transfer term and the diffusive mass transfer term
relate to different locations (i.e., the bulk and the GDE, respectively), Shi is not
an authentic Sherwood and is therefore defined as a Sherwood-like number. The
concentrations in the bulk depend linearly on the production rate Pi and the
characteristic diffusion time τi , and depend inversely on the pore volume Vphil
MODELING THE GDE IN ONE SPATIAL DIMENSION 117

and the Sherwood-like number Shi (to a lesser extent). Integrating equation
(7.27) over the hydrophilic pore volume of the GDE gives an expression for
the steady-state accumulation of species in the pores of the GDE that contain
electrolyte:
Rw
nGDE
RRR
i,ss = c (x) dV = ψAgeo 0 GDE ci,ss (x) dx
Vphil i,ss

  (7.30)
1 1
= ci,0 Vphil + Pi τi +
Shi 3
With this notation, it is clear that higher production (i.e., higher applied
current) and higher characteristic diffusion times (i.e., thicker electrodes or
lower effective diffusion coefficients) result in more accumulation in the electrode.
Additionally, the convective behavior of the flow-cell reactor becomes increasingly
less important for values of Shi above 3.

2.2 Analytical time-dependent model of the total accumula-


tion in the GDE

Using pH measurements (see experimental section), an indirect indicative


measure of the time-dependent accumulation of OH− in the pores of the GDE
is available. To make optimal use of these observations, a time-dependent
model for this accumulation imposes itself. Here, the aim is to simulate this
time-dependent behavior using an analytical model. The envisioned analytical
model should be able to easily demonstrate the overall expected behavior as
well as the role of the different parameters on this behavior.
The time-dependent equation without homogeneous reactions, taken far away
from the H2 O and H2 O2 equilibria (no reaction terms) can be written as:

∂ci ∂ 2 ci Pi
− Di 2 = (7.31)
∂t ∂x Vphil

This equation resembles the heat equation with a constant source term and is
often solved with separation of variables. Although this approach can potentially
give the exact solution, its form is needlessly complex and does not comply
with a vision to provide a straightforward solution. Because the very short
time range is not of special interest, but more the overall behavior, and more
specifically the times approaching the steady-state, Laplace transforms were
opted to solve the equation instead. As such, while solving the equations in the
Laplace domain, it is possible to solve the equations for the asymptotic behavior
(i.e., the limiting situation when s approaches zero, more or less equivalent to
118 ACCUMULATION OF ORR REACTION PRODUCTS IN POROUS GAS-DIFFUSION ELECTRODES

the time t going to infinity). The complete derivation is available in appendix


E. The solution of the time-dependent accumulation is:
  
 
1 1 t
nGDE
i,ss (t) = ci,0 Vphil + Pi τi + 1 − exp −   
Shi 3 τ 1
+1
i Shi 2
(7.32)
For times t going to infinity, the value of ni,GDE indeed reaches that of nGDE i,ss
(equation (7.30)). The time-dependent behavior is exponential and similar
to the steady-state value of the accumulation, the time-dependent behavior
depends mostly on the characteristic diffusion time τi . Similar to the steady-
state case, the effect of the value of Shi is decreasingly less important for values
above 2 (instead of 3). Although Shi can be estimated more realistically using
experiments (see below), a quick estimation learns that Shi will be in the range
of:
ki wGDE 5 10−5 300 10−6
Shi = ≈ ≈ 200 (7.33)
ψDi 0.5 0.29 5 10−10
This is much higher than 2 or 3, and therefore the Sherwood-like number will
have limited influence on the accumulation of species in the GDE, as long as no
fluid flow exists in the pores of the GDE.
The characteristic diffusion time has an important role in determining the
value of the steady-state accumulation as well as in the exponential behavior to
reach this steady-state value (equation (7.32)). The value of the characteristic
diffusion time strongly depends on the physical characteristics of the GDE (i.e.,
the penetration depth and effective diffusion coefficient). These characteristics
are inherently empirical, because of the stochastic variation on the production
of GDEs based on carbon powder/PTFE mixtures. Combing the model with
rightly designed experiments provides the necessary tools to estimate this
characteristic diffusion time and by extension the effective diffusion coefficient.

3 Experimental

The analytical solution of the time-dependent accumulation (equation (7.32))


demonstrated the three important parameters that determine this accumulation:
the production rate (Faraday-compensated current) Pi , the characteristic
diffusion time τi , and the Sherwood-like number Shi . The Faraday-compensated
volumetric current depends on the applied current and the HO− 2 efficiency.
The characteristic diffusion time and the Sherwood-like number depend on the
penetration depth wGDE , the mass transfer coefficient ki , the effective diffusion
coefficient Di , the porous media coverage ψ, and the hydrophilic porosity . The
EXPERIMENTAL 119

latter is known to be approximately 0.29 from previous characterizations [58].


The other parameters are unknown and need empirical estimations. Each
parameter is estimated by a different experiment. The experimental methods
are shortly described in this section.

3.1 Electrolyte solutions

For experiments to determine the HO-2 efficiency. 0.50 M NaCl solutions


with different initial proton concentration were made from a 5.00 M NaCl
(Merck) stock solution. The 5.00 M NaCl stock solution was made by adding
292.2 g NaCl salt to the required amount of demineralized water to complete a
1 L volumetric flask. The 0.50 M NaCl solutions were made by adding 0.100 L
of the 5.00 M NaCl stock solution and 0.050 L, 0.020 L, or 0.010 L of 1 M HCl
solution (Titrisol) to the required amount of demineralized water to complete a
1 L volumetric flask. This gave solutions with 0.050 M, 0.02 M, and 0.01 M
initial proton concentration.

For experiments to determine the penetration depth. A 1 M CuSO4 solution


of 0.500 L was prepared by dissolving 124.84 g of CuSO4 .5H2 O in 0.400 L

®
demineralized water. This solution was added to 0.050 L of 0.05 M H2 SO4
(Titrisol ) and the required amount of demineralized water to complete a 0.500
L volumetric flask. A copper salt was chosen because back-scattered electrons
in a scanning electron microscope give good contrast with the carbon phase in
the active carbon layer of the GDE.

For experiments to determine the mass transfer coefficient. The 0.001 M


iodide/triiodide stock solution was made by dissolving 0.127 g of I2 (Acros
Organics), 0.415 g of KI (Merck) and 25.25 g of KNO3 (Merck) in demineralized
water in a volumetric flask of 0.5 L.

For experiments to determine the accumulation and the effective diffusion


coefficient. A 2 M KNO3 stock solution of 1 L was prepared by dissolving
202.21 g KNO3 (VWR international) in the required amount of demineralized
water to complete a 1 L volumetric flask. In addition, for these experiments,
in-house calibration solutions were made for the pH meter. The first calibration
solution with a concentration of 0.01 M of H+ was made by dissolving 20.2
g of KNO3 (VWR international) and 0.001 L HCl solution (Titrisol) in the
required amount of demineralized water to complete a volumetric flask of 0.100
L. The 4 other calibration solutions (i.e., 0.005 M, 0.0025 M. 0.001 M, and
120 ACCUMULATION OF ORR REACTION PRODUCTS IN POROUS GAS-DIFFUSION ELECTRODES

0.0001 M HCl) were made by diluting the solution of 2 M KNO3 with 0.01 M
HCl with the 2 M KNO3 stock solution in the correct ratios. Using pH values
of 2, 2.301, 2.602, 3, and 4 to calibrate the 2 M KNO3 solutions of respective
HCl concentrations of 0.01 M, 0.005 M, 0.0025 M, 0.001 M, and 0.0001 M, it
is possible to artificially force the pH meter to measure proton concentrations,
rather than proton activities, in the highly supported electrolyte.

3.2 Gas-diffusion electrode

The gas-diffusion electrode was a commercially available VITO CoRE gas- ®


diffusion electrode. It separates the gas compartment from the catholyte
compartment. It consists of two different layers, cold-rolled together with a
stainless steel mesh. The gas-side layer is made of porous PTFE and has
an average thickness of approximately 300 µm. The active layer facing the
electrolyte is made of a 80/20 (% wt) mixture of active carbon/PTFE and
has an average thickness of approximately 300 µm. The stainless steel mesh
is pushed into the active layer. Cross-section samples were made by cutting
a piece from a used or an unused GDE sample and embedding it in epoxy
(Technovit 4004). The epoxy sample was ground and polished in several steps.
For SEM photographs (Philips XL 30 FEG), the polished surface was coated
with 2 nm Pt (Quorum Q150T S). Figure 7.3 demonstrates a front view and a
cross-section of the unused GDE.

3.3 Flow-cell reactor and process conditions

The default configuration. The default configuration of a GDEx reactor is a


two-compartment electrochemical flow-cell reactor. Figure 3.2a demonstrates
a typical flow-cell reactor set-up and figure 3.2b a scheme of the flow-cell

®
reactor. The flow-cell reactor compartment volumes are 0.010 L. The cathode
was a VITO CoRE gas-diffusion electrode. The anode was a flat platinized
tantalum plate. The geometric surface area of the anode and cathode were
10 cm2 . The anolyte and catholyte compartment are separated with an anion
®
exchange membrane (FUMASEP FAP-4130-PK). This type of membrane and
a sufficiently high chloride concentration allow a closed electric circuit. The
anolyte and catholyte recirculation reservoirs were 250 mL DURAN laboratory
bottles, with additional flask openings. The cap has openings to allow argon
purging. The reservoir solutions were stirred with a magnetic stirring bar of
3 cm on an IKA RCTbasic S1 stirrer at rotation level 7 (at rotation rates of
approximately 700-800 rpm). A double-headed peristaltic pump (Cole-Palmer
MasterflexTM L/s 7551-00 with Easyload II 77200-60 pump heads) was used to
EXPERIMENTAL 121

front side view


1 mm

stainless steel mesh active carbon

front side
electrolyte side
active carbon layer stainless steel mesh

model

porous gas-diffusion layer

0.2 mm gas side


cross-section

®
Figure 7.3: SEM photographs of the front view and cross-section of the VITO
CoRE electrode. The front side view demonstrates the stainless steel mesh
and the active carbon phase. The cross-section demonstrates the PTFE foam
layer (lighter phase), the carbon/PTFE active layer (darker phase) and the
stainless steel mesh (very light disks). The model is limited to the carbon/PTFE
active layer as indicated by the white line.
122 ACCUMULATION OF ORR REACTION PRODUCTS IN POROUS GAS-DIFFUSION ELECTRODES

pump the solutions in anolyte and catholyte around at 500 mL min−1 through
90 cm tubes with a diameter of 3.2 mm (Cole-Palmer MasterflexTM VersilonTM
L/s 06475-16). For 500 mL min−1 , the average residence time is 2.4 seconds
in the flow-cell and 0.8 seconds in a connection tube. The GDE was fed with
air at its hydrophobic gas-diffusion layer side with an airflow of 0.100 L min−1 .
The air back-pressure of approximately 30 mbar was provided by putting the
exit tube of the gas channel into a 30 cm high water column. All experiments
were carried out at room temperature (i.e., 22 ◦ C ± 2 ◦ C).

The one-compartment configuration. To determine the accumulation in the


GDE, a one-compartment configuration was used rather than a two-compartment
configuration. In the one-compartment configuration, the anion-exchange
membrane and the anolyte compartment and anolyte reservoir are removed,
leaving only one compartment and reservoir.

The electrochemical experiments. The electrochemical experiments were


carried out with a potentiostat (Biologic Biostat SP-300). The reference was
a 3 M KCl Ag/AgCl reference electrode (Radiometer Analytical REF321) in
connection with the working electrode (i.e., GDE) through a Luggin capillary
filled with 2 M KNO3 stock solution.

3.4 Determining the HO-2 production efficiency

In a two-compartment electrochemical flow-cell reactor set-up with an air-


fed GDE as cathode, the HO− 2 production efficiency ηHO− was calculated
2
from H2 O2 concentration measurements in the catholyte recirculation reservoir

® ®
every 5 seconds. The H2 O2 concentration was measured every minute using a
QUANTOFIX Relax with QUANTOFIX Peroxide 25 and QUANTOFIX ®
Peroxide 100 test strips. 0.100 L solutions with 0.050 M, 0.02 M and 0.01 M
initial proton concentration were respectively treated with current densities
of 500 A m−2 , 200 A m−2 , and 100 A m−2 . As such, each experiment could
be carried out for 1000 seconds before reaching high values of the pH (due
to the co-production of OH− in the air-fed GDE). The slope of the measured
d [H2 O2 ]
concentration of H2 O2 as a function of time [mol s−1 ], is equal to the
dt
H2 O2 production rate:

d [H2 O2 ]
PH2 O2 = (7.34)
dt
EXPERIMENTAL 123

Because of the sufficiently low pH in the bulk, all HO− 2 reaching the bulk would
react to H2 O2 as a result of the equilibrium reaction (equation (7.6)). Therefore
the production rates of H2 O2 and HO− −
2 are equivalent. The HO2 production
efficiency ηHO− is then calculated as follows:
2

PHO− PH2 O2 F
ηHO− = 2
= (7.35)
2 Ptot Iapp

3.5 Determining the electrolyte penetration depth

The electrolyte penetration depth was investigated by SEM photographs of


the cross-section of a pre-treated GDE. The GDE was pre-treated by flowing
CuSO4 at the catholyte side for 15 minutes at 100 L min−1 . This GDE was
taken from the flow-cell and dried in a desiccator for 24 hours. The excess salt
on top of the electrode was wiped off. The cross-section was prepared in the
same way as the unused GDE, explained before.

3.6 Determining the value of the mass-transfer coefficients

An estimation of the mass-transfer coefficient of the flow-cell reactor design was


made by a standard limiting-current method using iodide/triiodide solutions
[190] in the default two-compartment flow-cell configuration. Assuming no
electrolyte flow is flowing in the pores of the electrode, the mass-transfer
coefficient of a configuration with the GDE or with a flat plate should be
equivalent. Therefore, to avoid unnecessary side reactions in the GDE, for
these experiments, the cathode was a platinized tantalum plate, instead of
the GDE. A potential of 0 V vs. Ag/AgCl (3 M KCl) was determined
to carry out chronoamperometric measurements. For different flow rates,
chronoamperometric measurements of 60 seconds were carried out with 0.10 L
of 0.001 M iodide/triiodide solution. The current was measured every 1 second.
The average current of the last 20 seconds was taken as the current value for
the chosen flow rate. The reaction at the cathode is known to be:

I− −
3 + 2e
3I

(7.36)

The mass-transfer coefficient of triiodide was then calculated from the limiting
current as [190]:
q
Ilim
kIq− = (7.37)
3 2Fc∞,I− Ageo
3
124 ACCUMULATION OF ORR REACTION PRODUCTS IN POROUS GAS-DIFFUSION ELECTRODES

with kIq− [m s−1 ] the convective mass-transfer coefficient of triiodide for a given
3
q
flow rate q [m3 s−1 ], Ilim [A] the measured limiting current for a given flow
rate q and c∞,I− [mol m−3 ] the bulk concentration of triiodide. To recalculate
3
the mass-transfer coefficient of triiodide to OH− and HO−
2 , the definition of
the dimensionless average Sherwood number Shavg [-] of the electrochemical
flow-cell can be used:

kiq de kIq− de q
kOH− de
q
kHO− de
Shavg = = 3
= = 2
(7.38)
Di DI− DOH− DHO−
3 2

with kiq the mass-transfer coefficient of species i for a given flow rate q, de [m]
a characteristic length of the flow-cell reactor, and Di [m s−1 ] the diffusion
q q
coefficient of species i. kOH− and k
HO−
can be estimated for different flow rates
2
using the measured values of kIq− and literature values for DI− = 9.60 10−10
3 3

m2 s−1 [190], DOH− = 4.56 10−9 m2 s−1 [191] and DHO− = 1.71 10−9 m2 s−1
2
(roughly estimated by equalizing it to the value ofDH2 O2 from [192] due to the
lack of reliable sources for HO−
2 ).

3.7 Determining the accumulation of OH- and HO-2 in the


GDE and their effective diffusion coefficients

To estimate the accumulation of OH− in the pores of the GDE, the pH of the
bulk electrolyte was measured in a one-compartment configuration. Experiments
were conducted with 0.050 L solutions of the 2 M KNO3 stock solution for
different values of the current densities Iapp /Ageo of 50 A m−2 , 100 A m−2 , 200
A m−2 , 350 A m−2 , and 500 A m−2 . Because of the electrolyte choice of 2 M
KNO3 , it is assumed that no side reactions occurred (e.g., Cl2 evolution) and
therefore assumed a 100% efficiency for the production of H+ at the anode (i.e.,
every electron that passes through the electrochemical cell yields a proton at
the anode). At the GDE, the cathode, it is assumed that the ORRs produce
OH− and HO− 2 . Given a sufficiently low pH in the bulk electrolyte (below
pH 7), every such species would subsequently remove an H+ as the result
of their equilibrium reactions (equations (7.5) and (7.6)). The removal and
production rates of protons at anode and cathode, respectively, are theoretically
equivalent. Because of the one-compartment configuration, given the hypothesis
of accumulation of species in the GDE, an initial mismatch would exist between
the number of protons reaching the bulk via production at the anode and the
amount protons exiting the bulk via the equilibrium reactions with the OH− and
HO− 2 ions produced at the cathode (GDE). This mismatch would be observed
by a decrease in the pH (increase of the proton concentration) in the bulk
RESULTS AND DISCUSSION 125

Table 7.1: Summary of the measured or calculated parameters used in the


analytical models.
current density [A m−2 ]
parameters 50 100 200 350 500
ηHO− [-] - 0.088 0.127 - 0.184
2
wGDE [10−6 m] 300

−5
500
kOH − [10 m s−1 ] 37.5

−5
500
kHO − [10 m s−1 ] 14.1
2
DOH− [10 −10
m s2 −1
] 2.36 2.31 3.05 2.87 3.09
DOH−
δD = [-] 2.75
DHO−
2
E [V] 0.041 0.077 0.132 0.185 0.242

electrolyte. This pH was measured in the recirculation reservoir every 5 seconds


with a Metrohm 781 pH/ion meter equipped with a Metrohm Unitrode pH
electrode. The pH electrode was calibrated with in-house calibration solutions
(see electrolyte solutions section) so the pH could be easily recalculated to the
time-dependent H+ concentration (i.e., cH+ = − log pH). If the assumptions are
correct, this time-dependent H+ concentration multiplied by the bulk electrolyte
volume is equivalent to the sum of the amounts of OH− and HO− 2 accumulated
in the GDE. Given the values of the other model parameters, the effective
diffusion coefficient could be estimated by using it as the fitting parameter
between model and experiment. More details of certain data treatments are
given in the results and discussion section.

4 Results and discussion

The analytical model demonstrated the important parameters that determine


the concentration and accumulation of OH− and HO− 2 in the GDE. First, the
results of the experimental estimations of these parameters are demonstrated.
Subsequently, the experimental data for the bulk accumulation of H+ are
demonstrated and after elaborating on the data treatment, the fit of the
experimental data and the model is used to estimate the effective diffusion
coefficients of OH− and HO− 2 . Thereafter, these results are used to estimate the
concentration profiles of OH− and HO− 2 in the GDE. The obtained parameters
are summarized in table 7.1.
126 ACCUMULATION OF ORR REACTION PRODUCTS IN POROUS GAS-DIFFUSION ELECTRODES

0.25

HO2- production efficiency  HO - [-]


2
0.2

0.15

0.1

0.05

0
10 100 1000
𝐼𝑎𝑝𝑝
current density [A m-2]
𝐴𝑔𝑒𝑜

Figure 7.4: Hydrogen peroxide efficiency as a function of the current density.


The linear interpolation corresponds to ηHO− = 0.138 log (Iapp /Ageo ) − 0.190.
2

4.1 Hydrogen peroxide efficiency

The average efficiency for the production of HO− 2 , ηHO− , was calculated from
2
the bulk H2 O2 concentration measurements during galvanostatic experiments
in a two-compartment electrochemical flow-cell reactor using an air-fed GDE.
Figure 7.4 plots this hydrogen peroxide efficiency as a function of the current
density. The relationship is very well represented by the function ηHO− =
2
0.138 log (Iapp /Ageo )−0.190. The relationship between the production efficiency
and the current density is most probably the result of a combination of factors.
Such factors might be a higher Faradaic current efficiency and a higher degree
of migration of the HO− 2 (in insufficiently well-supported electrolytes), at higher
current densities, therefore increasing the outward flux of these species. The
latter might have an additional effect because a higher outward flux might
decrease the amount of subsequent reduction of HO− −
2 to OH in the GDE.

4.2 Electrolyte penetration depth

Figure 7.5 shows the penetration depth of electrolyte by showing the position of
CuSO4 salt crystals in the active carbon/PTFE layer of the GDE. The occurrence
RESULTS AND DISCUSSION 127

porous gas-diffusion layer

active carbon layer

50 m

solid CuSO4 crystals

200 m

Figure 7.5: SEM photograph of the cross-section of a GDE pre-treated by


penetration with CuSO4 electrolyte. Inset: clear CuSO4 crystallites (bright
white) on the interface (black dots) of the porous gas-diffusion layer (light
phase) and the active carbon layer (darker phase). These CuSO4 crystals are
distributed all over the active carbon layer, even at the interface with the porous
gas-diffusion layer, indicating the complete penetration of the GDE with CuSO4
electrolyte.

of CuSO4 salt crystals, at the interface of the porous PTFE gas-diffusion layer
and the active carbon/PTFE layer, indicate the complete penetration of the
active layer by the electrolyte. This estimates the penetration depth as the
total depth of the active layer (i.e., on average approximately 300 µm).
128 ACCUMULATION OF ORR REACTION PRODUCTS IN POROUS GAS-DIFFUSION ELECTRODES

9
8

I3- mass-tr. coeff. kI3- [10-5 m s-1]


7
6
5
4
3
2
1
0
0 100 200 300 400 500 600
flow rate q [mL min-1]

Figure 7.6: Measured mass-transfer coefficient of I−


3 and derived mass-transfer
coefficients of OH− and HO−
2 as a function of the flow rate.

4.3 Mass-transfer coefficient

Figure 7.6 demonstrates the measured mass-transfer coefficient for I− 3 . From


these data, the mass-transfer coefficients of OH− and HO− 2 as a function of
the flow rate were determined. Mass-transfer coefficients in the order of 10−5
m s−1 are commonly observed. Due to the Grotthuss mechanism [193], a
mechanism that allows the fast transport of protons in aqueous media, the
diffusion coefficient of H+ in water is unusually high. As a result of the chemical
equilibria, the diffusion coefficients of OH− and HO−2 are also higher compared to
most other species in aqueous solutions. Therefore, the mass-transfer coefficients
for OH− and HO− −
2 are also expected to be higher than for I3 , according to
equation (7.38). The values of the mass-transfer coefficient will be used further
below to calculate the Sherwood-like numbers of the electrochemical reactor.

4.4 Accumulation of OH− and HO−


2 in the porous GDE

Figure 7.7 shows the observed amount of protons in the bulk nbH+ as a function
of time. nbH+ [mol] is calculated from the measured pH values as:

nbH+ (t) = 10−pH(t) Vbulk (7.39)


RESULTS AND DISCUSSION 129

0.4
500 mA
A m-2
slope = rpar
350
350 mA
A m-2
A m-2
200 mA

[10-3 mol]
A m-2
100 mA
0.3
5050mA
A m-2
protons in the bulk

0.2

0.1

slope determination

0
0 500 1000
time t [s]

Figure 7.7: The observed amount of protons in the bulk electrolyte nbH+ as a
function of time for different values of the applied current density Iapp /Ageo .

with Vbulk [L] the volume of bulk electrolyte (0.05 L in the experiments). From
the assumptions, this would be equal to the total amount of species accumulated
in the GDE nGDEtot,ss [mol]:

nGDE nGDE
P
tot,ss (t) = i,ss (t) (7.40)
i

with nGDE
i,ss [mol] the accumulated amount of species i. The curves in figure
7.7 clearly show that a constant value is not reached within the observed time-
frame. Additional experiments were carried out to confirm that a constant
value steady-state was never reached, even for experiments up to 2000 seconds
(appendix F). However, the slope of the curve remained constant for times
higher than 800 seconds. Such constant sloped steady-state would be in line
with a consistent cathode H+ removal efficiency lower than the assumed 100%.
This possibly indicates a parasitic reaction in the GDE consuming electrons not
resulting in the removal of a proton.
To confirm the existence of a parasitic reaction, a linear scan voltammetry at
1 mV s−1 in the 2 M KNO3 stock solution was carried out. Figure 7.8 shows
the voltammogram, which demonstrates the occurrence of a parasitic reaction
resulting in an observable current before reaching the potential at which the
ORR starts. This parasitic current depends more or less linearly on the applied
130 ACCUMULATION OF ORR REACTION PRODUCTS IN POROUS GAS-DIFFUSION ELECTRODES

0
𝐼𝑝𝑎𝑟
= 81 𝐸 − 25.1
𝐴𝑔𝑒𝑜
-200
500 A m-2

current density i [A m-2]


350 A m-2
-400 200 A m-2
100 A m-2
50 A m-2
-600

-800

-1000

-1200
-0.5 -0.4 -0.3 -0.2 -0.1 0 0.1 0.2
GDE potential E vs. SHE [V]

Figure 7.8: Linear scan voltammogram for a scan rate of 1 mV s−1 . The
different values for the applied current density Iapp /Ageo are indicated with
diamonds. The black line indicates the slope of the line for the current not
originating from the ORRs.

potential. The observed amount of protons in the bulk nbH+ is thus not equal to
the accumulation of species in the GDE, but a combination of the accumulation
of the species in the GDE and a linear steady-state part, caused by a parasitic
reaction:
Ipar
nbH+ (t) = nGDE GDE
ss,tot,exp (t) + rpar t = nss,tot,exp (t) + t (7.41)
F
with rpar = Ipar /F [mol s−1 ] a constant rate of the parasitic reaction and Ipar
[A] the parasitic current (figure 7.8). The accumulation of species in the GDE
nGDE
tot,ss,exp can then supposedly be obtained by subtracting the linear part from
the total observed amount of protons in the bulk nbH+ (figure 7.10). The data
obtained by this treatment now represents the total accumulation of species in
the GDE, equal to the sum of the accumulation of OH− and the accumulation
of HO− 2 ions:

nGDE
ss,tot,exp (t) = nGDE
ss,OH− ,exp
(t) + nGDE
ss,HO−
(t)
2 ,exp
1

GDE GDE
 (7.42)
= α n ss,OH− ,mod
(t) + n ss,HO− ,mod
(t)
2

with nGDE [mol] the experimental accumulation of species i, nGDE


ss,i,exp ss,i,mod [mol]
the uncorrected modeled accumulation of species i, and α [-] a correction factor
RESULTS AND DISCUSSION 131

to handle the miss-match between the model and experiment which occurs as a
result of additional phenomena that are not accounted for in the model (e.g.,
migration of species, etc.). For values of α = 1, the modeled and experimental
accumulation would fit perfectly and the model would take into account all
occurring phenomena. Other values would indicate the occurrence of phenomena
not taken into account.
Both accumulation terms depend heavily on the effective diffusion coefficient of
the respective species (equation (7.32)). First, a parameter δD is defined as the
ratio of the two effective diffusion coefficients:
DOH−
δD = (7.43)
DHO−
2

Now, the model can be fitted using 3 fitting parameters. The first parameter
is the effective diffusion coefficient of OH− , DOH− , the second parameter is
the ratio of the effective diffusion coefficients of OH− and HO− 2 , δD , and the
third fitting parameter is the correction factor α. The procedure fits α and
DOH− in the model to experimental accumulation data for the 5 previously
discussed current densities. This is done for different values of δD (i.e., from 1
to 10). For each value of δD , the root mean square (RMS) of the residuals of
each current density experiment is calculated and summed up. These sums are
plotted as a function of δD to find the most appropriate value for δD . Figure
7.9 demonstrates these sums as a function of δD , demonstrating the best fit
for a value of δD = 2.75. The value for the ratio of the diffusion coefficients
found in the literature [191, 192] is equal to 2.67 (DOH− /DHO− = 4.56 10−9 m2
2
s−1 /1.71 10−9 m2 s−1 = 2.67). Therefore, this value for δD could be considered
reasonable. Figure 7.10 shows the treated experimental data from figure 7.7 as
the total accumulation in the GDE and compares it to the fitted model for a
value of δD = 2.75.
The fitted values of α and DOH− are plotted for the different current densities in
figure 7.11. The values of the effective diffusion coefficients (figure 7.11a) show
a slightly upwards trend with current density, but given the many assumptions,
it is probably safer to estimate the effective diffusion coefficient of OH− with
the average of these values. As such, the average effective diffusion coefficient of
OH− , DOH− , is equal to (2.73 ± 0.38) 10−10 m2 s−1 . Using δD and DOH− , it is
possible to estimate the average effective diffusion coefficient of HO−
2 to be equal
to (9.94 ± 0.14) 10−11 m2 s−1 . Using equation (7.29) and the experimentally
obtained values of kiq , wGDE , ψ and Di , the Sherwood-like numbers of OH−
and HO− 2 in the flow-cell at a flow rate q of 500 mL min
−1
can be estimated to
be ShOH− = 2902 and ShHO− = 2989, both significantly higher than 2 and 3,
2
demonstrating convection in the flow-cell reactor has very little influence on
the accumulation of species in the GDE.
132 ACCUMULATION OF ORR REACTION PRODUCTS IN POROUS GAS-DIFFUSION ELECTRODES

1.6

1.4

sum of RMS of residuals [-]


1.2

0.8

0.6
minimum at 2.75
0.4

0.2

0
0 1 2 3 4 5 6 7 8 9 10
D[-]

Figure 7.9: Sum of the 5 root mean squared (RMS) residuals of the fit
of 5 experimental and modeled accumulation curves as a function of δD =
DOH− /DHO− , the ratio of the effective diffusion coefficient of OH− to the
2
effective diffusion coefficient of HO−
2 . The lowest residuals are found for a ratio
of δD = 2.75.

The values of α show a significant dependence on the current density and are not
equal to 1 (figure 7.11b). This indicates the occurrence of phenomena not taken
into account by the model. It is clear that at higher currents, the corrected
accumulation (α = 0.3) is significantly smaller than 1. One very suitable
explanation is the fact that migration is not taken into account, something
increasingly important at higher current densities. This is especially true if the
concentration of the ions in the supporting electrolyte would have similar values
as the concentration of OH− or HO− 2 . To validate this, it is possible to calculate
the concentration profiles of OH− and HO− 2 in the GDE with equation (7.26)
using the values obtained from the experiments (table 7.1). This is shown in
Figure 7.12. In this case, the values of α were used as a correction factor to
the production term of equation (7.26). As such, the concentration profile is
corrected for the observed deviations for the accumulation in the GDE. From
figure 7.12 it is clear that for applied current density values between 100 A m−2
and 200 A m−2 , the maximum concentrations of OH− in the GDE surpasses
the 2 M mark, which is the concentration of the supporting electrolyte. In such
a case, the assumption that migration of these species is negligible is incorrect.
RESULTS AND DISCUSSION 133

0.25
500 mA
500 A m-2

total accumulation GDE nGDE,tot [10-3 mol]


350 mA
350 A m-2
200 mA
200 A m-2
0.2 100 mA
100 A m-2
5050mA
A m-2
modelfitfit
model
0.15

0.1

0.05

0
0 200 400 600 800 1000
time t [s]

Figure 7.10: The experimental and modeled time-dependent total accumulation


of OH− and HO− 2 in the GDE as a function of time, for different values of the
applied current density Iapp /Ageo .

10 0.7
a b
eff. diffusion coeff. 𝒟OH-[10-10 m2 s-1]

0.6
correction factor  [-]

0.5

average of 2.74 0.4

0.3

1 0.2
0 200 400 600 0 200 400 600
𝐼𝑎𝑝𝑝 𝐼𝑎𝑝𝑝
current density [A m-2] current density [A m-2]
𝐴𝑔𝑒𝑜 𝐴𝑔𝑒𝑜

Figure 7.11: Values of the effective diffusion coefficient of OH− (a) and correction
factor α (b) as a function of the applied current density Iapp /Ageo
134 ACCUMULATION OF ORR REACTION PRODUCTS IN POROUS GAS-DIFFUSION ELECTRODES

Therefore, it is not surprising that the value of α differs more from 1 if the
applied current density increases.
The accumulation of OH− above concentrations of 1 M results in a noticeable
decrease in the apparent OH− production efficiency. This translates itself
in observable delays in the time-dependent pH increase during galvanostatic
experiments. When GDEx is carried out in batch mode, this can result in
a significant increase of the time to remove all metal ions (as precipitates
or nanoparticles). The additional time necessary results in a higher power
consumption per amount of removed metal ion or per amount of produced
nanoparticle, which is undesirable. However, because the accumulation reaches
a steady state when GDEx would be carried out in continuous mode, the
accumulation could be considered part of the so-called start-up time, which
from our experiments appeared to be around 800 seconds. As stated before, to
become an industrially relevant process, GDEx must be carried out in continuous
mode, therefore the initial accumulation causes little harm to the overall process
efficiency. In addition to the accumulation, certain parasitic electrochemical
reactions also decrease the production efficiency of OH− . In contrast to the
initial accumulation, such reactions continuously weigh on the process efficiency.
Therefore, to optimize the process, special attention must be paid to avoid such
parasitic reactions as much as possible.
Although the initial accumulation initially decreases the apparent production
efficiency, it also results in very high OH− concentrations in the GDE (figure
7.12). The resulting high pH causes a very stable work potential of the GDE
(figure 7.13) for a very large range of current densities. This is caused by the
fact that the reduction potential of O2 strongly depends on the pH. Because the
pH is high, it remains stable over time. A stable potential is beneficial, because
it needs little control and results in less electric losses.
RESULTS AND DISCUSSION 135

0 x→ wGDE
5
OH-
4.5 HO2-

concentration of species [mol L-1]


4 500 A m-2
350 A m-2
3.5
200 A m-2
3 100 A m-2
50 A m-2
2.5

1.5

0.5

0
0 100 200 300
position in active layer of GDE [m]

Figure 7.12: Concentration profiles of OH− and HO−2 in the active layer of the
GDE for different values of the applied current density Iapp /Ageo , calculated
from the experimentally obtained parameters and the values of the fitted model
parameters (see table 7.1).

0.2
500 A m-2
Series3
350 A m-2
Series4
Series5
200 A m-2
0.1
GDE potential E vs. SHE [V]

Series6
100 A m-2
Series7
50 A m-2

-0.1

-0.2

-0.3
0 500 1000
time t [s]

Figure 7.13: GDE potential as a function of time for different values of the
applied current density Iapp /Ageo .
136 ACCUMULATION OF ORR REACTION PRODUCTS IN POROUS GAS-DIFFUSION ELECTRODES

5 Conclusions

In this work, the accumulation of OH− and HO− 2 in a porous air-fed gas-diffusion
electrode was investigated. An analytical steady-state model was derived for the
concentration profiles in the GDE and an analytical time-dependent model was
derived for the total accumulation of OH− and HO− 2 in the GDE. Experimental
data was gathered for the time-dependent accumulation for different current
densities. These data suggested the existence of parasitic reactions besides
the desired oxygen reduction reactions (ORRs). This existence was verified by
linear scan voltammetry. The experimental data could be fitted to the modeled
accumulation to obtain an estimate of the effective diffusion coefficients of OH−
and HO− 2 . The best fits were obtained for a ratio of the diffusion coefficients of
OH− and HO− 2 equal to 2.75, close to what values found in literature. Good
fits were only obtained when a correction factor was used, which indicated the
limitation of the analytical model to incorporate all experimental phenomena.
One such phenomenon is the fact that migration was not taken into account.
Since the model predicts concentrations of OH− and HO− 2 equal to or higher
than the concentration of the ions of the supporting electrolyte, this assumption
is invalid at high current densities. Overall, this work demonstrates the existence
of accumulating species and quantifying this accumulation in porous air-fed gas-
diffusion electrodes. It was possible to estimate the effective diffusion coefficients
of these species in relatively simple experiments.
Chapter 8

General conclusions

The general aim of this dissertation was to discover the fundamental aspects
of gas-diffusion electrocrystallization (GDEx). To achieve this purpose (1) an
extensive exploratory study of the GDEx process was carried out to demonstrate
similarities and differences with alkaline titration, a well-known method for
nanoparticle synthesis in addition to several other more in-depth investigations
to further develop the understanding of the physicoelectrochemical processes
in the GDEx process. These investigations were (2) the mathematical and
experimental analysis of alkaline titration of metal ions to reveal the effect of
different system parameters on the time-dependent behavior of metals during
alkaline precipitation, (3) the investigation of the effect of oxidants on the
precipitation behavior of oxidizable metal ions during alkaline precipitation,
discovering the phenomenon of oxidation-assisted alkaline precipitation, (4)
the demonstration of this newly discovered phenomenon in the confines of
the GDEx process by producing and comparing nanoparticles from different
metal-containing solutions and (5) the mathematical and experimental analysis
of the accumulation of electrochemically produced species in the pores of the
porous gas-diffusion electrode to increase understanding of how the processes in
the porous electrode affect the observations in the bulk electrolyte.
The motivation to investigate GDEx came from preliminary experiments at
VITO that demonstrated the potential of GDEx to fastly produce crystalline
metal oxide nanoparticles from metal-containing waste streams, even for streams
with low concentrations of metal ions. The combination of synthesis and metal
removal or recovery is desirable given the fact that society relies increasingly on
the development of new functional materials [90] and environmental concerns are
put higher on the political agendas by every new election [89]. In addition to this,

137
138 GENERAL CONCLUSIONS

the preliminary results demonstrated an interesting phenomenon, which was the


oxidation of metal ions in the catholyte of a two-compartment electrochemical
cell. This was considered an odd observation, given that the cathode is the
electrode at which reduction reactions occur, rather than oxidation reactions.
In light of this, the main aim was to understand GDEx with regards to how
metal ions are removed, how and why they precipitate as nanoparticles, and
how they are oxidized in a presumably reducing environment.
GDEx is an electrochemical synthesis method. Therefore, it is inherently
a process involving heterogeneous electrochemical reactions. Heterogeneous
processes are limited by the available reaction surface and are, at best, mass-
transfer limited. Because of this, the fast production of nanoparticles with
GDEx, even from solutions with low concentration, arose suspicion and questions
about the possible involvement of fast homogeneous reactions.
The exploratory study of GDEx was able to demonstrate the homogeneous
reactions that govern the GDEx process. These homogeneous reactions were
identified as the alkaline precipitation (AP) of metal (oxy)(hydr)oxides or metal
hydroxysalts in the bulk electrolyte. The AP of metal ions is caused by an
increase of the bulk electrolyte pH as a result of the production of OH− (i.e.,
removal of H+ ) in the porous gas-diffusion electrode (GDE). This OH− is
produced together with HO− 2 in oxygen reduction reactions (ORRs) that occur
at the surface of the GDE. Because of the production of OH− , the GDEx process
showed a lot of similarities with another more frequently used method for the
alkaline precipitation of metal ions: the alkaline titration (AT) with NaOH. AT
is a reasonably simple method and involves only the slow addition of NaOH to
an electrolyte in a closed batch reactor. The pH of the electrolyte during AT is
easily measured and therefore an ideal parameter for comparison of model and
experiment. The shape of the time-dependent pH curves can give information
on the processes that occur, and on the expected precipitates, among other
things. To our surprise, an in-depth investigation, involving a comparison of
modeled and experimental pH curves obtained during AT, was lacking in the
scientific literature. Therefore, because of the similarities of AT and GDEx,
such investigation was carried out in this work. The most important conclusion
of the comparison of model and experiment was that the AT of ZnCl2 solutions
is extremely well-suited to be modeled and that this system can, therefore,
act as a reference system against which other precipitation processes can be
compared. Deviation of the model from the experiments can carry interesting
information about the ongoing processes.
In contrast to alkaline titration, the exploratory study demonstrated that
GDEx produced significant amounts of H2 O2 . H2 O2 is a well-known oxidant
and its occurrence in the catholyte was a first indication of what caused the
GENERAL CONCLUSIONS 139

oxidation of metal ions in the catholyte during the preliminary research at


VITO. It was expected that this species might have important consequences
for the precipitation behavior. In light of this, the effect of H2 O2 on
the precipitation behavior of nanoparticles was investigated in a dedicated
study. This investigation uncovered the existence of oxidation-assisted alkaline
precipitation (Ox-AP), a process in which an oxidant interferes in the alkaline
precipitation process of a metal ion, resulting in nanoparticles of smaller sizes
with more narrow size-distributions than would be obtained without the oxidant.
Given the insights on Ox-AP, it was expected that GDEx was capable of eliciting
Ox-AP under certain circumstances for certain metal-containing solutions.
Therefore, in a subsequent study AT and GDEx were compared for the ZnCl2
reference system, and two other solutions containing oxidizable metal ions:
MnCl2 , and FeCl2 . ZnCl2 demonstrated to be a good reference system for
comparison of both techniques because AT and GDEx both produced Zn(OH)2
nanoparticles with the same size, morphology, and composition. In addition,
H2 O2 production was unaffected in this case, indicating that H2 O2 does not
interfere with the precipitation process. Because of these observations, AT
and GDEx could be considered equivalent in the case of ZnCl2 , making it
indeed an ideal reference case. For MnCl2 , AT and GDEx resulted both in
Mn3 O4 nanoparticles with similar characteristics. In contrast to the ZnCl2
case, no production of H2 O2 could be observed, indicating a certain interaction
between Mn-containing species and the H2 O2 . Remarkably, however, partial
oxidation of Mn(II) to Mn(III) was observed for both AT and GDEx. Also,
the pH observations indicated a very similar precipitation behavior. Therefore,
rather than GDEx eliciting Ox-AP in the case of MnCl2 , it was concluded that
it is more likely that H2 O2 is catalytically decomposed on the precipitated
manganese species, and that instead, the oxidation likely happens during post-
treatment of the precipitate. For FeCl2 , AT and GDE synthesized very different
nanoparticles. With AT, tile-shaped FeO nanoparticles were synthesized with a
size up to ten times larger than the Fe3 O4 nanoparticles synthesized with GDEx.
Like the case of MnCl2 , no production of H2 O2 was observed. However, in this
case, the pH observations clearly indicated a strong deviation of the precipitation
behavior. Using these insights, it was concluded that GDEx elicited Ox-AP of
the Fe2+ , which resulted in the synthesis of much smaller nanoparticles.
In addition to demonstrating the potential of GDEx to elicit Ox-AP under
certain conditions, the comparison of AT and GDEx also exposed the fact that
the production efficiency of OH− was lower than expected. This apparently-low
production efficiency was demonstrated to be caused by the accumulation of
OH− and HO− 2 in the pores of the porous GDE. It was demonstrated that
higher current densities caused more accumulation. Although it was expected
that a constant accumulation would be reached in the steady-state, instead
140 GENERAL CONCLUSIONS

the steady-state was characterized by a linear increase of the accumulation.


Evidence pointed towards parasitic reduction reactions in the GDE accounting
for over 10% of the applied current at low current densities.
In general, the aim of this thesis has been achieved. A first contribution towards
the fundamental understanding of the nanoparticle formation during GDEx
was made. GDEx was characterized as a nanoparticle synthesis method that,
under certain conditions, can elicit the newly discovered process of oxidation-
assisted alkaline precipitation (Ox-AP). Ox-AP was linked to the reaction of
the metal ion in solution with the OH− and H2 O2 produced in the oxygen
reduction reactions carried out at the GDE. And the produced H2 O2 was
connected to the odd occurrence of metal oxidation reactions in the catholyte of
an electrochemical reactor. Moreover, GDEx-elicited Ox-AP showed to result in
nanoparticles with smaller sizes than those achievable with alkaline precipitation.
This fact combined with the fact that the GDEx process can run entirely on
renewable energy, makes GDEx a viable future alternative for the production
of functional materials, possibly coupled to the removal or recovery of metal
ions from waste streams.

Outlook and open questions

On the one hand, an essential part of the GDEx process has been demystified.
Fortunately, or unfortunately, depending on the scientist’s frame of mind, many
open questions and ideas remain, as well as a few unfinished projects that
deserve some attention in this concluding section.
Although all of the work, related to GDEx, presented in this dissertation was
carried out in a flow-cell reactor operating in recirculation mode, GDEx would
be more practical as a continuous process. The first steps toward this were made
together with my close colleague Rafael Prato, who demonstrated the usefulness
of the GDEx process to controllably synthesize iron oxide nanoparticles with
different characteristics [169]. This work of him focused on the effect of the
initial iron concentration. However, besides the initial metal ion concentration,
the background electrolyte, the flow rate, and the current-density are expected to
also have important effects on the characteristics of the synthesized nanoparticles.
Especially because of their effect on the supersaturation of the precipitating
species. Additionally, the flow rate and current-density will have a significant
effect on the relative amount of H2 O2 available in solution, the species responsible
for eliciting Ox-AP in the GDEx process and thus the root cause for the
formation of nanoparticles with smaller sizes than obtainable with traditional
alkaline precipitation.
GENERAL CONCLUSIONS 141

In light of this, there is unfinished work in carrying out GDEx in a one-


compartment electrochemical flow-cell reactor. Although GDEx is capable
of producing oxidants in the catholyte, oxidants are generally produced at
the anode. Using only one compartment, oxidants can be produced by both
electrodes in a single compartment. One can estimate that the steady-state redox
potential of a solution in a one-compartment electrochemical flow-cell reactor
will be somewhere in between the redox potentials of the reactions carried out at
the anode and the cathode. Here, GDEx opens interesting possibilities. The use
of oxygen as a reagent introduces a relatively high cathode potential. Because
of this, the redox potential of the solution is pushed upwards significantly
compared to when protons are the reagent. Therefore, the one-compartment
GDEx process can be used to precipitate metal (oxy)(hydr)oxide nanoparticles
of metal ions which are not oxidized in the two-compartment GDEx set-up. Such
an example is the case of nickel. Only in the one-compartment flow-cell reactor,
using the oxygen reduction reaction at the cathode and the chlorine evolution
at the anode, the characteristic black NiOOH precipitate could be obtained.
In addition, first observations were made of a purple manganese solution after
treating an MnCl2 solution in the one-compartment GDEx reactor, indicating
the occurrence of the permanganate ion, a known higher oxide of the manganese.
This interesting investigation pathway is open for future work.
The GDEx process in this work was narrowly defined. Only aqueous electrolytes
were investigated, the GDE was only used as a cathode, and the feed gas was
always air. Many other electrolytes (e.g., organic-based liquids) and gasses are
available for reduction (e.g., O2 , CO2 , etc.).
Other unfinished work relates to a thorough essay on the sustainability
opportunities with GDEx. Small steps towards demonstrating the economic and
ecological benefits of GDEx were taken, especially with regards to demonstrating
a decrease in power consumption from using the oxygen reduction reaction, but
a more systematic approach urges itself.
Finally, one of the ultimate goals in understanding GDEx is to have an all-
encompassing model that can predict the behavior of a certain stream of metal
ions when treated by GDEx. This model would allow one to elaborately produce
nanoparticles with desired sizes, shapes, physical properties, among others,
so-called synthesis-by-design. The first steps were taken to model the time-
dependent pH in the catholyte flow-cell and reservoir. The observed issues with
the accumulation of OH− in the GDE put a halt on further development of this
model because in contrast to the good agreement between AT experiments and
models, the first GDEx pH model was unable to take into account the observed
delays. Without knowing the origin of the delays it was assumed best to pause
the development of the model. Our current understanding of the accumulation
allows us to incorporate a correction term for the accumulation in the electrode,
142 GENERAL CONCLUSIONS

which would help to improve the agreement between model and experiment. It
would also provide the possibility to model the continuous process and provide
the first steps towards choosing the right process parameters in function of,
among other things, the desired supersaturation, and H2 O2 concentration in
the reactor.
Besides these unfinished projects closest to my own interests, many more
applications can be imagined. The potential for GDEx as a new green
electrochemical process is high, both from a removal and a synthesis perspective,
with many opportunities for GDEx as a sustainable alternative to existing
processes. It is my hope and conviction that some of the ideas and concepts
that were worked out in this PhD are useful in future work regarding the
GDEx process specifically, but the field of (green) chemical and electrochemical
synthesis more generally.
Appendix A

Measuring the copper


concentration with a Cu-ISE

The Cu2+ concentration in a CuSO4 solution can be measured using a Cu ion-


selective electrode (Cu-ISE). The potential of the Cu-ISE is measured against
a Hg/Hg2 SO4 reference electrode. Conversion to concentrations is done by
making a calibration with known Cu2+ concentrations.
A 25 mL solution of 0.005 M CuSO4 with 0.25 M Na2 SO4 was used. A Cu-
ISE and a Hg/Hg2 SO4 reference electrode (Schott instruments B 3610+) were
immersed in this solution. The solution was diluted every 30 seconds with
increments of 1 mL 0.25 M Na2 SO4 solution. The potential of the Cu-ISE was
measured against the Hg/Hg2 SO4 reference electrode (Methrohm 691 pH/ion
meter). After each increment, the potential was registered. These potentials
were plotted against the logarithm of the known concentrations of Cu2+ after
each increment (figure A.1).
During the AT and GDEx experiments with 0.005 M CuSO4 , the Cu-ISE and
the reference electrode were introduced into the solution and the potential was
registered every 5 seconds. This potential was  then  recalculated to the Cu2+
2+
 
concentration with the calibration curve log Cu cal = 29.4EISE,cal + 3.3
using the following formula:

 2+  0.005
Cu (t) = 1029.4EISE (t)+3.3
29.4EISE (0)+3.3
(A.1)
10
with the multiplication factor 0.005/1029.4EISE (0)+3.3 a correction using the
firstly measured term to always have the correct initial concentration (necessary

143
144 MEASURING THE COPPER CONCENTRATION WITH A CU-ISE

-2.1
log [Cu2+] = 29.4 ECu-ISE +3.3
-2.3

-2.5

log [Cu2+]
-2.7

-2.9

-3.1

-3.3

-3.5
-0.23 -0.22 -0.21 -0.2 -0.19 -0.18
Cu-ISE potential vs. Hg/HgSO4 EISE [V]

Figure A.1: Calibration curve of the Cu-ISE, plotting the logarithm of the
known concentration of Cu2+ against the measured potential of the Cu-ISE.

due to measurement errors).


The measured potential and how it translates to the Cu2+ concentration during
a GDEx experiment is shown in figure A.2.

-0.15 6
Cu concentration [Cu2+] [10-3 mol L-1]

-0.2 5
Cu-ISE potential EISE [V]

-0.25 4

-0.3 3

-0.35 2

-0.4 1

-0.45 0
0 500 1000 1500
time t [s]

Figure A.2: Calibration curve of the Cu-ISE, plotting the logarithm of the
known concentration of Cu2+ against the measured potential of the Cu-ISE.
Appendix B

PHREEQC input file for


modeling PTCs of Zn(OH)2

#######################################################################################
##### This PHREEQC (.pqi) input file is developed by Sam Eggermont (2019). #####
##### It allows the calculation of the titration of a ZnCl2 solution with #####
##### a NaOH solution in a fixed amount of titration steps. It uses the #####
##### pitzer.dat database. In this input file the unavailable but necessary #####
##### additional phases and species are added from the Minteq.v4.dat database #####
##### This version is the adaptation for 100 mL 5 mM ZnCl with 0.5 M NaCl, #####
##### titrated with 20 mL of 0.1 M NaOH solution in 1222 steps. #####
##### #####
##### This file is easily adopted to other solutions allowing for different #####
##### metal ions, different anions, different salts, different titrant, etc. #####
#######################################################################################

#######################################################################################
##### Part 1: adding species that are not available in the Pitzer database #####
#######################################################################################

SOLUTION_MASTER_SPECIES ### Adding the metal ion

Zn Zn+2 0.0 Zn 65.39 ### Addition of a Zn master species to Pitzer


### database.
T T 0 T 0.00001 ### Artificial master species to act as
### titration step counter

SOLUTION_SPECIES ### Adding the thermodynamic data of the species


###(from minteq.v4 database) which are missing
### in Pitzer.dat
T = T
log_k = 0

Zn+2 = Zn+2
log_k 0

Zn+2 + H2O = ZnOH+ + H+ ### This block is excluded for NM1 and included for NM2
log_k -8.997

145
146 PHREEQC INPUT FILE FOR MODELING PTCS OF ZN(OH)2

delta_h 55.81 kJ
-gamma 0 0

Zn+2 + 2H2O = Zn(OH)2 + 2H+ ### This block is excluded for NM1 and included for NM2
log_k -17.794
delta_h 0 kJ
-gamma 0 0

Zn+2 + 3H2O = Zn(OH)3- + 3H+ ### This block is excluded for NM1 and included for NM2
log_k -28.091
delta_h 0 kJ
-gamma 0 0

Zn+2 + 4H2O = Zn(OH)4-2 + 4H+ ###This block is excluded for NM1 and included for NM2
log_k -40.488
delta_h 0 kJ
-gamma 0 0

Zn+2 + Cl- = ZnCl+ ###This block is excluded for NM1 and included for NM2
log_k 0.4
delta_h 5.4 kJ
-gamma 4 0

Zn+2 + 2Cl- = ZnCl2 ###This block is excluded for NM1 and included for NM2
log_k 0.6
delta_h 37 kJ
-gamma 0 0

Zn+2 + 3Cl- = ZnCl3- ###This block is excluded for NM1 and included for NM2
log_k 0.5
delta_h 39.999 kJ
-gamma 4 0

Zn+2 + 4Cl- = ZnCl4-2 ###This block is excluded for NM1 and included for NM2
log_k 0.199
delta_h 45.8566 kJ
-gamma 5 0

Zn+2 + H2O + Cl- = ZnOHCl + H+ ###This block is excluded for NM1 and included for NM2
log_k -7.48
delta_h 0 kJ
-gamma 0 0

PHASES ### Adding the thermodynamic data of the phases

fix_pH ### artificial phase to fix the pH to a certain value


H+ = H+
log_k 0.0
### Adding the thermodynamic data of the phases
### (from minteq.v4 database) which are missing in
### Pitzer.dat
ZnCl2
ZnCl2 = Zn+2 + 2Cl-
log_k 7.05
delta_h -72.5 kJ

ZincHydroxide
Zn(OH)2 + 2H+ = Zn+2 + 2H2O
log_k 12 ### 12 is the used value in the model for this publication,
delta_h -89.62 kJ ### the original value was 11.334 in Minteq.v4.dat database

PITZER ### Adding the Pitzer coefficients of Zn+2 which are missing
### in the pitzer.dat database
PHREEQC INPUT FILE FOR MODELING PTCS OF ZN(OH)2 147

-B0
Zn+2 Cl- 0.2602 ### ref. 1: Simoes et al. Estimation of the Pitzer
Parameters
-B1 ### ... for 1--1, 2--1, 3--1, 4--1, and 2--2 Single
### Electrolytes at 25* C
Zn+2 Cl- 1.6425 ### ref. 1
-C0
Zn+2 Cl- -0.088 ### ref. 1
END

###########################################################################################
##### Part 2: define initial solution - water with ZnCl2 and with or without NaCl #####
###########################################################################################

SOLUTION 1 ### Initiate solution: 100 mL of H2O


temp 25
pH 7
redox pe
units mol/l
density 1
water 0.1 L

USE SOLUTION 1
EQUILIBRIUM_PHASES ### Phases in equilibrium with water, to remove
### a phase delete the program line or add "#"
### at line start
Halite 0 0.05 ### 0.5 M NaCl in 100 mL
Melanothallite 0 0.0005 ### 0.005 M ZnCl2 in 100 mL
fix_pH -3 HCl ### Virtual phase to change initial pH to 3 with HCl
SAVE SOLUTION 1
END

##########################################################################################
##### Part 3: modeling stepwise addition of 20 mL of 0.1 M NaOH #####
##### solution, allow precipitation of Zn(OH)2 #####
##########################################################################################

USE SOLUTION 1
REACTION 2 ### Adding a total of 20 mL of 0.1 M NaOH via small
Na 0.1 ### steps
OH 0.1
H2O 55.55
T 0.1 ### Artificial counter to register the addition steps
0.02 moles in 1200 steps ### Steps are arbitrary, here they mimic 0.2 mL/min

EQUILIBRIUM_PHASES ### Defining phases that are in equilibrium w/ solution.


### "p" indicates precipitation for saturation index
Tenorite 0 0 p: ### above 1

USER_GRAPH 1 ### Generating the graph in-situ during calculations,


### demonstrating the pH vs. mol NaOH added
-headings 10_cells: pH
-axis_titles "NaOH added [mol]" "pH"
-axis_scale y_axis 0 14 1 1
-start
10 t =(sys("T"))
20 pH = -LA("H+")
30 GRAPH_X t
40 GRAPH_Y pH
Appendix C

Sample treatment for and


characterization with sp
ICP-MS

The samples obtained after the post-treatment explained in chapter 5 were first
100-fold diluted with 18MΩ ultra-pure water from a Milli-Q system (Millipore,
USA). Before analysis, the nanoparticle (NP) suspensions were sonicated using
a sonicator, ultrasonic processor (Qsonica LLC, USA) equipped with a microtip
1/16” working in pulse mode and with a maximum amplitude of 40% during 10
minutes to avoid NP agglomeration. After sonication, different sample dilutions
were prepared using ultra-pure water, followed by a gently stirred.
All measurements were carried out using a high-resolution sector-field ICP-MS
(HR-SF-ICP-MS) instrument (Nu AttoM ES, Nu Instruments, UK) equipped
with a conventional sample introduction system comprising a glass concentric
nebulizer mounted onto a baffled quartz cyclonic spray chamber. The instrument
was operated in single-particle (SP) mode with a dwell time of 50 µs. To
overcome spectral overlap, pseudo-medium resolution mode was selected for the
monitoring of 56 Fe, while a low resolution was used in the case of 65 Cu (see Table
S2). To calibrate the transport efficiency (TE), gold nanoparticles (AuNPs),
stabilized in a citrate buffer, of known diameter and ionic standards solutions
of Au were used. Additionally, the analyte sensitivity was also measured by
using Cu and Fe ionic standard solutions. Appropriate dilutions factors for each
of the samples were selected to obtain good statistics without compromising
the occurrence of double events (Poisson statistics). Data treatment was

149
150 SAMPLE TREATMENT FOR AND CHARACTERIZATION WITH SP ICP-MS

performed using the Nu Quant software (Nu Instruments, UK). This software
can automatically differentiate the signal spikes corresponding to individual
NPs and the continuous background. In this way, particle sizes can be obtained
using the calculation’s editor built into the software and based on different
algorithms [194]. Additional information required for data treatment, such as
shape, chemical composition and density of the particles, is shown in table C.1.
In all cases, a spherical shape was assumed, and thus, the particle spherical
equivalent diameter is reported.

Table C.1: Calculations of the supersaturation values at pH 12


Sample Particle Elemental Molecular Density Isotope Resolution Dilution
composition mass mass [g cm3 ] [108] measured mode factor
[g mol−1 ] [g mol−1 ]
65
CuCl2 CuO 63.55 79.454 6.31 Cu Low 2 106
65
CuCl + CuO 63.55 79.454 6.31 Cu Low 2 107
H2 O 2
56
FeCl3 FeOOH 55.85 88.852 4.26 Fe Pseudo- 2 108
medium
56
FeCl2 + FeOOH 55.85 88.852 4.26 Fe Pseudo- 2 108
H2 O 2 medium
Appendix D

Calculating supersaturations

At the Y-junction the metal chloride solution and the NaOH solution are mixed.
Just before the precipitation reaction takes place, the metal ion concentration
and the hydroxyl ion concentration can be calculated for the total effluent
volume. Using the solubility products of the assumed precursors (Cu(OH)2 and
Fe(OH)3 ), a theoretical value for the supersaturation can be calculated. The
metal ion concentration is equal to the metal ion concentration of the influent
divided by 2. The hydroxyl ion concentration is calculated similarly in the
case of CuCl2 and FeCl3 but subtracting the amount of NaOH to neutralize
the acidic pH and overcome the water equilibrium. For the cases of CuCl and
FeCl2 , where H2 O2 is added, the theoretical calculation should also consider
the hydroxyl ions produced from the oxidation reaction between the metal ion
and hydrogen peroxide (maximally 1 time the metal concentration; because
each metal ion is assumed to oxidize to only one unit of valence higher). For
CuCl2 , CuCl, FeCl3 and FeCl2 the metal ion concentration is 0.005 M and the
OH- concentrations are respectively:

ˆ For CuCl2 solution: (0.14 M – 0.1 M)/2 = 0.02 M


ˆ For CuCl solution: (0.13 M – 0.1 M + 0.01 M)/2 = 0.02 M
ˆ For FeCl3 solution: (0.15 M – 0.1 M)/2 = 0.025 M
ˆ For FeCl2 solution: (0.14 M – 0.1 M + 0.01 M)/2 = 0.025 M

As observable, the OH− concentration for the copper cases is equal as well as the
OH− concentrations for the iron cases. The end OH− concentration in the iron
cases is 0.005 M more than the copper case. This is expected because the same

151
152 CALCULATING SUPERSATURATIONS

end pH is desired but for each Fe3+ ion, 1 more OH− -ion needs to react with it
compared to the Cu2+ -ion, and the metal concentration after
n mixing is 0.005
M. With the formula for supersaturation S = Mn+ OH− /Ksp,M(OH)n , the
 

theoretical supersaturation values can be calculated as shown in table D.1.

Table D.1: Calculations of the supersaturation values at pH 12


Case Precursor Ksp Theoretical value for initial S(upersaturation)
precipitate
2 2
[Cu2+ ][OH− ] 5 10−3 (2 10−2 )
CuCl2 and CuCl2 2 10−20 [107] S= Ksp,Cu(OH)2 = 2 10−20 = 9.1 1013
CuCl or
2 3 2 3
[Cu2+ ] [OH− ] [Cl− ] (5 10−3 ) (210−2 ) 6.05 10−1
Cu2 OH3 Cl 2.34 10−35 [153] S= Ksp,Cu2 (OH)3 Cl = 2 10−20 = 5.17 1024
3 3
−39 [Fe3+ ][OH− ] 5 10−3 (2.5 10−2 ) 31
FeCl3 and Fe(OH)3 2.79 10 [107] S= Ksp,Fe(OH)3 = 2.79 10−39 = 2.8 10
FeCl2
Appendix E

Derivation of the analytical


time-dependent model of the
total accumulation in the
GDE

The time-dependent equation without homogeneous reactions, taken far away


from the H2 O and H2 O2 equilibria (no reaction terms) can be written as:

∂ci ∂ 2 ci Pi
− Di 2 = (E.1)
∂t ∂x Vp

This equation resembles the heat equation with a constant source term and
is often solved with separation of variables. Although this approach gives the
exact solution, its form is needlessly complex and does not comply with our
vision to provide a straightforward solution. Because we are not especially
interested in the very short time range, but more in the overall behavior, and
more specifically at times approaching the steady state, we opted to solve the
equation using Laplace transforms. As such, while solving the equations in the
Laplace domain, it is possible to solve the equations for the limiting situation
when s approaches zero (equivalent to the time t going to infinity). The Laplace
transform of equation (E.1) is:
∂ 2 ĉi (x, s) Pi
sĉi (x, s) − ci (x, 0) − Di 2
= (E.2)
∂x Vp s

153
154 DERIVATION OF THE ANALYTICAL TIME-DEPENDENT MODEL OF THE TOTAL
ACCUMULATION IN THE GDE

with s the Laplace complex number frequency parameter, ĉi the Laplace
transformed concentration, and ci (x, 0) = ci,0 , a space independent initial
concentration equal to that of the initial bulk concentration ci,∞ (0). The
general solution to this differential equation is:
Pi + Vp ci,0 s √  √ 
ĉi (x, s) = + C1 cosh √ s x + C2 sinh √ s x (E.3)
Vp s2 Di Di

with C1 and C2 integration constants. The Laplace transformed boundary


conditions (equations (7.19) and (7.23)) are written as:

∂ĉi (0, s)
−Di =0 (E.4)
∂x
 
∂ĉi (wGDE , s) −ki ĉi,∞ (s)
= ĉi (wGDE , s) − (E.5)
∂x ψDi s
From the first boundary condition (equation (E.4)) we find C2 = 0:

∂ĉi (0, s) √  √  √  √ 
−Di = C1 √Ds sinh √s 0
Di
+ C2 √Ds cosh √s 0
Di
∂x i i
(E.6)
= C1 0 + C2 1 = C2 = 0

Using the second boundary condition (equation (E.5)), we can find the value of
C1 to be:
 
ki Pi +Vp ĉi (wGDE ,0)s−Vp ĉi,∞ (0)s
− ψD i V p s2
C1 = √ √  √  (E.7)
√ s sinh √ s wGDE + ki cosh √ s wGDE
D i D i ψDi D i

With the concentration at t = 0 the same at every position ĉi (wGDE , 0) =


ĉi,∞ (0), the expression for ĉi (x, s) becomes:
 √  
ki √s x
ci,0 Pi  ψDi cosh Di
ĉi (x, s) = + 1− √ √  √ 
s Vp s2 √ s sinh √ s w + ki cosh √ s w
Di Di GDE ψDi Di GDE
(E.8)
DERIVATION OF THE ANALYTICAL TIME-DEPENDENT MODEL OF THE TOTAL ACCUMULATION IN
THE GDE 155

We can integrate ĉi (x, s) over the volume of the pores Vp to find an expression
of the Laplace-transformed accumulation in the GDE n̂GDE i (s):
ci,0 Rw
n̂GDE
R
i (s) = ĉ
Vp i
dVp = Vp + ψAgeo 0 GDE ĉi (x, s) dx
s
ci,0 Pi
= Vp + ψAgeo wGDE ...
s Vp s2
√ √
 
ki 
√Di sinh √ s wGDE
 ψDi wGDE s Di 
1 −
...  √ √  √ 
√s sinh √Ds wGDE + ψDki
cosh √Ds wGDE

Di i i i

(E.9)
with Vp = ψAgeo wGDE , the expression for n̂GDE
i (s) becomes:
ci,0
n̂GDE
i (s) = Vp + ...
s
√ √
 
ki 
√Di sinh √ s wGDE
Pi  ψDi wGDE s Di 
... 1 − √ √  √ 
s2  √s sinh √Ds wGDE + ψDki
cosh √Ds wGDE

Di i i i

(E.10)
Since we are interested in the long time-frames, approaching steady-state, we
can simplify the expression for n̂GDE
i (s) by looking at the asymptotic behavior
and taking the limit for s going to 0 (something equivalent to taking the limit
for t to infinity) of the sinh and cosh terms:
√  √ 3/2
lim sinh √Ds wGDE = √Ds wGDE + s 3/2 wGDE 3
+ ... (E.11)
s→0 i i Di
 √  2
swGDE
lim cosh √ s wGDE =1+ + ... (E.12)
s→0 Di Di

As such, equation (E.10) simplifies to:


2
  
GDE ci,0 Pi wGDE ψ wGDE 1
n̂i (s) = Vp + + (E.13)
s s ki 3Di 1 + sτs

with τ equal to:


2
wGDE ψ wGDE
τs = + (E.14)
ki 2Di
156 DERIVATION OF THE ANALYTICAL TIME-DEPENDENT MODEL OF THE TOTAL
ACCUMULATION IN THE GDE

Inverse Laplace transform of equation (E.13) gives the accumulation as a function


of time:
nGDE
i (t) = ci,0 Vp +   
2
 
wGDE ψ wGDE  
1 − exp − t 
... Pi +  wGDE ψ w2

ki 3Di 
GDE

+
ki 2Di
(E.15)
A more concise expression would be:
  
 
1 1   t 
nGDE
i (t) = ci,0 Vp + Pi τi + 1 − exp −   
Shi 3   1 1 
τi +
Shi 2
(E.16)
with τi and Shi defined as:
2
wGDE
τi = (E.17)
Di
ki wGDE
Shi = (E.18)
ψDi
Appendix F

Long-time experiment of the


bulk proton concentration

In figure F.1 we demonstrate the long time experiments of the observed amount
of protons in the bulk. It is easily observed that the steady-state has a constant
slope after approximately 800 seconds.

0.7
Series3
500 A m-2
Series4
350 A m-2
protons in the bulk nbulk,H+ [10-3 mol]

0.6 Series5
200 A m-2
Series6
100 A m-2
0.5 Series7
50 A m-2

0.4

0.3

0.2

0.1

0
0 500 1000 1500 2000
time t [s]

Figure F.1: Unrepeated long time experiments of the observed amount of


protons in the bulk electrolyte as a function of time for different values of the
I
applied current density Aapp
geo

157
Bibliography

[1] Frank Solomon. Electrochemical Production of Finely Divided Metal


Oxides, Metal Hydroxides and Metals, January 10 1978. US Patent
4,067,788.

[2] Atsushi Fujimaru, Toshifumi Mimura, Yutaka Kadowaki, and Katsuhiko


Mushiake. Method of Manufacturing Metal Hydroxides and Method of
Manufacturing ITO Sputtering Target, July 16 2015. US Patent App.
14/394,662.
[3] Jun Zhang, Xia Sheng, Xiqing Cheng, Liping Chen, Jian Jin, and Xinjian
Feng. Robust Electrochemical Metal Oxide Deposition Using an Electrode
with a Superhydrophobic Surface. Nanoscale, 9(1):87–90, 2017.
[4] Frederick Hunt. Production of Hydroxids and Oxids of Metals by
Electrolysis, January 1 1904. US Patent 748,609.
[5] G Helen Annal Therese and P Vishnu Kamath. Electrochemical Synthesis
of Metal Oxides and Hydroxides. Chemistry of Materials, 12(5):1195–1204,
2000.
[6] Igor Zhitomirsky. Cathodic Electrodeposition of Ceramic and
Organoceramic Materials. Fundamental Aspects. Advances in Colloid and
Interface Science, 97(1-3):279–317, 2002.

[7] Gao-Ren Li, Han Xu, Xue-Feng Lu, Jin-Xian Feng, Ye-Xiang Tong, and
Cheng-Yong Su. Electrochemical Synthesis of Nanostructured Materials
for Electrochemical Energy Conversion and Storage. Nanoscale, 5(10):4056–
4069, 2013.
[8] Lok-kun Tsui and Giovanni Zangari. Electrochemical Synthesis of Metal
Oxides for Energy Applications. Springer, New York, US, 2014.

159
160 BIBLIOGRAPHY

[9] Matthew J Lawrence, Adam Kolodziej, and Paramaconi Rodriguez.


Controllable Synthesis of Nanostructured Metal Oxide and Oxyhydroxide
Materials via Electrochemical Methods. Current Opinion in
Electrochemistry, 10:7–15, 2018.
[10] Mustafa Aghazadeh. Cathodic Electrochemical Deposition of Nanostruc-
tured Metal Oxides/Hydroxides and Their Composites for Supercapacitor
Applications: A Review. Analytical & Bioanalytical Electrochemistry,
2019.
[11] A Dierstein, H Natter, F Meyer, H-O Stephan, Ch Kropf, and
R Hempelmann. Electrochemical Deposition under Oxidizing Conditions
(EDOC): A New Synthesis For Nanocrystalline Metal Oxides. Scripta
Materialia, 44(8-9):2209–2212, 2001.
[12] Markus Niederberger, Georg Garnweitner, Jelena Buha, Julien Polleux,
Jianhua Ba, and Nicola Pinna. Nonaqueous Synthesis of Metal Oxide
Nanoparticles: Review and Indium Oxide as Case Study for the
Dependence of Particle Morphology on Precursors and Solvents. Journal
of Sol-Gel Science and Technology, 40(2-3):259–266, 2006.
[13] Georg Garnweitner and Markus Niederberger. Nonaqueous and Surfactant-
free Synthesis Routes to Metal Oxide Nanoparticles. Journal of the
American Ceramic Society, 89(6):1801–1808, 2006.
[14] Igor Djerdj, Denis Arčon, Zvonko Jagličić, and Markus Niederberger.
Nonaqueous Synthesis of Metal Oxide Nanoparticles: Short Review and
Doped Titanium Dioxide as Case Study for the Preparation of Transition
Metal-doped Oxide Nanoparticles. Journal of Solid State Chemistry,
181(7):1571–1581, 2008.
[15] Wei Chen, Debraj Ghosh, and Shaowei Chen. Large-Scale Electrochemical
Synthesis Of SnO2 Nanoparticles. Journal of Materials Science,
43(15):5291–5299, 2008.
[16] Andrea Pucci and Nicola Pinna. Non-aqueous Sol-Gel Routes to Metal
Oxide Nanocrystals under Solvothermal Conditions: Review and Case
Study on Doped Group IV Metal Oxides. Zeitschrift für Naturforschung
B, 65(8):1015–1023, 2010.
[17] AV Nikam, BLV Prasad, and AA Kulkarni. Wet Chemical Synthesis of
Metal Oxide Nanoparticles: a Review. CrystEngComm, 20(35):5091–5107,
2018.
[18] IUPAC. Compendium of Chemical Terminology, 2nd ed. (the ”Gold
Book”). Compiled by A. D. McNaught and A. Wilkinson., volume 528.
Blackwell Scientific Publications, Oxford, UK, 1997.
BIBLIOGRAPHY 161

[19] Jean-Pierre Jolivet, Marc Henry, and Jacques Livage. Metal Oxide
Chemistry and Synthesis: From Solution to Solid State. Wiley-Blackwell,
Chichester, UK, 2000.

[20] Marcel Pourbaix. Atlas of Electrochemical Equilibria in Aqueous Solution.


NACE, Houston, TX, 1974.
[21] Ahsan U Khan. Activated Oxygen: Singlet Molecular Oxygen and
Superoxide Anion. Photochemistry and Photobiology, 28(4-5):615–626,
1978.
[22] Donya Ramimoghadam, Samira Bagheri, and Sharifah Bee Abd
Hamid. Progress in Electrochemical Synthesis of Magnetic Iron Oxide
Nanoparticles. Journal of Magnetism and Magnetic Materials, 368:207–
229, 2014.

[23] Magdalena Skompska and Kamila Zarebska. Electrodeposition of ZnO


Nanorod Arrays on Transparent Conducting Substrates - A Review.
Electrochimica Acta, 127:467–488, 2014.
[24] Carl Rudolf. Process of Electrolytically Separating the Alloys of Silver
with Other Precious or Base Metals, October 9 1928. US Patent 1,687,056.

[25] Tung-Yu Ying, Sotira Yiacoumi, and Costas Tsouris. An Electrochemical


Method for the Formation of Magnetite Particles. Journal of Dispersion
Science and Technology, 23(4):569–576, 2002.
[26] Min Lai and D Jason Riley. Templated Electrosynthesis of Zinc Oxide
Nanorods. Chemistry of Materials, 18(9):2233–2237, 2006.

[27] L Cabrera, S Gutierrez, N Menendez, MP Morales, and P Herrasti.


Magnetite Nanoparticles: Electrochemical Synthesis and Characterization.
Electrochimica Acta, 53(8):3436–3441, 2008.
[28] Mustafa Aghazadeh, Ahmad Nozad Golikand, and Mehdi Ghaemi.
Synthesis, Characterization, and Electrochemical Properties of Ultrafine
β-Ni(OH)2 Nanoparticles. International Journal of Hydrogen Energy,
36(14):8674–8679, 2011.
[29] Peter Hing, Lim Chee Ming, Jung Sang Cheol, and Low Siaw Huei.
Method for Synthesis of Nano-Crystalline Metal Oxide Powders, March 22
2016. US Patent 9,293,230.

[30] Mustafa Aghazadeh, Isa Karimzadeh, Mohammad Reza Ganjali, and


Ahmad Behzad. Mn2+ -Doped Fe3 O4 Nanoparticles: A Novel Preparation
Method, Structural, Magnetic and Electrochemical Characterizations.
162 BIBLIOGRAPHY

Journal of Materials Science: Materials in Electronics, 28(23):18121–


18129, 2017.
[31] Mustafa Aghazadeh. One-Step Cathodic Electrosynthesis of Surface
Capped Fe3 O4 Ultra-Fine Nanoparticles from Ethanol Medium without
Using Coating Agent. Materials Letters, 211:225–229, 2018.
[32] Mustafa Aghazadeh and Mohammad Reza Ganjali. One-Step Electro-
Synthesis of Ni2+ -Doped Magnetite Nanoparticles and Study of Their
Supercapacitive and Superparamagnetic Behaviors. Journal of Materials
Science: Materials in Electronics, 29(6):4981–4991, 2018.
[33] Mustafa Aghazadeh. One-Step Facile Preparation of High Perfor-
mance Ni(OH)2 Nanoparticles/Mesoporous Carbon Nanocomposite:
Electrochemical Synthesis and Properties. Analytical & Bioanalytical
Electrochemistry, 10(5):554–565, 2018.
[34] Zhenhua Yan, Hongming Sun, Xiang Chen, Huanhuan Liu, Yaran Zhao,
Haixia Li, Wei Xie, Fangyi Cheng, and Jun Chen. Anion Insertion
Enhanced Electrodeposition of Robust Metal Hydroxide/Oxide Electrodes
for Oxygen Evolution. Nature Communications, 9(1):1–9, 2018.
[35] Kenneth N Brown. Argentous Oxide, Powder and Method for Making
Same, October 10 1961. US Patent 3,003,935.
[36] R Tom Coyle and Jay A Switzer. Electrochemical Synthesis of Ceramic
Films and Powders, November 21 1989. US Patent 4,882,014.
[37] Roger Lumbroso. Method of Producing Metal Hydroxides, January 24
1995. US Patent 5,384,017.
[38] Yanchun Zhou, Richard J Phillips, and Jay A Switzer. Electrochemical
Synthesis and Sintering of Nanocrystalline Cerium (IV) Oxide Powders.
Journal of the American Ceramic Society, 78(4):981–985, 1995.
[39] Bernd Bauer and Thomas Menzel. Electrochemical Process and Device for
the Production of Metallic Hydroxides and/or Metallic-Oxide Hydroxides,
August 26 1997. US Patent 5,660,709.
[40] Chikafumi Tanaka, Katsuyuki Tanabe, and Shingo Hoshi. Method of and
Apparatus for Producing Ultrafine Metal Compound Particles, May 22
2001. US Patent 6,235,185.
[41] Armin Olbrich, Astrid Görge, Frank Schrumpf, Juliane Meese-
Marktscheffel, Viktor Stoller, Gerhard Gille, Josef Schmoll, Michael Kruft,
and Dirk Naumann. Method for Producing Metal Hydroxides or Alkaline
Metal Carbonates, May 23 2006. US Patent 7,048,843.
BIBLIOGRAPHY 163

[42] Jacqueline E Cloud, Tara S Yoder, Nathan K Harvey, Kyle Snow, and
Yongan Yang. A Simple and Generic Approach for Synthesizing Colloidal
Metal and Metal Oxide Nanocrystals. Nanoscale, 5(16):7368–7378, 2013.

[43] Fabian Kubannek, Thomas Turek, and Ulrike Krewer. Modeling Oxygen
Gas Diffusion Electrodes for Various Technical Applications. Chemie
Ingenieur Technik, 91(6):720–733, 2019.
[44] Yanguang Li and Hongjie Dai. Recent Advances in Zinc–Air Batteries.
Chemical Society Reviews, 43(15):5257–5275, 2014.
[45] Linden David and B Reddy Thomas. Handbook of Batteries. McGraw-Hill
Professional, New York, US, 2001.
[46] F Bidault, DJL Brett, PH Middleton, and NP Brandon. Review of Gas
Diffusion Cathodes for Alkaline Fuel Cells. Journal of Power Sources,
187(1):39–48, 2009.
[47] L Cindrella, Arunachala Mada Kannan, JF Lin, K Saminathan, Y Ho,
CW Lin, and J Wertz. Gas Diffusion Layer for Proton Exchange Membrane
Fuel Cells—a Review. Journal of Power Sources, 194(1):146–160, 2009.

[48] Sehkyu Park, Jong-Won Lee, and Branko N Popov. A Review of Gas
Diffusion Layer in PEM Fuel Cells: Materials and Designs. International
Journal of Hydrogen Energy, 37(7):5850–5865, 2012.
[49] Jaeman Park, Hwanyeong Oh, Taehun Ha, Yoo Il Lee, and Kyoungdoug
Min. A Review of the Gas Diffusion Layer in Proton Exchange Membrane
Fuel Cells: Durability and Degradation. Applied Energy, 155:866–880,
2015.
[50] George W Heise. Air-Depolarized Primary Battery, February 28 1933.
US Patent 1,899,615.
[51] Roy J Plunkett. Tetrafluoroethylene Polymers, February 4 1941. US
Patent 2,230,654.
[52] Walter G Taschek. Electrode for Fuel Cell, July 23 1963. US Patent
3,098,772.
[53] Charles E Thompson. Electrodeposition of Polymers in Porous Electrodes,
December 3 1963. US Patent 3,113,048.

[54] Alain PO Blanchart, Gilbert JL Van Bogaert, Constantine WMVA


De Brandt, and Gustaaf JF Spaepen. Porous Electrode, December 7
1982. US Patent 4,362,790.
164 BIBLIOGRAPHY

[55] M Cifrain and K Kordesch. Hydrogen/Oxygen (Air) Fuel Cells with


Alkaline Electrolytes. Wiley, New York, 2010.

[56] Eddy De Geeter, Michael Mangan, Staf Spaepen, William Stinissen, and
Georges Vennekens. Alkaline Fuel Cells for Road Traction. Journal of
Power Sources, 80(1-2):207–212, 1999.
[57] K Tomantschger, F McClusky, L Oporto, A Reid, and K Kordesch.
Development of Low Cost Alkaline Fuel Cells. Journal of Power Sources,
18(4):317–335, 1986.
[58] Yolanda Alvarez-Gallego, Xochitl Dominguez-Benetton, Deepak Pant,
Ludo Diels, Karolien Vanbroekhoven, Inge Genné, and Philippe Vermeiren.
Development of Gas Diffusion Electrodes for Cogeneration of Chemicals
and Electricity. Electrochimica Acta, 82:415–426, 2012.

[59] Jurgen Garche, Chris K Dyer, Patrick T Moseley, Zempachi Ogumi,


David AJ Rand, and Bruno Scrosati. Encyclopedia of Electrochemical
Power Sources. Elsevier, Amsterdam, The Netherlands, 2013.
[60] RJ Taylor and AA Humffray. Electrochemical Studies on Glassy Carbon
Electrodes: II. Oxygen Reduction in Solutions of High pH (pH>10).
Journal of Electroanalytical Chemistry and Interfacial Electrochemistry,
64(1):63–84, 1975.
[61] JF Pérez, A Galia, MA Rodrigo, Javier Llanos, S Sabatino, C Sáez,
B Schiavo, and O Scialdone. Effect of Pressure on the Electrochemical
Generation of Hydrogen Peroxide in Undivided Cells on Carbon Felt
Electrodes. Electrochimica Acta, 248:169–177, 2017.
[62] Fangke Yu, Yang Chen, Yuwei Pan, Yang Yang, and Hongrui Ma. A Cost-
effective Production of Hydrogen Peroxide via Improved Mass Transfer
of Oxygen for Electro-Fenton Process Using the Vertical Flow Reactor.
Separation and Purification Technology, 241:116695, 2020.

[63] MR Tarasevich, Andrzej Sadkowski, and Ernest Yeager. Oxygen


Electrochemistry. Springer, New York, 1983.
[64] Kim Kinoshita. Electrochemical Oxygen Technology, volume 30. John
Wiley & Sons, New York, 1992.
[65] Yuanyuan Jiang, Pengjuan Ni, Chuanxia Chen, Yizhong Lu, Ping Yang,
Biao Kong, Adrian Fisher, and Xin Wang. Selective Electrochemical H2 O2
Production Through Two-Electron Oxygen Electrochemistry. Advanced
Energy Materials, 8(31):1801909, 2018.
BIBLIOGRAPHY 165

[66] UN Desa et al. Transforming Our World: The 2030 Agenda for Sustainable
Development. 2016.
[67] Aling Chen and Bao Yu Xia. Ambient Dinitrogen Electrocatalytic
Reduction for Ammonia Synthesis. Journal of Materials Chemistry A,
7(41):23416–23431, 2019.
[68] Xiaoxi Guo, Huitong Du, Fengli Qu, and Jinghong Li. Recent Progress in
Electrocatalytic Nitrogen Reduction. Journal of Materials Chemistry A,
7(8):3531–3543, 2019.
[69] Lin Hu, Zhuo Xing, and Xiaofeng Feng. Understanding the
Electrocatalytic Interface for Ambient Ammonia Synthesis. ACS Energy
Letters, 2020.
[70] Antonio José Martı́n, Tatsuya Shinagawa, and Javier Pérez-Ramı́rez.
Electrocatalytic Reduction of Nitrogen: From Haber-Bosch to Ammonia
Artificial Leaf. Chem, 5(2):263–283, 2019.
[71] Xiaoyang Cui, Cheng Tang, and Qiang Zhang. A Review of
Electrocatalytic Reduction of Dinitrogen to Ammonia Under Ambient
Conditions. Advanced Energy Materials, 8(22):1800369, 2018.
[72] Bryan HR Suryanto, Hoang-Long Du, Dabin Wang, Jun Chen, Alexandr N
Simonov, and Douglas R MacFarlane. Challenges and Prospects in the
Catalysis of Electroreduction of Nitrogen to Ammonia. Nature Catalysis,
page 1, 2019.
[73] Masami Shibata, Kohji Yoshida, and Nagakazu Furuya. Electrochemical
Synthesis of Urea at Gas-diffusion Electrodes I. Simultaneous Reduction
of Carbon Dioxide and Nitrite Ions at Zn Catalysts. Denki Kagaku Oyoki
Kogyo Butsuri Kagaku, 64(10):1068–1073, 1996.
[74] Masami Shibata, Kohji Yoshida, and Nagakazu Furuya. Electrochemical
Synthesis of Urea at Gas-Diffusion Electrodes: Part II. Simultaneous
Reduction of Carbon Dioxide and Nitrite Ions at CU, Ag and AU Catalysts.
Journal of Electroanalytical Chemistry, 442(1-2):67–72, 1998.
[75] Masami Shibata, Kohji Yoshida, and Nagakazu Furuya. Electrochemical
Synthesis of Urea at Gas-Diffusion Electrodes III. Simultaneous Reduction
of Carbon Dioxide and Nitrite Ions with Various Metal Catalysts. Journal
of the Electrochemical Society, 145(2):595–600, 1998.
[76] Masami Shibata, Kohji Yoshida, and Nagakazu Furuya. Electrochemical
Synthesis of Urea at Gas-Diffusion Electrodes IV. Simultaneous Reduction
of Carbon Dioxide and Nitrate Ions with Various Metal Catalysts. Journal
of the Electrochemical Society, 145(7):2348–2353, 1998.
166 BIBLIOGRAPHY

[77] Masami Shibata, Kohji Yoshida, and Nagakazu Furuya. Electrochemical


Synthesis of Urea at Gas-Diffusion Electrodes V. Simultaneous Reduction
of Carbon Dioxide and Nitrite Ions with Various Boride Catalysts. Denki
Kagaku oyobi Kogyo Butsuri Kagaku, 66(6):584–589, 1998.
[78] Masami Shibata and Nagakazu Furuya. Electrochemical Synthesis of Urea
at Gas-Diffusion Electrodes: Part VI. Simultaneous Reduction of Carbon
Dioxide and Nitrite Ions with Various Metallophthalocyanine Catalysts.
Journal of Electroanalytical Chemistry, 507(1):177–184, 2001.

[79] Duraisamy Saravanakumar, Jieun Song, Sunhye Lee, Nam Hwi Hur, and
Woonsup Shin. Electrocatalytic Conversion of Carbon Dioxide and Nitrate
Ions to Urea by a Titania-Nafion Composite Electrode. ChemSusChem,
10(20):3999–4003, 2017.

[80] Ronald L Cook, Robert C MacDuff, and Anthony F Sammells. High Rate
Gas Phase CO2 Reduction to Ethylene and Methane Using Gas Diffusion
Electrodes. Journal of the Electrochemical Society, 137(2):607–608, 1990.
[81] Maria Jitaru. Electrochemical Carbon Dioxide Reduction-Fundamental
and Applied Topics. Journal of the University of Chemical Technology
and Metallurgy, 42(4):333–344, 2007.

[82] Devin T Whipple and Paul JA Kenis. Prospects of CO2 Utilization via
Direct Heterogeneous Electrochemical Reduction. The Journal of Physical
Chemistry Letters, 1(24):3451–3458, 2010.
[83] Jinli Qiao, Yuyu Liu, Feng Hong, and Jiujun Zhang. A Review of Catalysts
for the Electroreduction of Carbon Dioxide to Produce Low-Carbon Fuels.
Chemical Society Reviews, 43(2):631–675, 2014.
[84] Sichao Ma, Masaaki Sadakiyo, Raymond Luo, Minako Heima, Miho
Yamauchi, and Paul JA Kenis. One-Step Electrosynthesis of Ethylene
and Ethanol from CO2 in an Alkaline Electrolyzer. Journal of Power
Sources, 301:219–228, 2016.

[85] Jun Li, Guangxu Chen, Yangying Zhu, Zheng Liang, Allen Pei, Chun-Lan
Wu, Hongxia Wang, Hye Ryoung Lee, Kai Liu, Steven Chu, et al. Efficient
Electrocatalytic CO2 Reduction on a Three-Phase Interface. Nature
Catalysis, 1(8):592, 2018.

[86] Phil De Luna, Christopher Hahn, Drew Higgins, Shaffiq A Jaffer,


Thomas F Jaramillo, and Edward H Sargent. What Would It Take for
Renewably Powered Electrosynthesis to Displace Petrochemical Processes?
Science, 364(6438):eaav3506, 2019.
BIBLIOGRAPHY 167

[87] Harry Eccles. Treatment of Metal-Contaminated Wastes: Why Select a


Biological Process? Trends in Biotechnology, 17(12):462–465, 1999.

[88] Jawwad A Darr, Jingyi Zhang, Neel M Makwana, and Xiaole


Weng. Continuous Hydrothermal Synthesis of Inorganic Nanoparticles:
Applications and Future Directions. Chemical Reviews, 117(17):11125–
11238, 2017.
[89] GE Breeman and Arco Timmermans. Political Attention to Environmental
Issues: Analyzing Policy Punctuations in the Netherlands. 2009.
[90] Maria Wächtler, Leticia González, Benjamin Dietzek, Andrey Turchanin,
and Christina Roth. Functional Materials: Making the World Go Round.
Physical Chemistry Chemical Physics, 21(18):8988–8991, 2019.
[91] Kilaru Harsha Vardhan, Ponnusamy Senthil Kumar, and Rames C Panda.
A Review on Heavy Metal Pollution, Toxicity and Remedial Measures:
Current Trends and Future Perspectives. Journal of Molecular Liquids,
page 111197, 2019.
[92] Raymond A Wuana and Felix E Okieimen. Heavy Metals in Contaminated
Soils: A Review of Sources, Chemistry, Risks and Best Available Strategies
for Remediation. Isrn Ecology, 2011, 2011.
[93] Madhu Agarwal and Karam Singh. Heavy Metal Removal from Wastewater
Using Various Adsorbents: A Review. Journal of Water Reuse and
Desalination, 7(4):387–419, 2016.
[94] Francisco R Peligro, Ivana Pavlovic, Ricardo Rojas, and Cristobalina
Barriga. Removal of Heavy Metals from Simulated Wastewater by in Situ
Formation of Layered Double Hydroxides. Chemical Engineering Journal,
306:1035–1040, 2016.
[95] C Femina Carolin, P Senthil Kumar, A Saravanan, G Janet Joshiba,
and Mu Naushad. Efficient Techniques for the Removal of Toxic Heavy
Metals from Aquatic Environment: A Review. Journal of Environmental
Chemical Engineering, 5(3):2782–2799, 2017.
[96] Jaya Mary Jacob, Chinnannan Karthik, Rijuta Ganesh Saratale, Smita S
Kumar, Desika Prabakar, K Kadirvelu, and Arivalagan Pugazhendhi.
Biological Approaches to Tackle Heavy Metal Pollution: A Survey of
Literature. Journal of Environmental Management, 217:56–70, 2018.
[97] Young Ku and In-Liang Jung. Photocatalytic Reduction of Cr(VI) in
Aqueous Solutions by UV Irradiation with the Presence of Titanium
Dioxide. Water Research, 35(1):135–142, 2001.
168 BIBLIOGRAPHY

[98] Fenglian Fu and Qi Wang. Removal of Heavy Metal Ions from Wastewaters:
A Review. Journal of Environmental Management, 92(3):407–418, 2011.
[99] Arezoo Azimi, Ahmad Azari, Mashallah Rezakazemi, and Meisam
Ansarpour. Removal of Heavy Metals from Industrial Wastewaters: A
Review. ChemBioEng Reviews, 4(1):37–59, 2017.
[100] Jian-Feng Chen, Yu-Hong Wang, Fen Guo, Xin-Ming Wang, and Chong
Zheng. Synthesis of Nanoparticles with Novel Technology: High-Gravity
Reactive Precipitation. Industrial & Engineering Chemistry Research,
39(4):948–954, 2000.
[101] Marcos Fernández-Garcı́a and José A Rodriguez. Metal Oxide
Nanoparticles. Wiley, New Jersey, 2011.
[102] SK Tripathi, Ramneek Kaur, and Mamta Rani. Oxide Nanomaterials and
Their Applications as a Memristor. In Solid State Phenomena, volume
222, pages 67–97. Trans Tech Publ, 2015.
[103] Margherita Izzi, Maria Chiara Sportelli, Nicoletta Ditaranto, Rosaria Anna
Picca, Massimo Innocenti, Luigia Sabbatini, and Nicola Cioffi. On the
Pros and Cons of the Sacrificial Anode Electrolysis for the Preparation of
Transition Metal Colloids. ChemElectroChem, 2019.
[104] David R Lide and Henry V Kehiaian. CRC Handbook of Thermophysical
and Thermochemical Data. CRC Press, Boca Raton, FL, 1994.
[105] GWC Kaye and TH Laby. Tables of Physical and Chemical Constants
15th ed. Longman, New York, 1986.
[106] Kathryn C Sole. The Evolution of Cobalt–Nickel Separation and
Purification Technologies: Fifty Years of Solvent Extraction and Ion
Exchange. Springer, Cham, 2018.
[107] Pradyot Patnaik. Handbook of Inorganic Chemicals. McGraw-Hill, New
York, 2003.
[108] David Lide. CRC Handbook of Chemistry and Physics, 90th Edition. Boca
Raton, FL, 2009.
[109] Jerzy Chlistunoff. Advanced Chlor-Alkali Technology. Final Technical
Report, 2004.
[110] Imad Moussallem, Jakob Jörissen, Ulrich Kunz, Stefan Pinnow, and
Thomas Turek. Chlor-Alkali Electrolysis with Oxygen Depolarized
Cathodes: History, Present Status and Future Prospects. Journal of
Applied Electrochemistry, 38(9):1177–1194, 2008.
BIBLIOGRAPHY 169

[111] Shyam Lakshmanan and Thanapalan Murugesan. The Chlor-Alkali


Process: Work in Progress. Clean Technologies and Environmental Policy,
16(2):225–234, 2014.

[112] Johannes Jung, Sarah Postels, and André Bardow. Cleaner Chlorine
Production Using Oxygen Depolarized Cathodes? A Life Cycle
Assessment. Journal of Cleaner Production, 80:46–56, 2014.
[113] Juergen Kintrup, Marta Millaruelo, Vinh Trieu, Andreas Bulan, and
Ernesto Silva Mojica. Gas Diffusion Electrodes for Efficient Manufacturing
of Chlorine and Other Chemicals. The Electrochemical Society Interface,
26(2):73–76, 2017.
[114] Fourth Edition. Guidelines for Drinking-Water Quality. WHO Chronicle,
38(4):104–8, 2011.

[115] Sameeh A Mansour. Heavy Metal Contamination as a Global Problem


and the Need for Prevention/Reduction Measurements. John Wiley &
Sons, Ltd., Chichester, UK, 2014.
[116] Jacco L Huisman, Gerard Schouten, and Carl Schultz. Biologically
Produced Sulphide for Purification of Process Streams, Effluent
Treatment and Recovery of Metals in the Metal and Mining Industry.
Hydrometallurgy, 83(1-4):106–113, 2006.
[117] Karlis A Baltpurvins, Robert C Burns, Geoffrey A Lawrance, and Alan D
Stuart. Use of the Solubility Domain Approach for the Modeling of the
Hydroxide Precipitation of Heavy Metals from Wastewater. Environmental
Science & Technology, 30(5):1493–1499, 1996.
[118] Nathalie Meunier, Patrick Drogui, Camille Montané, Robert Hausler, Guy
Mercier, and Jean-François Blais. Comparison Between Electrocoagulation
and Chemical Precipitation for Metals Removal from Acidic Soil Leachate.
Journal of Hazardous Materials, 137(1):581–590, 2006.

[119] Daniel C Harris. Quantitative Chemical Analysis. Macmillan, New York,


2010.
[120] Oliver Totsche, Andrew Fyson, Margarete Kalin, and Christian Steinberg.
Titration Curves: A Useful Instrument for Assessing the Buffer Systems
of Acidic Mining Waters. Environmental Science and Pollution Research,
13(4):215–224, 2006.
[121] J Sánchez España, E López Pamo, E Santofimia Pastor, J Reyes
Andrés, and JA Martı́n Rubı́. The Removal of Dissolved Metals by
Hydroxysulphate Precipitates During Oxidation and Neutralization of
170 BIBLIOGRAPHY

Acid Mine Waters, Iberian Pyrite Belt. Aquatic Geochemistry, 12(3):269–


298, 2006.
[122] Robert de Levie. Explicit Expressions of the General Form of the
Titration Curve in Terms of Concentration: Writing a Single Closed-
Form Expression for the Titration Curve for a Variety of Titrations
Without Using Approximations or Segmentation. Journal of Chemical
Education, 70(3):209, 1993.
[123] Daniel A Morales. Mathematical Modeling of Titration Curves. Journal
of Chemometrics, 16(5):247–260, 2002.
[124] David L Parkhurst and CAJ Appelo. User’s Guide to Phreeqc (Version 2)–
A Computer Program for Speciation, Batch-Reaction, One-Dimensional
Transport, and Inverse Geochemical Calculations. Water-Resources
Investigations Report, 99(4259):312, 1999.
[125] Jerry D Allison, David S Brown, and Kevin J Novo-Gradac.
MINTEQA2/PROFEDA2, a Geochemical Assessment Model for
Environmental Systems: Version 3.0 User’s Manual. Environmental
Research Laboratory, Office of Research and Development, Georgia, 1991.
[126] Virginia Herndon. MINTEQA2/PRODEFA2, A Geochemical Assessment
Model for Environmental Systems: User Manual Supplement for Version
4.0. Georgia, US, 1998.
[127] Niels Henrik Abel. Mémoire sur les Équations Algébrique: où on Démontre
l’Impossiblité de la Résolution de l’Equation Générale du Cinquième Dégré.
Librarian, Faculty of Science, University of Oslo, 1824.
[128] Galo J de AA Soler-Iltia, Matias Jobbagy, Roberto J Candal, Alberto E
Regazzoni, and Miguel A Blesa. Synthesis of Metal Oxide Particles from
Aqueous Media: The Homogeneous Alkalinization Method. Journal of
Dispersion Science and Technology, 19(2-3):207–228, 1998.
[129] Galo J de AA Soler-Illia, Matı́as Jobbágy, Alberto E Regazzoni, and
Miguel A Blesa. Synthesis of Nickel Hydroxide by Homogeneous
Alkalinization. Precipitation Mechanism. Chemistry of Materials,
11(11):3140–3146, 1999.
[130] Roger Gordon Bates. Determination of pH: Theory and Practice. John
Wiley & Sons, Inc., New York, London, Sydney, 1964.
[131] Marcus C Simoes, Kevin J Hughes, Derek B Ingham, Lin Ma, and
Mohamed Pourkashanian. Estimation of the Pitzer Parameters for 1–1,
2–1, 3–1, 4–1, and 2–2 Single Electrolytes at 25 ◦ C. Journal of Chemical
& Engineering Data, 61(7):2536–2554, 2016.
BIBLIOGRAPHY 171

[132] Qiaobao Zhang, Kaili Zhang, Daguo Xu, Guangcheng Yang, Hui Huang,
Fude Nie, Chenmin Liu, and Shihe Yang. Cuo Nanostructures: Synthesis,
Characterization, Growth Mechanisms, Fundamental Properties, and
Applications. Progress in Materials Science, 60:208–337, 2014.
[133] M Mohapatra and S Anand. Synthesis and Applications of Nano-
Structured Iron Oxides/Hydroxides–A Review. International Journal
of Engineering, Science and Technology, 2(8), 2010.
[134] Victor K LaMer and Robert H Dinegar. Theory, Production and
Mechanism of Formation of Monodispersed Hydrosols. Journal of the
American Chemical Society, 72(11):4847–4854, 1950.
[135] Arne E Nielsen. Kinetics of Precipitation, volume 18. Pergamon, Oxford,
New York, 1964.
[136] Tadao Sugimoto. Preparation of Monodispersed Colloidal Particles.
Advances in Colloid and Interface Science, 28:65–108, 1987.
[137] JA Dirksen and TA Ring. Fundamentals of Crystallization: Kinetic Effects
on Particle Size Distributions and Morphology. Chemical Engineering
Science, 46(10):2389–2427, 1991.
[138] JP Klein and R David. Reaction Crystallization. Marcel Dekker, New
York, 1995.
[139] GP Demopoulos. Aqueous Precipitation and Crystallization for the
Production of Particulate Solids with Desired Properties. Hydrometallurgy,
96(3):199–214, 2009.
[140] James J De Yoreo, Pupa UPA Gilbert, Nico AJM Sommerdijk, R Lee
Penn, Stephen Whitelam, Derk Joester, Hengzhong Zhang, Jeffrey D
Rimer, Alexandra Navrotsky, Jillian F Banfield, et al. Crystallization by
Particle Attachment in Synthetic, Biogenic, and Geologic Environments.
Science, 349(6247):aaa6760, 2015.
[141] Nguyen TK Thanh, N Maclean, and S Mahiddine. Mechanisms of
Nucleation and Growth of Nanoparticles in Solution. Chemical Reviews,
114(15):7610–7630, 2014.
[142] Renguo Xie, Zheng Li, and Xiaogang Peng. Nucleation Kinetics vs.
Chemical Kinetics in the Initial Formation of Semiconductor Nanocrystals.
Journal of the American Chemical Society, 131(42):15457–15466, 2009.
[143] Jean-Pierre Jolivet, Sophie Cassaignon, Corinne Chanéac, David Chiche,
and Elisabeth Tronc. Design of Oxide Nanoparticles by Aqueous Chemistry.
Journal of Sol-Gel Science and Technology, 46(3):299–305, 2008.
172 BIBLIOGRAPHY

[144] Manuel D Montaño, John W Olesik, Angela G Barber, Katie Challis, and
James F Ranville. Single Particle ICP-MS: Advances Toward Routine
Analysis of Nanomaterials. Analytical and Bioanalytical Chemistry,
408(19):5053–5074, 2016.
[145] A Ortega and J Garcıa de la Torre. Hydrodynamic Properties of Rodlike
and Disklike Particles in Dilute Solution. The Journal of Chemical Physics,
119(18):9914–9919, 2003.
[146] Richard C Murdock, Laura Braydich-Stolle, Amanda M Schrand, John J
Schlager, and Saber M Hussain. Characterization of Nanomaterial
Dispersion in Solution Prior to in Vitro Exposure Using Dynamic Light
Scattering Technique. Toxicological Sciences, 101(2):239–253, 2008.
[147] Puthusserickal A Hassan, Suman Rana, and Gunjan Verma. Making
Sense of Brownian Motion: Colloid Characterization by Dynamic Light
Scattering. Langmuir, 31(1):3–12, 2014.
[148] Harry B Weiser, WO Milligan, and EL Cook. Hydrous Cupric Hydroxide
and Basic Cupric Sulfates. Journal of the American Chemical Society,
64(3):503–508, 1942.
[149] Yannick Cudennec and André Lecerf. The Transformation of Cu(OH)2
Into CuO, Revisited. Solid State Sciences, 5(11-12):1471–1474, 2003.
[150] Hui Liu, Ping Li, Bin Lu, Yu Wei, and Yuhan Sun. Transformation of
Ferrihydrite in the Presence or Absence of Trace Fe (II): The Effect of
Preparation Procedures of Ferrihydrite. Journal of Solid State Chemistry,
182(7):1767–1771, 2009.
[151] Bai-Sheng Zhu, Yong Jia, Zhen Jin, Bai Sun, Tao Luo, Ling-Tao Kong,
and Jin-Huai Liu. A Facile Precipitation Synthesis of Mesoporous 2-Line
Ferrihydrite with Good Fluoride Removal Properties. RSC Advances,
5(103):84389–84397, 2015.
[152] Soumya Das, M Jim Hendry, and Joseph Essilfie-Dughan. Transformation
of Two-Line Ferrihydrite to Goethite and Hematite as a Function of pH
and Temperature. Environmental Science & Technology, 45(1):268–275,
2010.
[153] James W Ball, D Kirk Nordstrom, et al. User’s Manual for wateq4f,
with Revised Thermodynamic Data Base and Test Cases for Calculating
Speciation of Major, Trace, and Redox Elements in Natural Waters. 1991.
[154] John O Edwards and Ralph G Pearson. The Factors Determining
Nucleophilic Reactivities. Journal of the American Chemical Society,
84(1):16–24, 1962.
BIBLIOGRAPHY 173

[155] E Neyens and J Baeyens. A Review of Classic Fenton’s Peroxidation


as an Advanced Oxidation Technique. Journal of Hazardous Materials,
98(1-3):33–50, 2003.
[156] Alok D Bokare and Wonyong Choi. Review of Iron-Free Fenton-Like
Systems for Activating h2o2 in Advanced Oxidation Processes. Journal
of Hazardous Materials, 275:121–135, 2014.
[157] Hanna Laura Wiegand, Christian Timon Orths, Klaus Kerpen,
Holger Volker Lutze, and Torsten Claus Schmidt. Investigation of
the Iron–Peroxo Complex in the Fenton Reaction: Kinetic Indication,
Decay Kinetics, and Hydroxyl Radical Yields. Environmental Science &
Technology, 51(24):14321–14329, 2017.
[158] Geoffrey Wilkinson, Robert D Gillard, and Jon A McCleverty.
Comprehensive Coordination Chemistry. The Synthesis, Reactions,
Properties and Applications of Coordination Compounds. V. 3. Main
Group and Early Transition Elements. Pergamon, Oxford, UK, 1987.
[159] Rochelle M Cornell, Rudolf Giovanoli, and Walter Schneider. Review of
the Hydrolysis of Iron (III) and the Crystallization of Amorphous Iron
(III) Hydroxide Hydrate. Journal of Chemical Technology & Biotechnology,
46(2):115–134, 1989.

[160] Andri Stefánsson. Iron (III) Hydrolysis and Solubility at 25 C.
Environmental Science & Technology, 41(17):6117–6123, 2007.
[161] David T Richens. Ligand Substitution Reactions at Inorganic Centers.
Chemical Reviews, 105(6):1961–2002, 2005.
[162] Shadpour Mallakpour and Maryam Madani. A Review of Current Coupling
Agents for Modification of Metal Oxide Nanoparticles. Progress in Organic
Coatings, 86:194–207, 2015.
[163] José A Rodrı́guez and Marcos Fernández-Garcı́a. Synthesis, Properties,
and Applications of Oxide Nanomaterials. John Wiley & Sons, New York,
2007.
[164] Sam GF Eggermont, Ana Rua-Ibarz, Kristof Tirez, Xochitl Dominguez-
Benetton, and Jan Fransaer. Oxidation-Assisted Alkaline Precipitation:
The Effect of H2 O2 on the Size of CuO and FeOOH Nanoparticles. RSC
Advances, 9(51):29902–29908, 2019.
[165] Catherine Henrist, J-P Mathieu, C Vogels, André Rulmont, and Rudi
Cloots. Morphological Study of Magnesium Hydroxide Nanoparticles
Precipitated in Dilute Aqueous Solution. Journal of Crystal Growth,
249(1-2):321–330, 2003.
174 BIBLIOGRAPHY

[166] Balasubramaniam Gnana Sundara Raj, Abdullah M Asiri, Jerry J Wu, and
Sambandam Anandan. Synthesis of Mn3 O4 Nanoparticles via Chemical
Precipitation Approach for Supercapacitor Application. Journal of Alloys
and Compounds, 636:234–240, 2015.
[167] SF Hasany, I Ahmed, J Rajan, and A Rehman. Systematic Review of the
Preparation Techniques of Iron Oxide Magnetic Nanoparticles. Nanosci.
Nanotechnol, 2(6):148–158, 2012.
[168] Sam GF Eggermont, Rafael Prato, Xochitl Dominguez-Benetton, and
Jan Fransaer. Metal Removal from Aqueous Solutions: Insights from
Modeling Precipitation Titration Curves. Journal of Environmental
Chemical Engineering, page 103569, 2019.
[169] Rafael A Prato, Vincent Van Vught, Sam Eggermont, Guillermo Pozo,
Pilar Marin, Jan Fransaer, and Xochitl Dominguez-Benetton. Gas
Diffusion Electrodes on the Electrosynthesis of Controllable Iron Oxide
Nanoparticles. Scientific Reports, 9(1):1–11, 2019.
[170] Yohannes Kiros, Massoud Pirjamali, and Martin Bursell. Oxygen
Reduction Electrodes for Electrolysis in Chlor-Alkali Cells. Electrochimica
Acta, 51(16):3346–3350, 2006.
[171] Peter Woltering, Philipp Hofmann, Frank Funck, Randolf Kiefer, Ulf-
Steffen Baeumer, Dmitri Donst, and Carsten Schmitt. Energy-Saving
Chlorine Production. Chlor-Alkali Electrolysis Using Innovative Cathode
Technology; Energiesparende Chlorpoduktion. Chlor-Alkali-Elektrolyse
Mit Innovativer Kathoden-Technologie. 2013.
[172] M Paidar, V Fateev, and K Bouzek. Membrane Electrolysis—History,
Current Status and Perspective. Electrochimica Acta, 209:737–756, 2016.
[173] Hella Boumaiza, Romain Coustel, Ghouti Medjahdi, Christian Ruby, and
Latifa Bergaoui. Conditions for the Formation of Pure Birnessite During
the Oxidation of Mn (II) Cations in Aqueous Alkaline Medium. Journal
of Solid State Chemistry, 248:18–25, 2017.
[174] Zheng-Rong Tian, Wei Tong, Jin-Yun Wang, Nian-Gao Duan,
Venkatesan V Krishnan, and Steven L Suib. Manganese Oxide
Mesoporous Structures: Mixed-Valent Semiconducting Catalysts. Science,
276(5314):926–930, 1997.
[175] Darminto, Machida N Cholishoh, Feby A Perdana, Malik A Baqiya,
Mashuri, Yoyok Cahyono, and Triwikantoro. Preparing Fe3 O4
Nanoparticles from Fe2+ Ions Source by Co-Precipitation Process in
Various pH, volume 1415. 2011.
BIBLIOGRAPHY 175

[176] Kai Sundmacher and Thorsten Schultz. Macrokinetic Analysis of


Polarisation Characteristics of Gas-Diffusion Electrodes in Contact with
Liquid Electrolytes, Part II: Oxygen Reduction as Example for a Higher
Order Reaction. Journal of Applied Electrochemistry, 35(12):1171–1181,
2005.
[177] Rene Massart. Preparation of Aqueous Magnetic Liquids in Alkaline and
Acidic Media. IEEE Transactions on Magnetics, 17(2):1247–1248, 1981.

[178] Mingsong Wang, Lingxia Jiang, Eui Jung Kim, and Sung Hong
Hahn. Electronic Structure and Optical Properties of Zn(OH)2 :
LDA+U Calculations and Intense Yellow Luminescence. RSC Advances,
5(106):87496–87503, 2015.
[179] A Khorsand Zak, WH Abd Majid, M Ebrahimizadeh Abrishami, and
Ramin Yousefi. X-Ray Analysis of ZnO Nanoparticles by Williamson–Hall
and Size–Strain Plot Methods. Solid State Sciences, 13(1):251–256, 2011.
[180] Mingsong Wang, Yajun Zhou, Yiping Zhang, Sung Hong Hahn, and
Eui Jung Kim. From Zn(OH)2 to Zno: A Study on the Mechanism of
Phase Transformation. CrystEngComm, 13(20):6024–6026, 2011.

[181] YQ Chang, XY Xu, XH Luo, CP Chen, and DP Yu. Synthesis and


Characterization of Mn3 O4 Nanoparticles. Journal of Crystal Growth,
264(1-3):232–236, 2004.
[182] Je-Hun Jang and Susan L Brantley. Investigation of Wustite (FeO)
Dissolution: Implications for Reductive Dissolution of Ferric Oxides.
Environmental Science & Technology, 43(4):1086–1090, 2009.
[183] SK Sharma, JM Vargas, KR Pirota, Shalendra Kumar, CG Lee, and
M Knobel. Synthesis and Ageing Effect in FeO Nanoparticles: Transfor-
mation to Core–Shell FeO/Fe3 O4 and Their Magnetic Characterization.
Journal of Alloys and Compounds, 509(22):6414–6417, 2011.

[184] Wei Jiang, Fei Liang, Jianwei Wang, Lei Su, Yaoming Wu, and Limin
Wang. Enhanced Electrochemical Performances of FeO X–Graphene
Nanocomposites as Anode Materials for Alkaline Nickel–Iron Batteries.
RSC Advances, 4(30):15394–15399, 2014.
[185] Jing Xu, Haibin Yang, Wuyou Fu, Kai Du, Yongming Sui, Jiuju Chen,
Yi Zeng, Minghui Li, and Guangtian Zou. Preparation and Magnetic
Properties of Magnetite Nanoparticles by Sol–Gel Method. Journal of
Magnetism and Magnetic Materials, 309(2):307–311, 2007.
176 BIBLIOGRAPHY

[186] SP Jiang, WR Ashton, and ACC Tseung. An Observation of Homogeneous


and Heterogeneous Catalysis Processes in the Decomposition of H2 O2
Over MnO2 and Mn(OH)2 . Journal of Catalysis, 131(1):88–93, 1991.

[187] James W Ball and D Kirk Nordstrom. Wateq4f–User’s Manual with


Revised Thermodynamic Data Base and Test Cases for Calculating
Speciation of Major, Trace and Redox Elements in Natural Waters.
Technical report, 1991.

[188] J Johnson, G Anderson, and D Parkhurst. Database “thermo. com. V8. R6.
230,” Rev. 1-11. Lawrence Livermore Natl. Lab., Livermore, California,
2000.
[189] John Newman and Karen E Thomas-Alyea. Electrochemical Systems.
John Wiley & Sons, New York, 2012.

[190] Ewald Heitz and Gerhard Kreysa. Principles of Electrochemical


Engineering. VCH Verlagsgesellschaft, Weinheim, 1986.
[191] Song Hi Lee and Jayendran C Rasaiah. Proton Transfer and the Mobilities
of the H+ and OH− Ions from Studies of a Dissociating Model for Water.
The Journal of Chemical Physics, 135(12):124505, 2011.

[192] David MH Kern. The Polarography and Standard Potential of the Oxygen-
Hydrogen Peroxide Couple. Journal of the American Chemical Society,
76(16):4208–4214, 1954.
[193] Noam Agmon. The Grotthuss Mechanism. Chemical Physics Letters,
244(5-6):456–462, 1995.

[194] P Shaw and A Donard. Nano-Particle Analysis Using Dwell Times


Between 10 µS and 70 µS with an Upper Counting Limit of Greater than
3× 107 CPS and a Gold Nanoparticle Detection Limit of Less than 10 nm
Diameter. Journal of Analytical Atomic Spectrometry, 31(6):1234–1242,
2016.
Curriculum Vitae

Personalia
Eggermont Sam
1989/08/04
Belgium

Education
2016−2020 KU Leuven Doctoral student of Engineering - Materials Engineering

2012−2013 KU Leuven Advanced Master of Safety Engineering


2010−2012 KU Leuven Master of Engineering - Materials Engineering
2008−2010 KU Leuven Bachelor of Engineering - Materials Engineering
2007−2008 KU Leuven Bachelor of Physics - Option Engineering
KULAK

Positions
2015−2016 KU Leuven Academic researcher
2013−2014 Jan De Nul QHSSE advisor of dredging activities

177
178 CURRICULUM VITAE

Teaching activities

2016−2017 KU Leuven P & O module 5 - Chemical analytic techniques


2016−2018 KU Leuven Exercises - Mechanical behavior of materials
2016−2019 KU Leuven Exercises - Introduction to materials science
KULAK
2017−2019 KU Leuven Support - Introduction to material science

Institutional responsibilities

2016−2017 KU Leuven POC council


2016−2019 KU Leuven Department council
2017−2018 KU Leuven Chair MAC meeting
2018 KU Leuven Department board

Membership of scientific societies

2016−2018 ECS student chapter Belgium


2019 − ... Royal Society of Chemistry
List of publications

This is a list of all publications to which the author has contributed up until
now:

Eggermont, S.G., Rua-Ibarz, A., Tirez, K., Dominguez-Benetton, X.,


Fransaer, J. Oxidation-assisted Alkaline Precipitation: the Effect of
H2 O2 on the Size of CuO and FeOOH Nanoparticles. RSC Advances.
2019;9(51):29902-8. doi: 10.1039/C9RA03086G
Lippiatt, N., Ling, T.C., Eggermont, S.G. Combining Hydration
and Carbonation of Cement Using Super-saturated Aqueous CO2
Aolution. Construction and Building Materials. 2019;(229):116825. doi:
10.1016/j.conbuildmat.2019.116825
Prato, R.A., Van Vught, V., Eggermont, S.G., Pozo, G., Marin, P.,
Fransaer, J., Dominguez-Benetton, X. Gas-Diffusion Electrodes on the
Electrosynthesis of Controllable Iron Oxide Nanoparticles. Scientific
Reports. (2019); 9(1), 1-11. doi: 10.1038/s41598-019-51185-x
Eggermont, S.G., Prato, R.A., Dominguez-Benetton, X., Fransaer,
J. Metal Removal from Aqueous Solutions: Insights from Modeling
Precipitation Titration Curves. Journal of Environmental Chemical
Engineering. doi: 10.1016/j.jece.2019.103596

This is a list of all the conferences to which the author has contributed up until
now:

Eggermont, S.G., Prato, R.F., Dominguez-Benetton, X., Fransaer, J.


Models and Experiments for the Analysis of ORR Product Accumulation
within Porous Oxygen Gas-Diffusion Electrodes. Symposium on
Insights into Gas-Diffusion Electrodes: from Fundamentals to Industrial
Applications. 2019/5/23

179
180 LIST OF PUBLICATIONS

Giels, M., Alarcón, R.M., Eggermont, S.G., Hertel, T., Pontikes, Y.


Heavy Metal Release from Porous Inorganic Polymers Made from Fe-rich
Slag: Effect of Al Content and Curing Time. In Proceedings of the 6th
international Slag Valorisation Symposium. (2019); 183-186.
FACULTY OF ENGINEERING SCIENCE
DEPARTMENT OF MATERIALS ENGINEERING
SURFACE AND INTERFACE ENGINEERED MATERIALS
Kasteelpark Arenberg 44 - box 2450
B-3001 Leuven
www.mtm.kuleuven.be/Onderzoeksgroepen/SIEM

You might also like