Download as pdf or txt
Download as pdf or txt
You are on page 1of 235

University of Calgary

PRISM Repository https://prism.ucalgary.ca


The Vault Open Theses and Dissertations

2018-06-11

Regeneration of Carbon Materials


Saturated with Organic Pollutants by
Fenton Oxidation Based Technologies

Xiao, Ye

Xiao, Y. (2018). Regeneration of Carbon Materials Saturated with Organic Pollutants by Fenton
Oxidation Based Technologies (Doctoral thesis, University of Calgary, Calgary, Canada).
Retrieved from https://prism.ucalgary.ca. doi:10.11575/PRISM/31999
http://hdl.handle.net/1880/106773
Downloaded from PRISM Repository, University of Calgary
UNIVERSITY OF CALGARY

Regeneration of Carbon Materials Saturated with Organic Pollutants by

Fenton Oxidation Based Technologies

by

Ye Xiao

A THESIS

SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

IN PARTIAL FULFILMENT OF THE REQUIREMENTS FOR THE

DEGREE OF DOCTOR OF PHILOSOPHY

GRADUATE PROGRAM IN CHEMICAL AND PETROLEUM ENGINEERING

CALGARY, ALBERTA

JUNE, 2018

© Ye Xiao 2018
Abstract

Organic pollutants, which may be carcinogenic and endocrine disrupting, are frequently

detected in different water bodies. Carbon adsorption is an effective and efficient approach to

remove organic pollutants from water even for the toxic and bio-refractory ones. After

adsorption, it is economically and environmentally beneficial to regenerate the carbon materials.

Fenton and electro-Fenton oxidation are promising technologies for regeneration because they

can regenerate the adsorbents and degrade the adsorbed organic pollutants simultaneously under

ambient conditions. In this thesis, the Fenton and electro-Fenton oxidation were studied for the

regeneration of carbon saturated with methyl orange (MO) with a focus on the impact of

adsorbent properties and operational parameters.

First, the impact of pore structure on Fenton regeneration was investigated using γ-Fe2O3

loaded carbon with different microporosity. The fresh and regenerated carbon materials were

characterized by adsorption isotherm and kinetics measurements. The regeneration efficiency

decreased with increasing microporosity; essentially complete regeneration was achieved with

the mesoporous carbon. The results were consistent with the MO adsorbed in micropores being

strongly bond and difficult to remove using the Fenton regeneration method.

As an attempt to regenerate the adsorption sites in micropores, carbon materials were

regenerated using electro-Fenton oxidation. The regeneration efficiency was improved with

higher cathodic potential, better contact with cathode, and lower microporosity. Compared with

Fenton oxidation, electro-Fenton oxidation could regenerate the adsorption sites in the

micropores < 1 nm. The results indicated that electro-Fenton regeneration was controlled by the

electro-desorption process.

i
Because better contact with the cathode was necessary for electro-Fenton regeneration,

activated carbon - polytetrafluoroethylene electrodes were proposed to act as both the adsorbent

for adsorption and the cathode for electro-Fenton regeneration. MO adsorption capacity of 176

mg/g and regeneration efficiency of 81% were obtained using one of these electrodes. A higher

MO adsorption rate and regeneration efficiency were obtained with a higher electrode

hydrophilicity.

Overall, this thesis provides mechanistic insights on the regeneration of carbon materials by

the Fenton and electro-Fenton oxidation, which can guide further research and practical

application of these methods in carbon regeneration.

ii
Acknowledgements

First and foremost, I would like to express my sincere thanks to my supervisor, Dr. Josephine

Mary Hill. Dr. Hill is a respected scientist and a supportive supervisor. Not only did she provided

valuable suggestions on my research, but she also guided me to a better researcher. Every time

meeting with her, she illustrated new ideas and advices in detail on several pieces of paper so

that I could fully understand. Every time after meeting with her, I felt more clear and confident

in my research. In addition, she also spent plenty of time to teach me how to do research and

how to advance in academia. Thank you, Dr. Hill, for your precious guidance.

Secondly, I would like to convey my gratitude to my committee members Dr. Edward Roberts

and Dr. Sathish Ponnurangam for their time and valuable suggestions through my PhD research.

I also want to acknowledge Dr. Gordon Chua, Dr. Maen Husein, Dr. Angus Chu, Dr. Charles Q.

Jia for being examiners of my candidacy exam and final defense. I am grateful for the

scholarship from the Department of Chemical and Petroleum Engineering, University of

Calgary. The financial support from the Canada Research Chairs program for equipment and

operating funds is gratefully acknowledged.

The kind help and support from Arthur de Vera and Suha Abusalim during my whole PhD study

is very much appreciated. My sincere thanks also go to everyone in LECA group: Dr. Luis Daniel

Virla Alvarado, Ross Arnold, Dr. Tazul Islam Bhuiyan, Dr. Melanie Hazlett, Qing Huang, Dr.

Vincente Montes Jimenez, Richard Gerald Kaldenhoven, Sip Chen Liew, Javier Mora, Sreelakshimi

Puthoor, Sebastian Sessarego, Jingfeng Wu. Special thanks go to Sip and Tazul for helping me

setting up computer and buying chemicals, to Ross for teaching me how to use TGA equipment, to

Jingfeng for teaching me how to use TriStar, to Sebastian and Vincente for all the discussions about

iii
my research. I am also grateful for the support and encouragement from my previous supervisor

Jianguo Jiang from Tsinghua University and Wenpo Shan from Chinese Academy of Science.

I am also very thankful to my parents and other family members for their endless,

unconditional, and unreserved love. Their encouragement is always a source of my motivation.

Special thanks to my beautiful, adorable, and talented fiancée for her constant encouragement

and support.

iv
Dedication

To my parents.

v
Table of Contents

Abstract ........................................................................................................................... i

Acknowledgements ....................................................................................................... iii

Dedication .......................................................................................................................v

Table of Contents .......................................................................................................... vi

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

List of Figures ............................................................................................................. xiii

List of Abbreviations.................................................................................................. xvii

CHAPTER ONE: INTRODUCTION ..................................................................................1

CHAPTER TWO: LITERATURE REVIEW ....................................................................10

2.1 Organic pollutants in water system ..........................................................................10

2.2 Adsorption of organic pollutants .............................................................................11

2.2.1 Adsorption system ...........................................................................................11

2.2.2 Activated carbon ..............................................................................................14

2.2.3 Biochar ............................................................................................................15

2.2.4 Emerging carbon materials ..............................................................................16

2.2.5 Mechanism of carbon adsorption ....................................................................17

2.2.6 Challenge of carbon adsorption and strategies ................................................20

2.3 Regeneration ............................................................................................................21

2.3.1 Principle and classification ..............................................................................21

2.3.2 Regeneration methods .....................................................................................22

2.3.3 Regeneration methods comparison..................................................................23

2.4 Fenton oxidation ......................................................................................................26

2.4.1 Homogeneous ..................................................................................................26

vi
2.4.2 Heterogenous ...................................................................................................27

2.4.3 Fenton oxidation regeneration .........................................................................28

2.5 Electro-Fenton oxidation .........................................................................................30

2.5.1 Principles .........................................................................................................30

2.5.2 Cathode ............................................................................................................32

2.5.3 Cathodic regeneration ......................................................................................35

2.6 Knowledge gaps .......................................................................................................39

CHAPTER THREE: MATERIALS AND METHODS ....................................................40

3.1 Methodology ............................................................................................................40

3.2 Materials ..................................................................................................................41

3.2.1 Methyl orange ..................................................................................................41

3.2.2 Preparation of magnetic carbon materials .......................................................42

3.2.3 Preparation of carbon-PTFE electrodes ...........................................................43

3.3 Characterization of materials ...................................................................................44

3.3.1 Nitrogen adsorption-desorption isotherms ......................................................44

3.3.2 X-ray diffraction ..............................................................................................46

3.3.3 Thermogravimetric analysis ............................................................................46

3.3.4 Fourier-transform infrared spectroscopy .........................................................47

3.4 Analysis of solutions ................................................................................................47

3.4.1 MO solution and regeneration solution analysis .............................................47

3.4.2 H2O2 concentration analysis ............................................................................49

3.4.3 Adsorbent extraction analysis ..........................................................................51

3.5 Adsorption experiments ...........................................................................................51

3.5.1 Adsorption kinetics ..........................................................................................51

3.5.2 Adsorption isotherms.......................................................................................52

vii
3.6 Regeneration experiments........................................................................................53

3.6.1 Fenton oxidation regeneration .........................................................................53

3.6.2 EF regeneration using graphite felt cathode ....................................................54

3.6.3 EF regeneration-cathode as adsorbent .............................................................56

3.7 Experimental error analysis .....................................................................................56

CHAPTER FOUR: IMPACT OF PORE SIZE ON FENTON OXIDATION OF METHYL


ORANGE ADSORBED ON MAGNETIC CARBON MATERIALS: TRADE-OFF
BETWEEN CAPACITY AND REGENERABILITY .............................................59

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

4.2 Materials and methods .............................................................................................62

4.2.1 Materials ..........................................................................................................62

4.2.2 Preparation and characterization of magnetic samples ...................................62

4.2.3 Adsorption isotherms.......................................................................................63

4.2.4 Adsorption and Fenton oxidation regeneration ...............................................64

4.3 Results and discussion .............................................................................................65

4.3.1 Characterization of samples ............................................................................65

4.3.2 Adsorption kinetics ..........................................................................................70

4.3.3 Adsorption isotherms.......................................................................................72

4.3.4 Effect of loading Fe .........................................................................................75

4.3.5 Effect of adsorption and regeneration conditions ............................................76

4.3.6 Effect of particle size .......................................................................................81

4.3.7 Multiple adsorption-regeneration cycles .........................................................83

4.3.8 Relationship between regeneration efficiency and pore structure ...................84

4.4 Conclusions ..............................................................................................................88

CHAPTER FIVE: MECHANISTIC INSIGHTS FOR THE ELECTRO-FENTON


REGENERATION OF METHYL ORANGE SATURATED CARBON MATERIALS
...................................................................................................................................89
viii
5.1 Introduction ..............................................................................................................89

5.2 Materials and methods .............................................................................................92

5.2.1 Electrodes ........................................................................................................92

5.2.2 H2O2 generation ...............................................................................................93

5.2.3 Carbon material preparation, adsorption, and Electro-Fenton regeneration ...93

5.2.4 Characterization of carbon materials ...............................................................95

5.3 Results and discussion .............................................................................................95

5.3.1 Effect of cathodic potential .............................................................................95

5.3.2 Effect of carbon pore structure-comparison with Fenton oxidation regeneration


........................................................................................................................102

5.3.3 Effect of electrodes ........................................................................................107

5.3.4 Mineralization of MO ....................................................................................111

5.3.5 Regeneration and mineralization mechanisms ..............................................113

5.4 Conclusions ............................................................................................................120

CHAPTER SIX: INVESTIGATION OF THE IMPACT OF HYDROPHILICITY ON


ACTIVATED CARBON-POLYTETRAFLUOROETHYLENE ELECTRODE
ADSORPTION AND ELECTRO-FENTON REGENERATION..........................121

6.1 Introduction ............................................................................................................122

6.2 Materials and methods ...........................................................................................124

6.2.1 Preparation of AC-PTFE electrodes ..............................................................124

6.2.2 Characterization of AC-PTFE electrodes ......................................................124

6.2.3 Adsorption and regeneration .........................................................................125

6.2.4 Regeneration cycles .......................................................................................126

6.3 Results and discussion ...........................................................................................127

6.3.1 Characterization of AC-PTFE electrodes ......................................................127

6.3.2 Hydrogen peroxide generation ......................................................................131

ix
6.3.3 Electro-Fenton oxidation of MO ...................................................................132

6.3.4 MO adsorption before and after electro-Fenton regeneration .......................135

6.3.5 Effect of adsorption time ...............................................................................140

6.3.6 Regeneration cycles .......................................................................................142

6.4 Conclusions ............................................................................................................147

CHAPTER SEVEN: CONCLUSIONS AND RECOMMENDATIONS ........................149

7.1 Conclusions ............................................................................................................149

7.2 Recommendations for future research ...................................................................152

REFERENCES ................................................................................................................155

APPENDIX A: PRELIMINARY RESULTS-REGENERATION OF IRON AMENDED


CARBON BY HETEROGENEOUS FENTON OXIDATION ..............................180

A.1. Introduction ..........................................................................................................180

A.2. Materials and methods .........................................................................................180

A.2.1. Materials ......................................................................................................180

A.2.2. Carbon modification ....................................................................................180

A.2.3. Adsorption ...................................................................................................181

A.2.4. Regeneration ................................................................................................181

A.3. Results and discussion .........................................................................................182

A.3.1. Adsorption isotherms of modified carbon materials ...................................182

A.3.2. Effect of regeneration conditions ................................................................183

A.3.3. Effect of carbon modification ......................................................................187

A.4. Conclusions ..........................................................................................................188

A.5. Problems with G5 powder activated carbon ........................................................189

APPENDIX B: ERRONEOUS APPLICATION OF PSEUDO SECOND ORDER


ADSORPTION KINETICS MODEL: IGNORED ASSUMPTIONS AND SPURIOUS
CORRELATIONS ..................................................................................................190

x
B.1. Introduction ..........................................................................................................190

B.2. Results and discussion..........................................................................................191

B.2.1. Pseudo second order model development ....................................................191

B.2.2. Linearization of the model and subsequent model fitting............................194

B.2.3. Recommendations for fitting adsorption data ..............................................198

B.3. Conclusions ..........................................................................................................201

APPENDIX C: ELECTRO-FENTON WATER TREATMENT RESULTS IN THE


LITERATURE ........................................................................................................203

APPENDIX D: METHYL ORANGE ADSORPTION MECHANISM PROPOSED IN THE


LITERATURE ........................................................................................................208

APPENDIX E: COMPARISON OF NITROGEN ADSORPTION ISOTHERM AND


CARBON DIOXIDE ADSORPTION ISOTHERM DATA ..................................210

APPENDIX F: FENTON OXIDATION OF METHYL ORANGE ................................212

APPENDIX G: COPYRIGHT PERMISSIONS ..............................................................214

xi
List of Tables

Table 3-1. The MO adsorption capacity of Fe2O3/G5 tested at different time. ............................ 58

Table 4-1. Physical properties and metal oxide loadings of the samples. .................................... 67

Table 4-2. WM diffusion model fits for the adsorption kinetics of the fresh and regenerated
magnetic samples. ................................................................................................................. 72

Table 4-3. The Freundlich isotherm and Langmuir isotherm parameters of MO adsorption for
different magnetic samples. .................................................................................................. 74

Table 4-4. Effect of iron loading on the adsorption capacity of fresh and regenerated carbon
samples and magnetic samples. ............................................................................................ 76

Table 5-1. The influence of applied charge on regeneration efficiency under different EF
conditions. ........................................................................................................................... 100

Table 5-2. Physical properties of adsorbents as determined by nitrogen adsorption.................. 102

Table 5-3. Effect of cathode type on EF regeneration of Fe2O3/G5 saturated with MO. ........... 108

Table 5-4. Effect of anodes and air flow on MO adsorption capacity of regenerated carbon
and regeneration efficiency. ................................................................................................ 111

Table 6-1. The pore structure properties of the G5-PTFE electrode materials. .......................... 128

Table 6-2. Pseudo first-order rate constant of MO removal by EF using different G5-PTFE
cathodes............................................................................................................................... 133

Table 6-3. The applied charge for the regeneration of G5-PTFE-7-1 for each cycle when
fresh electrolyte was supplied for each cycle...................................................................... 143

Table 6-4. The applied charge for the regeneration of G5-PTFE-7-1 for each cycle when the
electrolyte was reused. ........................................................................................................ 147

xii
List of Figures

Figure 1-1. Schemes of (a) AOP application for wastewater treatment directly and (b) AOP
application for adsorbent regeneration. ................................................................................... 3

Figure 1-2. Mechanism of the EF process with a focus on the cathodic reactions. ........................ 5

Figure 1-3. Scheme of adsorption-regeneration cycle for water treatment. ................................... 6

Figure 2-1. Schematic relationship between solution pH and adsorbent surface charge. ............. 18

Figure 2-2. The impact of relationships among water pH, adsorbate pKa, and adsorbent
pHpzc on the interaction during adsorption. ......................................................................... 19

Figure 2-3. The effect of pore size on the adsorption potential energy. ....................................... 20

Figure 2-4. Comparison between different decomposition based regeneration methods. ............ 25

Figure 2-5. Scheme of Fenton oxidation regeneration mechanism. ............................................. 29

Figure 2-6. Two possible desorption processes on a cathodic polarized surface. ........................ 38

Figure 3-1. Experimental strategy used in this dissertation. ......................................................... 41

Figure 3-2. Molecular structure of MO......................................................................................... 42

Figure 3-3. Molecular structures of MO anion under acid and base solution. .............................. 42

Figure 3-4. Procedure for the preparation of magnetic (γ-Fe2O3) carbon materials. .................... 43

Figure 3-5. Procedure for the preparation of carbon-PTFE electrodes. ........................................ 44

Figure 3-6. The UV-Vis spectra of an aqueous MO solution at pH 3 and pH 7........................... 48

Figure 3-7. The calibration curves for MO solutions at pH 3 and pH 7. ...................................... 48

Figure 3-8. The UV-Vis spectra of solutions with H2O2 (~150 mg/L) and TiOSO4 (0.25 mL)
as well as deionized water and TiOSO4 (0.25 mL). .............................................................. 50

Figure 3-9. The calibration curve for H2O2 concentration determined using TiOSO4 as an
indicator. ............................................................................................................................... 50

Figure 3-10. Experimental flow diagram of Fenton oxidation regeneration. ............................... 53

Figure 3-11. Potential overestimation of regeneration efficiency under some adsorption


conditions. ............................................................................................................................. 54

Figure 3-12. Schematic diagram of EF regeneration system. ....................................................... 55


xiii
Figure 3-13. The cable connections of Solartron SI1287 potentiostate for a three-electrode
electrochemical cell shown in Figure 3-12. .......................................................................... 56

Figure 3-14. The UV-Vis absorbance of the MO standard solutions at 464 nm prepared at
different time. ........................................................................................................................ 57

Figure 4-1. The H2O2 concentration (■) during regeneration of Fe2O3/NORIT. ........................ 65

Figure 4-2. XRD patterns of prepared samples. ........................................................................... 66

Figure 4-3. Pore size distributions of SMC and Fe2O3/SMC. ...................................................... 68

Figure 4-4. TGA of samples before and after iron oxide loading and regeneration. .................... 69

Figure 4-5. Pore size distributions for fresh (○) and regenerated (●) magnetic samples: (a)
Fe2O3/SMC, (b) Fe2O3/G5, and (c) Fe2O3/NORIT. .............................................................. 70

Figure 4-6. Rate of MO adsorption at 20 °C for fresh (open symbols) and regenerated (filled
symbols) magnetic samples: Fe2O3/SMC (∆, ▲), Fe2O3/G5 (□, ■), and Fe2O3/NORIT
(○, ●). ................................................................................................................................. 71

Figure 4-7. MO adsorption isotherms on magnetic samples (a) as prepared and (b) after
regeneration: Fe2O3/NORIT (○, ●), Fe2O3/G5 (□, ■) and Fe2O3/SMC (∆, ▲). ............. 73

Figure 4-8. Impact of regeneration conditions on Fe2O3/SMC adsorption capacity. ................... 77

Figure 4-9. Desorption kinetics of MO saturated carbon materials. ............................................. 78

Figure 4-10. Impact of time, H2O2 concentration, and H2O2 amount during regeneration of
Fe2O3/G5. .............................................................................................................................. 79

Figure 4-11. Regeneration kinetics of Fe2O3/G5 with 250 rpm shaking (○) and without
shaking (●). ......................................................................................................................... 80

Figure 4-12. Impact of initial MO concentration on adsorption capacity of Fe2O3/NORIT. ....... 81

Figure 4-13. MO adsorption kinetics of fresh and regenerated Fe2O3/NORIT with different
carbon support sizes. ............................................................................................................. 82

Figure 4-14. The MO adsorption capacity of (a) Fe2O3/SMC and (b) Fe2O3/G5 during 6
consecutive adsorption-regeneration cycles.......................................................................... 84

Figure 4-15. Relationship between regeneration efficiency and pore structure: (a)
micoporosity; (b)-(d) intraparticle diffusion rate constants. ................................................. 85

Figure 4-16. Schematic of MO adsorption in two pores of different diameters, and


subsequent reaction with H2O2. ............................................................................................ 87

xiv
Figure 5-1. H2O2 concentration as a function of time generated with CB-PTFE (carbon black-
PTFE) electrode and GF (graphite felt) at different cathodic potential. ............................... 96

Figure 5-2. Cyclic voltammetry (CV) curve of graphite felt. ....................................................... 97

Figure 5-3. MO removal kinetics by EF oxidation process. ......................................................... 97

Figure 5-4. Current efficiency of H2O2 electrochemical generation by graphite felt at


different cathodic potentials. ................................................................................................. 98

Figure 5-5. Effect of cathodic potential on the adsorption capacity of regenerated Fe2O3/G5
and regeneration efficiency. .................................................................................................. 99

Figure 5-6. The potentiostatic polarization curves of graphite felt during H2O2 generation and
regeneration at different cathodic potentials (vs Ag/AgCl). ............................................... 100

Figure 5-7. The adsorption capacity (■) of regenerated Fe2O3/G5 and applied charge (▲)
during EF regeneration at different time. ............................................................................ 101

Figure 5-8. 2D-NLDFT pore size distributions of fresh and regenerated Fe2O3/G5. ................. 104

Figure 5-9. Regeneration efficiency as a function of adsorbent microporosity at different


applied potentials for the Fenton and EF (-0.8 V and -3 V) oxidation of carbon
adsorbents saturated with MO............................................................................................. 105

Figure 5-10. Adsorption capacity and regeneration efficiency (■) of different adsorbates on
different carbon materials. .................................................................................................. 106

Figure 5-11. The relationship between regeneration efficiency and the recovery of carbon
pore structure properties...................................................................................................... 107

Figure 5-12. Reactor pictures before (a) and after (b) EF regeneration using graphite felt as
cathode. ............................................................................................................................... 109

Figure 5-13. Pictures of graphite felt before (a) and after (b) regeneration. ............................... 109

Figure 5-14. The UV-vis spectra of the electrolyte solution during the EF regeneration of
Fe2O3/NORIT. ..................................................................................................................... 112

Figure 5-15. The UV-visible spectra of methanol solution after extraction of EF regenerated
carbon. ................................................................................................................................. 113

Figure 5-16. The fraction of BA and benzoate at different pH. .................................................. 117

Figure 5-17. Relationship between H2O2 generation rate (derived from the initial 20 min) and
regeneration efficiency of Fe2O3/G5 after EF regeneration. ............................................... 119

Figure 5-18. Proposed scheme for the EF regeneration of carbon materials. ............................. 120
xv
Figure 6-1. The nitrogen adsorption and desorption isotherms (a) and pore size distribution
(b) of G5-PTFE electrode materials. ................................................................................... 127

Figure 6-2. Measurement of contact angle with deionized water for electrodes with different
G5 to PTFE ratios. .............................................................................................................. 129

Figure 6-3. The SEM and EDX images of fresh G5-PTFE-1-1 and G5-PTFE-7-1. .................. 130

Figure 6-4. The FTIR spectrum of G5-PTFE electrode materials. ............................................. 131

Figure 6-5. The H2O2 production kinetics (a) of different AC-PTFE electrodes at applied
cathodic potential of -0.8 V (vs Ag/AgCl) and the relationship between H2O2
concentration at 2 h and hydrophilicity (represented by cosθ). .......................................... 132

Figure 6-6. EF oxidation kinetics of MO with different G5-PTFE cathodes. ............................ 133

Figure 6-7. The UV-Vis spectra of electrolyte during MO EF oxidation using different G5-
PTFE cathodes. ................................................................................................................... 135

Figure 6-8. MO adsorption kinetics of fresh (a) and regenerated (b) G5-PTFE electrodes. ...... 136

Figure 6-9. The relationship between WM diffusion rate constant and G5-PTFE electrodes
hydrophilicity (represented by cosθ)................................................................................... 138

Figure 6-10. The relationship between EF regeneration efficiency and G5-PTFE electrode
hydrophilicity. ..................................................................................................................... 139

Figure 6-11. MO adsorption kinetics G5-PTFE electrodes for different adsorption times. ....... 140

Figure 6-12. MO adsorption kinetics of G5-PTFE-7-1 in different adsorption-regeneration


cycles. .................................................................................................................................. 142

Figure 6-13. Adsorption-regeneration cycles of G5-PTFE-7-1 without changing electrolyte. .. 144

Figure 6-14. Deposition of iron species (red) on the G5-PTFE-7-1 cathode after the sixth (a)
and eighth (b) regeneration cycles. ..................................................................................... 145

Figure 6-15. The UV-Vis spectra of the electrolyte after each regeneration of G5-PTFE-7-1
when the electrolyte was reused in each cycle. ................................................................... 146

xvi
List of Abbreviations

Abbreviations Definition
2D-NLDFT Two-dimensional nonlocal density functional theory
ATR Attenuated total reflection
AC Activated carbon
AOP Advanced oxidation process
BA Benzoic acid
BDD Boron doped diamond
BET Brunauer–Emmett–Teller
CB Carbon black
CNT Carbon nanotube
DFT Density functional theory
EDX Energy-dispersive X-ray spectroscopy
EF Electro-Fenton
FTIR Fourier-transform infrared spectroscopy
GAC Granular activated carbon
GDE Gas diffusion electrode
GO Graphene oxide
H2O2 Hydrogen peroxide
IUPAC International Union of Pure and Applied Chemistry
MB Methylene blue
MO Methyl orange
MTBE Methyl-tert-butyl ether
NOM Natural organic matter
•OH Hydroxyl radical
ORR Oxygen reduction reaction
PAC Powdered activated carbon
PTFE Polytetrafluoroethylene
PZC Point of zero charge
SEM Scanning electron microscope
TGA Thermogravimetric analysis
TOC Total organic carbon
UV Ultraviolet light
UV-Vis Ultraviolet-visible spectroscopy
WM Weber-Morris
e charge of electron
Ka Acidity constant
NA Avogadro's number

xvii
Chapter One: Introduction

With the rapid industrialization of the global society, more and more organic chemicals are

being synthesized and used in our lives, which will inevitably be released into the environment

through human activities. Specifically, the ubiquitous presence of various organic pollutants in

surface water, groundwater, and wastewater has drawn tremendous attention from the public and

scientists.

In order to abate the influence of these organic pollutants, numerous technologies have been

developed to treat water contaminated by toxic organic pollutants, such as Fenton oxidation [1],

electrochemical methods [2, 3], photocatalysis [4, 5], ultrasonic processes [6], and adsorption [7-

9]. Among these technologies, adsorption processes have the advantages of low capital

investment, simple design, simple installation, ease of operation, insensitivity to toxic substrates,

and high efficiency (up to ~100%) for contaminants removal, even for those at low

concentrations (such as < 1 mg/L) [10, 11]. Among all the adsorbents, activated carbon (AC) has

been intensively investigated because of its high surface area, well developed internal porosity,

and presence of various functional groups [12]. After use, it is desirable to regenerate and recycle

the used AC, rather than to burn or dispose of it in a landfill because of the high cost and

potential of secondary pollution. For instance, the estimated annual AC cost for a wastewater

treatment plant using adsorption process with average daily flow of 100,000 m3/d would be over

10 million USD [13-15].

Numerous studies have been carried out for the regeneration of AC saturated with organic

pollutants, see [13, 16] and references therein. Among all the regeneration technologies

investigated, thermal regeneration method is the primary commercialized technology for

1
granular activated carbon (GAC) regeneration in a wastewater treatment plant [14]. However,

thermal regeneration technology has some disadvantages [17], which include (1) decrease in the

ability of removing small molecular contaminants owing to the enlargement of micropores, (2)

~5-20% loss in carbon weight, and (3) high energy requirements. To avoid these drawbacks,

researchers have investigated several alternative methods such as extraction and regeneration by

microwave, microbiology, chemical oxidation, and electrochemical oxidation. According to the

fate of the adsorbed organic pollutants, regeneration technologies can be classified into

desorption-based technologies and decomposition-based technologies [18]. To eliminate the

potential for secondary pollution, decomposition-based technologies are used to completely

mineralize the adsorbed organic pollutants.

As a promising substitution for thermal regeneration, advanced oxidation process (AOP) have

drawn tremendous attention because of their capabilities to regenerate carbon materials under

ambient conditions. Broadly speaking, the AOP is a kind of chemical treatment method which

can generate highly active hydroxyl radicals (•OH) that can degrade and mineralize the organic

pollutants [19]. The AOP includes, but is not limited to, Fenton oxidation, electro-Fenton (EF)

oxidation, anodic oxidation, ozonation, and photocatalysis [19, 20]. The AOP can be directly

applied for the treatment of wastewater, but large amount of chemicals is required to adjust

conditions for AOP and neutralize the treated water for discharge (Figure 1-1a) [1]. On the

contrary, much smaller amount of chemicals may be needed if the AOP is used for the

regeneration of adsorbents because the organic pollutants are accumulated and concentrated

during the adsorption process (Figure 1-1b).

2
(a)

(b)
Figure 1-1. Schemes of (a) AOP application for wastewater treatment directly and (b) AOP

application for adsorbent regeneration.

Some of the AOP has been applied in the regeneration of carbon materials, with a high

regeneration efficiency (> 90%) reported [13, 21-23]. However, as indicated in many

investigations [17, 24, 25], the AOP regeneration methods still have some problems for practical

applications, but they merit further investigation. In this dissertation, Fenton oxidation and EF

oxidation regeneration of saturated carbon materials were investigated.

Fenton oxidation is an oxidation method that utilizes the •OH, which is generated from

Fenton reagents, i.e., hydrogen peroxide (H2O2) and ferrous iron (Fe2+), through the following

reaction (1-1).
3
𝐻2 𝑂2 + 𝐹𝑒 2+ + 𝐻 + →• 𝑂𝐻 + 𝐻2 𝑂 + 𝐹𝑒 3+ (1-1)

The •OH radical then reacts with contaminates producing various decomposition products. The

iron reagent is regenerated to complete the catalytic cycle by the following reaction (1-2).

𝐹𝑒 3+ + 𝐻2 𝑂2 → 𝐹𝑒 2+ + 𝐻𝑂2• + 𝐻 + (1-2)

Although Fenton oxidation is capable of regenerating mesoporous (pore size between 2-50

nm) carbon materials [26], the Fenton oxidation regeneration can partially (mostly <~60%)

regenerated ACs with a high microporosity (pore size < 2 nm) [25, 27, 28]. In view of this, the

impact of pore structure of carbon materials on Fenton oxidation regeneration was studied in this

thesis. The pore structure of carbon materials can affect the adsorption kinetics and isotherms of

an adsorption process by limiting diffusion of adsorbates and increasing the interaction energy

between adsorbates and adsorbents. As a consequence, the pore structure can influence

desorption of the adsorbed adsorbates, and affect the reaction rate between the adsorbates and

regeneration reagents because of the limited accessibility of some reaction sites.

As an attempt to improve the regeneration efficiency of microporous materials, the EF

oxidation method was studied and compared to Fenton oxidation regeneration. EF oxidation is a

Fenton oxidation-based technology, which is typically applied in water treatment as illustrated in

Figure 1-2.

4
Figure 1-2. Mechanism of the EF process with a focus on the cathodic reactions.

In this process, H2O2 is electrochemically produced rather than externally added, and the Fe

catalyst is electrochemical regenerated. EF regeneration was reported to achieve almost complete

regeneration of carbon materials [23]. The regeneration efficiency, however, may be

overestimated because the carbon materials were not saturated before regeneration [23]. Besides,

the mechanism and pathways of EF regeneration were not clear owing to scant information about

the impact of different parameters. Different factors, including cathodic potential, cathodic type,

carbon pore structure, catalyst loading, anode type, and air flow can have significant influence on

the EF oxidation treatment of wastewater [29]. In contrast, the impact of these factors on EF

regeneration was not investigated. Therefore, it is difficult to understand the behavior of

adsorbates and the mechanism during EF regeneration. Although oxidation was proposed to be

the main pathway for EF regeneration [23], the fact that cathodic regeneration, which was similar

to EF regeneration except no oxygen and Fe catalyst was introduced, could regenerate phenol
5
saturated carbon materials [30] indicated that electro-desorption may also contribute to the EF

regeneration process.

Meanwhile, if the adsorption and Fenton or EF oxidation regeneration cycles are considered

for removal of organic pollutants in a water flow, it is beneficial to separate the regeneration

process from the adsorption as shown in Figure 1-3.

Figure 1-3. Scheme of adsorption-regeneration cycle for water treatment.

During the Fenton or EF oxidation regeneration process, the aqueous solution needs to be

adjusted to the optimal conditions (e.g. pH of 3.0, presence of iron catalyst, and addition of

electrolyte). If regeneration is carried out in the same reactor as the adsorption process,

additional chemicals and treatments are required to meet the emission standards. One way to

facilitate the separation of carbon materials from water is to load magnetic iron oxides (such as

γ-Fe2O3 and Fe3O4) [31] on the carbon surface, and the loaded iron oxides can also act as

catalysts for the Fenton reaction. It would be advantageous if the carbonaceous electrodes

applied in EF regeneration can also act as the adsorbents for adsorption. In this case, the
6
separation procedure could be eliminated. During the EF oxidation process, carbon materials

such as graphite felt, carbon-polytetrafluoroethylene (PTFE) material, and reticulated vitreous

carbon were often used as the cathodes [29]. These cathodes often have low adsorption

capacities for organic pollutants because of their low surface area. Thus, further research on

combination of adsorbents and electrodes is required.

Research has been carried out to investigate the application of a carbon nanotube (CNT)-

PTFE electrode as both the adsorbent for adsorption and as the cathode for regeneration in an

electro-peroxone process [32]. Although high regeneration efficiency (~100%) of the electrode

and high mineralization (~100%) of adsorbed pollutant have been reported, the adsorption

capacity was low at ~1.0 mg/g for this CNTs-PTFE electrode [32]. Therefore, it is necessary to

prepare electrodes with a high surface area and thus high adsorption capacity for the adsorption

and regeneration process. AC-PTFE cathodes with high surface area have been applied for

oxygen reduction reactions (ORRs) [33], but they have not been used for adsorption and EF

regeneration. The introduction of PTFE into the electrode can reduce the hydrophilicity of the

electrode [33] and thus may affect the adsorption and regeneration processes. Therefore, it is

important to determine the proper hydrophilicity that an electrode should have for its application

in an adsorption and regeneration process.

This thesis describes the regeneration of carbon materials by Fenton and EF oxidation

methods and contains 7 chapters. In Chapter 2, a literature review on the adsorption of carbon

materials and regeneration methods with a focus on regeneration by AOP was carried out to

identify the knowledge gaps. In Chapter 3, the methodology for this research and the materials

7
and methods used in this study are listed and explained. In addition, the sources of experimental

errors are also summarized in Chapter 3.

In Chapter 4, the regeneration of carbon materials by the Fenton oxidation method was

investigated with a focus on the important impact of pore structure properties on the regeneration

of carbon materials saturated with methyl orange (MO). Most of the results in this chapter were

published in the journal of Environmental Science & Technology (Xiao, Y., & Hill, J. M. (2017).

Environmental Science & Technology, 51(8), 4567-4575).

In Chapter 5, by using a graphite felt cathode, the EF oxidation method was applied to

regenerate MO saturated carbon materials, focusing on the influence of different parameters on

regeneration. Most of the results in this chapter have been included in the manuscript submitted

to the journal of Water Research.

In Chapter 6, by using AC-PTFE electrodes as both the adsorbent for adsorption and the

cathode for EF regeneration, the effect of hydrophilicity on adsorption and regeneration has been

investigated. Most of the results in this chapter have been included in a manuscript in preparation

for submission to Environmental Science & Technology.

In Chapter 7, the main conclusions of this dissertation have been listed and summarized. In

addition, suggestions on AOP regeneration methods have been made.

In addition, some appendices are included at the end of this thesis. In Appendix A,

preliminary results on Fenton oxidation regeneration of iron loaded carbon materials saturated

with MO are presented. In Appendix B, the erroneous application of pseudo-second-order

adsorption kinetics model with a focus on the ignored assumptions and spurious correlations is

8
discussed. The content in Appendix B was published in the journal of “Industrial & Engineering

Chemistry Research” (Xiao, Y., Azaiez, J., & Hill, J. M. (2018). Industrial & Engineering

Chemistry Research, 57(7), 2705-2709). In Appendix C, some studies on EF oxidation of

organic pollutants are summarized. In Appendix D, the MO adsorption mechanisms reported in

the literature are summarized. In Appendix E, carbon dioxide adsorption isotherms of some

carbon materials are included and compared with nitrogen adsorption isotherms. In Appendix F,

the results for Fenton oxidation of MO by three different carbon materials are presented. The last

Appendix F includes the copyright permissions.

9
Chapter Two: Literature review

2.1 Organic pollutants in water system

Clean water is a crucial resource for ecosystems and humans, but is limited in quantity and in

quality [34]. Organic pollutants present in all kinds of water systems, and are reported to be one

of the major concerns for water safety. With the wide application of organic chemicals in our

daily life, numerous emerging organic pollutants are released into water systems through

domestic wastewater, industrial wastewater, and agriculture activities. These organic pollutants

include, but are not limited to pesticides, surfactants, synthetic dyes, pharmaceuticals, personal

care products, and flammable retardants. For instance, naphthenic acids are considered the main

toxic contaminants in oil sands process-affected water in Alberta, Canada [35]. Perfluorooctane

sulfonate was detected at the level of 10-1000 pg/L in the ocean water [36]. Pharmaceuticals and

hormones were detected in the effluent of most wastewater treatment plants with concentrations

up to tens of μg/L [37, 38]. Although most of the emerging pollutants are at low concentrations,

a large amount of these organic pollutants is released into the environment everyday considering

that the treatment capacity of a wastewater treatment plant can be larger than 1 million m3/d.

These organic pollutants, especially synthetic chemicals, can be toxic or cause reproductive,

neurologic, endocrine, and immunologic adverse health effects on wild species [39]. For

instance, naphthenic acids solutions at concentrations in tens of mg/L had toxic and inhibitory

effects on different organisms such as plants, fish, and bacteria [40]. Pharmaceuticals and

hormones at low concentration levels (< 1 μg/L) can have chronic effects on the aquatic biota

and increase the possibility that bacteria develop resistance [38]. In addition, some of these

chemicals can remain in the environment for a long time with a half-life ranging from years to

>20 years [41]. The accumulation of these chemicals through food-chain can increase the

10
exposure to humans. For example, perfluorooctane sulfonate concentrations as high as 4000 ng/g

were detected in some marine mammal species and polar bears [36, 42]. Therefore, it is

important to develop reliable technologies to remove organic pollutants from water systems.

Numerous technologies have been developed and investigated for the treatment of water

contaminated by organic pollutants, which can be classified as biological, chemical, and physical

methods. Among these, biological methods may have low efficiencies and slow rates for removal

of recalcitrant organic chemicals and micro-pollutants from wastewater [43], for example only

<30% of organic pollutants were removed from water in a 3 days biological treatment [44]. On

the other hand, chemical and physical methods often require high energy and/or chemical inputs

to remove or degrade the organic pollutants in water. Adsorption is one of the most feasible

technologies to use in practice because it can remove a wide range of organic chemicals even at

low concentrations with high efficiency.

2.2 Adsorption of organic pollutants

2.2.1 Adsorption system

According to the International Union of Pure and Applied Chemistry (IUPAC) definition,

adsorption is “an increase in the concentration of a dissolved substance at the interface of a

condensed and a liquid phase due to the operation of surface forces” [45]. In this definition, the

dissolved substance is the adsorbate while the condensed phase is the adsorbent. Generally,

adsorption isotherms and kinetics analysis are collected to characterize an adsorption system

with specific adsorbates and adsorbents.

An adsorption isotherm demonstrates the relationship between the adsorption capacity and

adsorbate concentration at equilibrium and constant temperature. The two major adsorption

11
isotherm models - Langmuir and Freundlich - are often applied for the adsorption of organic

pollutants from water [10, 46, 47]. These models are given in equations (2-1) and (2-2),

𝑞𝑚𝑎𝑥 𝐾𝐿 𝐶𝑒
𝑞𝑒 = (2-1)
1 + 𝐾𝐿 𝐶𝑒

1/𝑛
𝑞𝑒 = 𝐾𝐹 𝐶𝑒 (2-2)

where 𝑞𝑒 is the equilibrium adsorption capacity, 𝑞𝑚𝑎𝑥 is the maximum monolayer adsorption

capacity, 𝐶𝑒 is the adsorbate concentration in water phase at equilibrium, 𝐾𝐿 is the Langmuir

constant, 𝐾𝐹 is the Freundlich constant, and 𝑛 is the heterogeneity. The Langmuir adsorption

isotherm model has assumptions of homogeneous and monolayer adsorption [46]. Although it

has been widely applied in different adsorption processes in water treatment, these assumptions

were not carefully checked in most of the investigations. The Freundlich adsorption isotherm

model is an empirical model with the basic assumption of heterogeneous adsorption [47], which

is the case in most adsorption systems [7].

Adsorption kinetics is the relationship between the amount of adsorbed adsorbate on an

adsorbent surface and the adsorption time. Based on the rate limiting step of an adsorption

process, the adsorption kinetics can be classified into two categories, which are reaction (the

reaction between the adsorbate and adsorbent) controlled and diffusion controlled kinetics

models [48]. The two widely used reaction controlled adsorption kinetics models are the pseudo-

first-order model developed by Largergren [49] and the pseudo-second-order model developed

by Ho and McKay [50] as described in equations (2-3) and (2-4).

𝑘1
log(𝑞𝑒 − 𝑞𝑡 ) = log 𝑞𝑒 − 𝑡 (2-3)
2.303

12
1 1
= + 𝑘2 𝑡 (2-4)
(𝑞𝑒 − 𝑞𝑡 ) 𝑞𝑒

where 𝑞𝑒 is the equilibrium adsorption capacity, 𝑞𝑡 is the adsorbed adsorbate at time 𝑡, 𝑘1 is the

pseudo-first-order rate constant, and 𝑘2 is the pseudo-second-order rate constant. The assumption

of reaction control was often ignored when these models were employed for analysis. In

particular, the application of the pseudo second order kinetics model is becoming more and more

popular due to the high correlation coefficients through a linear fitting, but it can result in

spurious correlations. A detailed analysis of the issues encountered with applying the pseudo

second order kinetics model is shown in Appendix B. Diffusion controlled kinetics models, on

the other hand, often have complicated formulae [48] and are difficult to solve. Weber-Morris

(WM) diffusion model [51] in equation (2-5),

𝑞𝑡 = 𝑘𝑊 𝑡1/2 + 𝐶 (2-5)

where 𝑘𝑊 is the WM diffusion rate constant, is an empirical diffusion controlled kinetics model

with a simple expression, which has been widely used for adsorption of contaminates from

water.

For a water system with specified organic pollutants, the adsorption process, including its

isotherm and kinetics, is mainly determined by the adsorbent. Numerous materials have been

investigated for the adsorption of organic pollutants from water, including carbon materials [7,

12, 15, 52-56], zeolites [54, 57], clays [58], and metal organic frameworks [59]. Among these

adsorbents, carbon materials have drawn intensive attention for organic pollutants removal since

they have high surface area (hundreds to several thousand m2/g), well developed internal

porosity, and various functional groups [12].

13
2.2.2 Activated carbon

AC materials are highly porous (> 0.7 cm3/g) amorphous structures [60]. AC is generally

prepared from materials with a high carbon content such as biomass, coal, and petcoke [61].

After pyrolysis and carbonization under no or low oxygen conditions at elevated temperatures,

the carbon materials undergo an activation process in the presence of oxidation reagents [60].

There are two kinds of activation methods, physical activation and chemical activation. During

physical activation, steam or carbon dioxide is applied to partially oxidize the carbon materials at

a high temperature (above 800 °C) to create pores in the carbon materials [55]. On the other

hand, for chemical activation, chemicals such as phosphoric acid, hydroxides, and zinc chloride

are mixed with carbon materials and react at high temperatures to increase the porosity [55]. On

some occasions, physical and chemical activation methods are also combined to increase the

porosity as well as the pore size of the obtained AC [62].

The physical properties of AC - surface area > 500 m2/g and porosity > 0.7 cm3/g [60]- make

it favorable for the adsorption of organic pollutants. Oxygen functional groups such as phenolic,

carboxylic, ketone, and quinone groups may be present on the carbon surface [63, 64] because of

the oxidation atmosphere during the activation process. The AC may contain some ash such as

silica, potassium oxide, iron oxides, and calcium oxides [65], which originate from the raw

materials. The presence of oxygen functional groups and ash provide hydrophilicity to the AC so

that they can be suspended in water instead of aggregating or adhering to the vessel wall [60].

The adsorption capacity of the AC for some organic pollutants can be improved in the presence

of these functional groups and ash components because of the enhanced interactions between

adsorbates and adsorbents under some conditions [66, 67].

14
In general, the adsorption capacity of AC increased with increased surface area and porosity.

For instance, an AC prepared from pine cones by phosphoric acid activation with a BET surface

area of 734 m2/g adsorbed up to 404 mg MO per gram [68], while AC from magnolia leaves by

potassium hydroxide activation with a BET surface area of 2834 m2/g (pore volume of 1.58

cm3/g) adsorbed 869 mg MO/g [69]. The adsorption capacity of an AC was less affected by a

change in pH compared with some other materials such as biochar and oxides. That is, the

adsorption capacity changed by less than 20% when the pH was changed from acidic to basic

[27, 69].

2.2.3 Biochar

Different definitions exist for biochar in the literature. According to the guidelines of the

European Biochar Certificate, biochar is a substance with abundant aromatic carbon and

minerals [70]. Biochar was also defined by Lehmann and Joseph as a carbon-rich product that

originated from biomass after heating, with little or no available air [71]. In general, biochar is

produced by biomass pyrolysis at high temperatures (350 °C to 800 °C) in a low or no oxygen

atmosphere [15, 70, 72].

Depending on the pyrolysis conditions and feedstock, the obtained biochar had significant

differences in physical and chemical properties [73]. In most cases, biochar had a low specific

surface area and pore volume as summarized by Ahmad et al., with a median BET surface area

of 15.0 m2/g and a median pore volume of 0.035 cm3/g [72]. For the chemical properties, biochar

generally has a high content of hydrogen (H/C up to 1.5, atomic ratio) and oxygen (O/C up to

0.9, atomic ratio) after pyrolysis at lower temperatures (200-400 °C) [71, 73]. Higher pyrolysis

temperatures (> 600 °C) can increase the carbon content through the dehydration and

15
deoxygenation reaction of biomass [72]. Overall, the adsorption capacity of biochar was low

(lower than 10 mg/g in most cases [15]), and could be highly affected by the pH of water [74].

2.2.4 Emerging carbon materials

Besides AC and biochar, numerous other carbon materials have also been developed and

synthesized for potential applications in adsorption. These carbon materials include carbon

nanotubes (CNT), ordered mesoporous carbon materials, graphene or graphene oxide, three-

dimensional graphene-based macrostructures, and hierarchically porous carbon [52, 75, 76].

Methods such as chemical oxidation and reduction, high temperature vapor deposition, and

templating are applied to synthesize these carbon materials.

These carbon materials generally have a high percentage of mesopores (up to 100%), which

would favor the fast adsorption of organic pollutants from water due to the lower diffusion

limitations. Lower BET surface areas of these carbon materials, however, could result in low

adsorption capacities. For example, a CNT prepared from chemical vapor deposition with a BET

surface area of 160 m2/g and average pore size of 20 nm was applied for MO adsorption [77].

The MO adsorption capacity was lower at ~50 mg/g compared with >400 mg/g for AC, but the

equilibrium was obtained in less than 3 h compared with 1 day for AC [77, 78]. The faster

adsorption kinetics of these engineering nanocarbon materials could also be attributed to the

open and interconnected three-dimensional pore structures [79]. However, it is also possible to

obtain high adsorption capacity for organic pollutants using some of these engineering carbon

nanomaterials. A graphene based three-dimensional structure reported by Wang et al. was

capable of adsorbing dyes from 115 mg/g for MO to 1260 mg/g for calcein [80]. Besides, it is

possible to increase the adsorption capacity of these materials by introducing functional groups

16
and microporosity [7]. The adsorption capacity of MO by a CNT was improved from 46 mg/g to

118 mg/g after introducing oxygen functional groups and increasing the surface area with

potassium hydroxide activation [81].

2.2.5 Mechanism of carbon adsorption

The adsorption mechanism can vary significantly depending on the properties of the

adsorbates, the carbon material properties, and the water chemistry [7, 82]. As interpreted in the

literature [7, 82], the mechanism for adsorption of organic pollutants onto carbon materials

included, but were not limited to, electrostatic interactions, hydrogen bonding, π-π interactions,

van der Waals interactions, dipole-dipole interactions, and dipole-ion interactions. Almost all

these interactions are intermolecular without formation and/or dissociation of chemical bonds so

that the adsorption of organic pollutants by carbon material is generally physical adsorption. The

pH of water influences electrostatic interactions but it has little effect on the other interactions.

Therefore, adsorption experiments at different pH were often conducted to determine the main

mechanism of the adsorption process [69, 81, 83-86].

The adsorption capacity can be influenced by the pH of water during adsorption if

electrostatic interactions are the main force for adsorption [87, 88]. For an adsorbent surface with

functional groups, it can be positively charged at low pH (reaction (2-6)) and negatively charged

at high pH (reaction (2-7)), as shown in Figure 2-1 [89].

≡ 𝑆 − 𝑂𝐻 + 𝐻 + ⇌≡ 𝑆 − 𝑂𝐻2+ (2-6)

≡ 𝑆 − 𝑂𝐻 + 𝑂𝐻 − ⇌≡ 𝑆 − 𝑂− + 𝐻2 𝑂 (2-7)

where ≡ 𝑆 stands for the adsorbent surface. When the surface charge of an adsorbent reaches

zero, this pH point is its point of zero charge (pzc).

17
Figure 2-1. Schematic relationship between solution pH and adsorbent surface charge.

Depending on the relationships of pH with the pKa (Ka is the acidity constant, and pKa = -

log10(Ka)) of the adsorbate and the pHpzc of the adsorbent, the interactions between the adsorbate

and the adsorbent can be either attractive or repulsive as shown in Figure 2-2 [89]. Therefore, the

adsorption capacity can be increased or decreased with the variation of pH for these adsorption

systems. The dependence of adsorption capacity on pH indicates that the electrostatic interaction

plays an important role in these adsorption processes. In some other adsorption systems,

however, the adsorption capacity was not significantly affected by the pH of water [69, 81],

indicating that other interactions such as a dispersion force instead of electrostatic interactions

could be the dominant adsorption mechanism.

18
(a) (b)

Figure 2-2. The impact of relationships among water pH, adsorbate pKa, and adsorbent pHpzc
on the interaction during adsorption. (a) anionic adsorbate; (b) cationic adsorbate.

The strength of different interactions can be interpreted by the potential energies. For the

electrostatic interaction, the potential energy can be estimated according to the zeta potential (𝜑)

of the adsorbent according to equation (2-8).

𝐸 = 𝜑 × 𝑒 × 𝑁𝐴 = 96.5𝜑 kJ/mol (2-8)

where 𝜑 is the zeta potential (V), 𝑒 is the charge of one electron (1.602 × 10-19 C), 𝑁𝐴 is

Avogadro’s number (6.022 × 1023 /mol). In the literature, the zeta potential of an adsorbent

normally ranges from -100 mV to 100 mV, so the potential energy of electrostatic interaction is

generally less than 10 kJ/mol, and can change significantly depending on the water pH. For the

van de Waals interactions and/or hydrogen bonding, the potential energy can vary from ~2.0

to >40 kJ/mol.

The potential energy during adsorption can be affected by the pore size of the carbon

adsorbent as shown in Figure 2-3. In big pores (pore width is larger than several times of

molecular size), the potential energy curve is well developed so that the interaction between the
19
carbon surface and the adsorbate is similar to that on the external surface. In narrow pores (pore

width similar to 1-2 times of molecular size), however, the potential energy can reach twice the

magnitude compared with the external surface.

Figure 2-3. The effect of pore size on the adsorption potential energy.

2.2.6 Challenge of carbon adsorption and strategies

With their possible high adsorption capacity and fast organic pollutants uptake rate, novel

carbon nanomaterials are attractive for the adsorption process. However, some of these carbon

materials are too expensive to be used in a wastewater treatment project. For instance, the

average price of single walled CNTs was in the range of $45,000-$140,000/kg [90], which

hinders their applications in water treatment.

Researchers have investigated two different strategies to reduce the cost of carbon materials

for adsorption. The first strategy is the application of low-cost adsorbents, such as biochar or

biomass directly, and the other is the regeneration and recycling of the spent adsorbents after

20
saturation [91]. Although the low cost of ~$0.250/kg [15] for biochar and the abundant resource

of biomass makes the low-cost adsorbent an economically-attractive approach for water

treatment, the main concern for this method is the potential of releasing adsorbed pollutants into

the environment if the adsorbents are not properly disposed [91]. Carbon regeneration also has

high economic benefits. For example, the cost for thermal regeneration of AC was only 50% of

the cost for fresh AC, and electrochemical regeneration was estimated to reduce the cost by more

than 95% [14, 92]. In addition, some of the regeneration technologies were able to degrade and

mineralize the adsorbed organic pollutants [18]. Therefore, carbon regeneration is a more

economically attractive and environmentally-benign approach to make the adsorption process

more applicable and feasible in water treatment.

2.3 Regeneration

2.3.1 Principle and classification

To regenerate carbon materials, the adsorbed organic pollutants must be removed through

desorption and/or degradation [18]. The desorption based regeneration process includes methods

such as solvent extraction, pH variation, temperature shifting, and pressure swing [13]. The main

principle of these methods is to reduce the interaction energy between the carbon surface and

adsorbate or to increase the interaction energy between the solvent and adsorbate. A feasible

desorption based regeneration method could be proposed based on the adsorption mechanism. If

the electrostatic interaction (attractive) is the major force for adsorption, then a pH change may

be effective to regenerate most of the adsorption sites on the adsorbent. On the other hand, if a

weak dispersion force is the dominant interaction between the adsorbed organic pollutants and

the adsorbent, then application of a solvent with high affinity to the adsorbed organic pollutants

would be effective to regenerate the adsorption sites. However, the desorption based

21
technologies just solve the problem temporarily, and further treatment of the other phase is

required to prevent secondary pollution or to recover the solvent and/or the adsorbate.

The best scenario for a regeneration process is to recover the adsorbent surface and to degrade

the adsorbed pollutants simultaneously [18], which is possible to achieve by decomposition

based regeneration methods. Some of these regeneration methods are described in following

section. The descriptions will focus on the use of these methods for carbon-based adsorbents.

2.3.2 Regeneration methods

Thermal regeneration is the only commercial technology that has been used in a water

treatment plant. During the thermal regeneration process at an industrial scale, the following

procedures are used: steaming, pyrolysis under inert atmosphere, and gasification and

reactivation by oxidizing gases such as steam or carbon dioxide [93].

Wet oxidation process, which was developed as a substitution of thermal regeneration [94],

was normally carried out under subcritical conditions at elevated temperature (200-320 °C) and

pressure (2-20 MPa) with an initial oxygen partial pressure ranging 10-50 bar [94, 95].

Advanced oxidation processes (AOP) are a set of chemical methods which can oxidize and

mineralize organic pollutants in the presence of •OH [19]. Because of the high standard

reduction potential of •OH in acidic aqueous solution, which is 2.80 V (vs. Standard Hydrogen

Electrode) higher than most other oxidants, it can oxidize almost all organic pollutants to carbon

dioxide and water except some of the simplest organic compounds [96]. There are numerous

AOP that have been developed for the water purification [96], and some of these technologies

have also been utilized in carbon regeneration either directly for AC or indirectly by modifying

carbon composites [13, 21, 97, 98].


22
Electrochemical regeneration occurs in an electrochemical cell contained the adsorbent, an

electrolyte, anode and cathode [13]. High regeneration efficiencies up to 95% or even an increase

in adsorption capacity was observed during electrochemical regeneration processes [30, 99].

Sometimes, electrochemical regeneration methods are classified as AOP. Technologies such as

EF, anodic oxidation, and electro-peroxone are AOP, because they generate homogeneous or

heterogeneous •OH in the electrochemical systems to oxidize the contaminants. However, in

some other situations, electrochemical regeneration was mainly the desorption of contaminants

from carbon surface due to the electrode polarization or local pH change as evidenced by the fast

initial increase in the total organic carbon (TOC) concentration in the electrolyte solution during

the regeneration process [24].

Microbial regeneration or bioregeneration method regenerates the saturated carbon material

either in an offline mode by mixing the bacteria with the spent carbon or accompanied by the

adsorption in the wastewater treatment process such as powder activated carbon (PAC) treatment

and biological AC treatment [100].

2.3.3 Regeneration methods comparison

All these decomposition-based technologies for the carbon regeneration may have some

advantages and disadvantages. For the thermal regeneration, it is the most mature and the only

commercialized technology. However, because of its high energy intensity, the cost for

regeneration is ~50% of the cost for purchasing the fresh carbon [14]. Additionally, some of the

porous structure and adsorption capacity will be lost after each regeneration cycle. Therefore, it

is highly desirable to develop alternatives and substitution technologies for the carbon

regeneration.

23
The wet oxidation process has been investigated as a substitution for thermal regeneration

since 1961 [101], however, it also suffers from carbon loss. For example, the weight loss for four

continuous regeneration cycles was ~40% for GAC saturated with chemictive brilliant blue R

and cibacron turquoise blue G at a reaction temperature of 200 °C [102]. In addition to the high

carbon loss, the equipment requirement for wet oxidation is complex because the reaction

happens under elevated temperature and high pressure (2-20 MPa).

In comparison to the thermal regeneration and wet oxidation process, AOP, electrochemical

regeneration, and microbial regeneration processes can take place under ambient conditions.

Because of the long regeneration time (longer than 2 days), low regeneration efficiency (only

15% for some adsorbates), and carbon fouling during the microbial regeneration process [13],

microbial regeneration is not suitable for the saturated carbon reactivation. Combined with

conventional activated sludge process, microbial regeneration facilitates the treatment of

wastewater by improving sludge thickening/dewatering and reducing the impact of organic

shock loading [100].

The regeneration efficiency of AOP varied from < 50% to more than 90% [13, 98, 103]. For

some of the AOP, commercial chemicals are required either to degrade the organic pollutants or

to adjust the solution chemistry, while others may require a special energy source such as

ultraviolet (UV) light. The limitation on chemicals supply and the attenuation of UV light in

presence of carbon will make some of the AOP less attractive for carbon regeneration.

For the electrochemical regeneration methods, typically high regeneration efficiencies (>80%)

were obtained [30, 99, 104-107], and modular design of electrochemical cells makes them

attractive to decentralized facilities [108]. However, low efficiencies (< 30%) [17, 109] and

24
generation of toxic byproducts were observed under some conditions for some electrochemical

regeneration studies [110, 111]. To overcome these disadvantages and improve regeneration

efficiency, the research on combining electrochemical methods with AOP technologies is

necessary.

In conclusion, the comparison of different carbon regeneration technologies is summarized in

Figure 2-4. Generally speaking, the regeneration efficiencies of different technologies increased

in the following order: microbial regeneration < AOP ≤ electrochemical regeneration ≤ wet

oxidation process < thermal regeneration. The energy consumptions and the carbon weight loss

increased following the same order. As an emerging regeneration technology that can possibly

make a balance between regeneration efficiency and energy input, electrochemical regeneration

combined with AOP merits further investigation to improve efficiency and treatment capacity.

Electro-Fenton oxidation is the process of interest in this thesis. The following sections describes

first Fenton oxidation and then EF oxidation.

Figure 2-4. Comparison between different decomposition based regeneration methods.

25
2.4 Fenton oxidation

2.4.1 Homogeneous

The mechanism of Fenton oxidation process is initiated with the formation of •OH by the

Fenton’s reaction between Fenton reagents, i.e., H2O2 and Fe2+, equation (2-9). The Fe2+ is then

regenerated from the Fenton-like reaction between Fe3+ and H2O2, equation (2-10). However,

reaction (2-10) is much slower than Fenton’s reaction (2-9). Fortunately, the Fe2+ can be

regenerated more quickly through reactions (2-11) and (2-12) by 𝐻𝑂2• and 𝑂2•− , which are

generated from reactions (2-13) and (2-14) [29].

𝐹𝑒 2+ + 𝐻2 𝑂2 + 𝐻 + → 𝐹𝑒 3+ + 𝐻2 𝑂 +• 𝑂𝐻 (2-9)

𝐹𝑒 3+ + 𝐻2 𝑂2 → 𝐹𝑒 2+ + 𝐻𝑂2• + 𝐻 + (2-10)

𝐹𝑒 3+ + 𝐻𝑂2• → 𝐹𝑒 2+ + 𝑂2 + 𝐻 + (2-11)

𝐹𝑒 3+ + 𝑂2•− → 𝐹𝑒 2+ + 𝑂2 (2-12)

𝐻2 𝑂2 +• 𝑂𝐻 → 𝐻2 𝑂 + 𝐻𝑂2• (2-13)

𝐻𝑂2• ⇌ 𝐻 + + 𝑂2•− (2-14)

𝐹𝑒 2+ +• 𝑂𝐻 → 𝐹𝑒 3+ + 𝑂𝐻 − (2-15)

Among these reactions, H2O2 can be decomposed by reactions (2-10) and (2-13), which are

side reactions competing with Fenton’s reaction. In addition, the generated •OH are consumed

through reactions (2-13) and (2-15) by Fenton’s reagent, which will lead to a decrease in the

oxidizing ability of the Fenton system [29]. Therefore, the concentrations or dosages of Fenton

reagents should be optimized for application. On the other hand, several disadvantages of Fenton

oxidation process have been emphasized for the conventional homogeneous Fenton oxidation,

including the relatively high cost and potential risks for the transportation and handling the

commercial concentrated H2O2, limited working pH range with the optimal pH of 2.8-3.0,
26
accumulation of iron sludge, and the impossible overall mineralization of organic pollutants

because the formation of complexes of Fe(III) with generated carboxylic acid cannot be

destroyed by bulk •OH [29, 112].

2.4.2 Heterogenous

Instead of using the soluble Fe ions as the catalysts, heterogeneous catalysts were also applied

in the Fenton oxidation process which was called the heterogeneous Fenton oxidation. The

catalysts were prepared by loading iron species onto different supports such as zeolite, alumina,

silica, and carbon materials [113]. In order to facilitate separation of the catalysts from a reaction

medium after the treatment, magnetic iron species including zero-valent iron, magnetite (Fe3O4),

and maghemite (γ-Fe2O3) were loaded on the supports through co-precipitation, hydrothermal, or

thermal treatment methods [113]. Application of heterogeneous catalysts can solve some

problems related to the conventional homogeneous Fenton oxidation process, for instance

broader pH range (neutral and basic pH), and no requirement for Fe recovery after treatment

[114].

The generation of •OH during heterogeneous Fenton oxidation could happen in two ways:

through the homogeneous reaction of dissolved Fe ions and through the surface catalytic

reactions [115]. Fe leaching can deactivate the catalyst by attrition [116], as the dissolved Fe ions

only contribute in a minor part to the whole oxidation process [21, 27]. The mechanism of •OH

generation on the heterogeneous catalyst surface with Fe(III) species was proposed and

summarized by different researchers as shown by equations (2-16) to (2-19) [117-119], which

could be well described by a Langmuir-Hinshelwood mechanism [117].

27
≡ 𝐹𝑒 𝐼𝐼𝐼 + 𝐻2 𝑂2 ↔≡ 𝐹𝑒 𝐼𝐼𝐼 𝐻2 𝑂2 (2-16)

≡ 𝐹𝑒 𝐼𝐼𝐼 𝐻2 𝑂2 →≡ 𝐹𝑒 𝐼𝐼 + 𝐻 + + 𝐻𝑂2• (2-17)

≡ 𝐹𝑒 𝐼𝐼𝐼 + 𝐻𝑂2• →≡ 𝐹𝑒 𝐼𝐼 + 𝐻 + + 𝑂2 (2-18)

≡ 𝐹𝑒 𝐼𝐼 + 𝐻2 𝑂2 →≡ 𝐹𝑒 𝐼𝐼𝐼 + 𝑂𝐻 − + 𝑂𝐻 • (2-19)

The oxidation of organic pollutants by •OH could occur on the surface of catalyst or in the

bulk solution. In most cases, the degradation kinetics were derived based on the concentration of

•OH, indicating the importance of oxidation in solution especially for the catalysts having low

adsorption capacity [118, 119].

2.4.3 Fenton oxidation regeneration

As illustrated in Figure 2-5, several reactions can occur during regeneration with Fenton

oxidation. First, adsorbed organic pollutants can desorb into solution to regenerate the adsorption

sites directly (1). The desorbed organic pollutants and intermediates can experience

homogeneous oxidation by •OH to induce more desorption (2). Then, the adsorbed organic

pollutants may be oxidized to CO2 and H2O on the surface to realize regeneration (3). Finally,

the adsorbed organic pollutants could be partially decomposed to intermediates (4), which can be

desorbed from the surface to recover the adsorption sites (5).

28
Figure 2-5. Scheme of Fenton oxidation regeneration mechanism. 1-desorption of adsorbed
species; 2-oxidation of desorbed species; 3, 4-oxidation of adsorbed species; 5-desorption of
intermediates.

Because of the effectiveness of this process, Fenton oxidation has been used to regenerate

carbon adsorbents [21, 120, 121]. The carbon particle size was found to have a significant effect

on the Fenton oxidation regeneration. For example, the regeneration efficiency decreased by

~8.5% per mm with increasing particle diameter of AC saturated with methyl-tert-butyl ether

(MTBE) [28]. This result was attributed to the shorter diffusion distance of H2O2 and MTBE

during regeneration when smaller AC particles were used.

Iron influences the Fenton oxidation regeneration process through the loading, the

distribution, and the specific species present. In general, loading Fe species (such as Fe2O3,

Fe3O4, and Fe salt) on AC increased regeneration efficiency [27, 120]. However, the optimal Fe

loading depended on the specific adsorbate/adsorbent system. For instance, the optimal Fe

loading concentration was obtained to be ~0.7% for regeneration of MTBE saturated AC with

29
the highest MTBE removal and maximum Fe oxidation efficiency [21]. Via acid pretreatment,

Fe was more evenly distributed, which resulted in higher oxidation of adsorbed organic

pollutants [28, 120]. In addition, the presence of Fe2+ in the catalyst (such as Fe3O4) may lead to

higher regeneration efficiency because of the higher Fenton oxidation activity [122]. For

example, Sarasidis et al. found ~4% ferrihydrite-impregnated AC only improved the

regeneration efficiency by ~7% for a diclofenac contaminated AC [123], but in the research of

Do et al. 10% of Fe3O4 loaded on AC increased the regeneration efficiency from <10% to ~70%

[27]. As illustrated in the literature [28, 120], only when the pollutants, H2O2, and the Fe species

meet can Fe contributes to the regeneration and oxidation process.

The porosity of the support for the iron species is also an important factor. Higher

regeneration efficiencies were obtained using mesoporous materials [25, 26, 124]. For example,

Fenton oxidation regeneration of a rhodamine B saturated Fe3O4/RGO with pore size ranging

from 10-20 nm achieved more than 90% after 2 regeneration cycles [26]. Regeneration

efficiency of 90% was reported for the Fenton oxidation regeneration of a methylene blue (MB)

saturated Fe3O4-loaded microsphere resin with average pore size of 23 nm [124].

2.5 Electro-Fenton oxidation

2.5.1 Principles

EF oxidation is a Fenton oxidation based technology, in which the H2O2 is generated in-situ

electrochemically through the two-electron oxygen reduction reaction (ORR) on the cathode [29]

according to equation (2-20).

𝑂2 + 2𝐻 + + 2𝑒 − → 𝐻2 𝑂2 (2-20)

𝐹𝑒 2+ + 𝐻2 𝑂2 + 𝐻 + → 𝐹𝑒 3+ + 𝐻2 𝑂 +• 𝑂𝐻 (2-9)

30
The Fe3+ ion generated from the Fenton’s reaction (2-9) is regenerated to Fe2+ on the cathode

[125] through reaction (2-21).

𝐹𝑒 3+ + 𝑒 − → 𝐹𝑒 2+ (2-21)

Two types of EF reactors have been applied for the treatment of water– namely, divided and

undivided electrochemical cells [29]. The benefit of the divided cells is that the oxidation of

H2O2 to O2 at the anode through reactions (2-22) and (2-23) was avoided [126].

𝐻2 𝑂2 → 𝐻𝑂2• + 𝐻 + + 𝑒 − (2-22)

𝐻𝑂2• → 𝑂2(𝑔) + 𝐻 + + 𝑒 − (2-23)

These two reactions reduce the concentration of H2O2 in the system. In addition, the influence

of the anode on the concentration of Fe2+ through reaction (2-24) is avoided [29].

𝐹𝑒 2+ → 𝐹𝑒 3+ + 𝑒 − (2-24)

However, the overall mineralization of the organic pollutants was difficult to achieve because

of the generation of Fe(III)-carboxylate complexes [29]. In the undivided EF cells, it is possible

to decompose the Fe(III)-carboxylate complexes by oxidant produced from the anode, especially

in the cases when boron doped diamond (BDD) is used as anode [127]. The heterogeneous

BDD(•OH) has a higher oxidation ability than the homogeneous •OH formed by Fenton’s

reaction in bulk solution [128]. The Fe(III)-carboxylate complexes, such as Fe3+-oxalato

complexes, could be efficiently decomposed by UV light in photoelectron-Fenton process with

the combination of photolysis and EF oxidation [129, 130].

31
2.5.2 Cathode

As indicated above, the main reactions of EF oxidation occur on the cathode. Various

materials, including mercury, graphite, carbon-PTFE, graphite felt, carbon sponge, and

reticulated vitreous carbon, have been utilized as cathodes for the EF reaction [29]. Carbon

materials can catalyze two-electron ORR process [131, 132], and have low catalytic activity for

the hydrogen evolution reaction (2-25) and H2O2 decomposition reaction (2-26). Therefore,

carbon cathodes have been widely used for EF oxidation [133].

2𝐻 + + 2𝑒 − → 𝐻2 (2-25)

𝐻2 𝑂2 + 2𝐻 + + 2𝑒 − → 2𝐻2 𝑂 (2-26)

Normally, pure carbon materials, including carbon black, graphene, and CNTs are used as

catalyst supports for the ORR process because of their inert catalytic ability [134, 135].

However, some research indicated that pure carbon materials had intrinsic activity for the two-

electron ORR process to generate H2O2. Using scanning electrochemical cell microscopy

technology, Byers et al. found that there was an intrinsic activity of unmodified defects-free

metallic single wall nanotube for two-electron ORR to H2O2, which was higher than other sp2

carbon material such as graphite and even comparable to standard gold electrocatalysts [135].

Intrinsic activity of CNTs with 2-3 walls for ORR to H2O2 was also observed, which was

attributed to the enhanced charge transfer through the electron tunneling between the outer wall

and inner tubes [132]. Besides the activity for ORR to H2O2 on the basal plane of sp2 carbon

materials, much higher activity was observed at the defects region as well as edge plane of the

carbon materials. In the research of Randviir et al., the peroxide yield on different carbon

materials decreased in order of edge plane pyrolytic graphite > graphene oxide > basal plane

pyrolytic graphite > Q-graphene > pristine graphene [136]. Whereas in the research of Byers et
32
al., a 5-fold increase in electrochemical current for ORR was observed at kinked regions with

defects [135]. In addition, Liu et al. attributed the high content of sp3 carbon and defects to the

high H2O2 production in their system [137].

Except for the intrinsic activity of sp2 carbon and the improved activity at the carbon defects

region for ORR to H2O2, introduction of nonmetal heteroatoms into the carbon materials will

also increase the ORR activity. Improved activity for ORR to H2O2 was observed when the

carbon materials were treated by oxidation [135, 136, 138]. For instance, Miao et al. improved

the H2O2 generation efficiency of graphite felt by 6 times after electrochemically oxidizing in

H2SO4 solution, and the importance of –COOH groups was highlighted in the research [138].

The importance of -COOH and C-O-C functional groups on the H2O2 electrochemical generation

was also demonstrated by Lu et al. [139]. For the oxygen reduction to H2O2 through two

electrons ORR on the carbon materials having oxygen containing functional groups, Jurmann et

al. proposed that the quinone-type groups were the active sites for ORR [140, 141]. Mesoporous

nitrogen-doped carbon materials were also applied for H2O2 electrochemical generation with

high efficiency and selectivity [134, 142, 143]. In most of these studies, the cathode was

prepared by mixing carbon materials with binders, such as PTFE and Nafion, and the H2O2

generation rate (up to >2000 mg/L h) was much higher compared with conventional carbon

materials (< 300 mg/L h) (e.g. graphite felt, carbon felt, and carbon sponge) [137].

Different cathode materials were also tested in a few studies for the electrochemical reduction

of Fe3+ [144]. Petrucci et al. found the iron regeneration rate decreased in the order of reticulated

vitreous carbon > carbon felt > graphite block, and a layer of iron oxide on the graphite cathode

surface was observed when high current was applied [144]. As indicated in the research of Sires

33
et al., carbon felt had much faster Fe2+ regeneration rate compared with gas diffusion electrode

(GDE), which was also demonstrated in some other investigations [145-147].

A low regeneration rate could also be a result of the high concentration of accumulated H2O2,

which can oxidize the Fe2+ back to Fe3+ through Fenton’s reaction and other side reactions [147].

The Fe3+ precipitates as Fe(OH)3 at a lower pH compared with Fe2+. The deposition of iron

species has been observed due to the high local pH on the cathode surface and resulted in the

depletion of soluble Fe in the electrolyte [144]. Overall, the apparent Fe2+ regeneration rate was

lower on the GDE compared with conventional carbon materials.

Instead of using pure carbon materials in the cathode, studies were also carried out by loading

iron species on the cathode for EF reaction. For instance, Zhao et al. used Fe3O4@Fe2O3/AC

aerogel as the cathode during the EF oxidation of pesticide [148]. While in the research of

Ganiyu et al., a Co and Fe loaded carbon felt was applied as the cathode for EF oxidation of an

azo dye [149]. In these research, the electrochemical generation of H2O2 was not reported due to

the fast transformation of H2O2 into •OH and superoxide anions on the cathode surface. As a

consequence, a heterogeneous EF mechanism, different from conventional homogeneous EF

oxidation, was proposed for these cathodes.

The most widely used cathodes for EF oxidation water treatment were carbon/graphite felts

and a GDE made from Vulcan XC-72 carbon because of its high catalytic activity of the Vulcan

carbon [150]. In order to maximize the decomposition of pollutants, different parameters have

been investigated for the EF oxidation process, such as cathodic potential (or applied current)

[127, 151], electrolyte composition and concentration [133], electrolyte pH [128], catalyst types

and concentration [130], oxygen flow rate [133], and anode type [20, 29, 152]. Some of the EF

34
oxidation results for water treatment are listed in Appendix C. In summary, the following two

strategies were applied to improve the EF oxidation performance. The first one was

maximization of the production of •OH with optimal H2O2 generation and Fe2+ regeneration by

controlling cathodic parameters. The other was the use of anode oxidation by using anodes with

high oxidation power, such as BDD anode [127, 153, 154].

2.5.3 Cathodic regeneration

Depending on the compartment where saturated carbon was put in an electrochemical cell, the

electrochemical regeneration can be classified into anodic regeneration (in anodic compartment)

and cathodic regeneration (in cathodic compartment). As indicated in many investigations,

cathodic regeneration had higher regeneration efficiency than anodic regeneration method [30,

155, 156].

Narbaitz and Cen [30] investigated electrochemical regeneration of GAC saturated with

phenol in the cathodic compartment of an electrochemical reactor. According to their research,

cathodic regeneration was ~5-10% higher than anodic regeneration. Based on Narbaitz and Cen’s

research on the cathodic regeneration of GAC saturated with phenol [30], the following

conclusions were extracted. During cathodic regeneration, the regeneration efficiency increased

with increasing current and treatment time. By increasing the current to ~200 mA, the

regeneration efficiency remained at the maximum level of ~90% while the residual phenol was

reduced to non-detectable levels. The regeneration efficiency increased significantly by

increasing the electrolyte concentration of sodium chloride from 0.01% up to 1% beyond which

there was no apparent increase in regeneration efficiency. Among different electrolytes including

NaCl, NaHCO3, Na2SO4, and CH3COONa, the regeneration efficiency reduced in the following

35
order: NaCl > CH3COONa > Na2SO4 >> NaHCO3. In addition, the regeneration efficiency of

GAC increased with decreasing particle size which indicated that the electrochemical

regeneration process was diffusion limited [30]. Zhang [106] also investigated the cathodic

regeneration of GAC by fixing it on the cathode. Similar to experiments of Narbaitz and Cen

[30], several electrochemical parameters including cathodic or anodic regeneration, electrolyte

type or concentration, current density, and treatment time were investigated, and similar results

and conclusions were obtained. In addition, Zhang concluded that the regeneration efficiency

was increased notably in a stirred reactor, which also confirmed that the electrochemical

regeneration was diffusion limited [106, 157].

Narbaitz and McEwen [17] tested the feasibility of electrochemical regeneration on field

spent GAC (adsorption column) saturated with natural organic matter (NOM) obtained from two

water treatment plants in Canada and US. According to their results, the cathodic regeneration

regenerated only 8-15% of NOM adsorption capacity of the spent carbon [17], though a

regeneration efficiency of 80% was achieved for NOM-loaded AC in laboratory scale [104]. In

this study, the desorption of organic pollutants owing to the high pH of the cathodic

compartment was proposed to be the main mechanism for cathodic regeneration. The field spent

GAC had high percentage of irreversibly adsorbed NOM which was resistant to the pH induced

desorption, hence, low regeneration efficiency was obtained for the GAC [17].

As a special case of cathodic regeneration, EF oxidation regeneration was different from

cathodic regeneration with the introducing of oxygen in the system, and it has been applied for

the regeneration of carbon materials. Bañuelos et al. [23] introduced the EF method for

regeneration of AC. In this process, the cathodically generated H2O2 from the reduction of

36
oxygen interacted with the iron ion in the electrolyte or Fe-loaded resins in the cathode to

generate hydroxide radicals, which can decompose molecules adsorbed on the AC. According to

their results, the addition of 0.05 mM FeSO4 in the system can significantly increase the O2

reduction current from 15% to 70%, and the loaded Fe in the cathode was as effective as iron

ions in the electrolyte to catalytically generate hydroxide radicals and decompose organic

pollutants [23]. The same group also applied the EF oxidation for the regeneration of orange II

saturated AC, and complete regeneration was reported for several cycles [158]. Meanwhile, Roth

et al. built a tubular electrochemical cell with a CNT-based GDE and used the cell in a cyclic

adsorption‒EF process to remove Acid Red 14, and 97-100% regeneration of the GDE was

observed during the EF process [159]. The carbon materials were not saturated before

regeneration in those studies and without further information on the saturation adsorption

capacity, the regeneration efficiency could be overestimated.

Based on the research of Bañuelos et al., oxidation was proposed to be the main pathway for

EF regeneration due to the improved regeneration efficiency with the addtion of iron catalyst

[23]. However, as indicated above, the cathodic regeneration without any production of H2O2

and •OH was also able to regenerate carbon materials with high regeneration efficiency. Besides,

in a process similar to EF, where O3 instead of O2 was introduced, the adsorbed p-nitrophenol

was first desorbed from the AC fibre and then oxidized in the electrolyte [24]. Therefore, the EF

regeneration of carbon materials could also result from the desorption of organic pollutants. A

possible mechanism of EF oxidation regeneration was shown in Figure 2-5. The desorption

mechanism could be due to pH induced desorption and cathodic polarization as shown in Figure

2-6. In the first mechanism, hydroxyl ions are generated from the cathodic reactions, and they

can react with the functional groups on adsorbent surface and in the adsorbate molecules. The
37
functional groups become negatively charged and thus impose electrostatic repulsion forces

between the adsorbent and the adsorbate, resulting in desorption. In the second mechanism,

application of a cathodic potential creates an electrical double layer on the surface. The electrical

field within the double layer has a direction pointing to the surface, which can repel negative

charges and lead to the desorption of anions.

Figure 2-6. Two possible desorption processes on a cathodic polarized surface. (a) pH induced
desorption; (b) electric field near surface.

In most EF oxidation processes, graphite/carbon felt or GDE acted as cathode for organic

pollutants removal. These cathodes normally had a low specific surface area, and the

contribution of adsorption to the removal of organic pollutants was often negligible. Trellu et al.

used EF to remove adsorbed organic pollutants on the cathode during treatment of humic acids
38
contaminated water [160]. Therefore, it is possible to use the electrode material as both the

adsorbent during adsorption and as the cathode during an EF regeneration process.

2.6 Knowledge gaps

Although AOP based technologies have been widely applied and investigated for the

regeneration of carbon materials, there are still some limitations and knowledge gaps for the

practical application of these technologies.

(1) Although numerous studies have been carried out for the regeneration of carbon materials

by Fenton oxidation method, the influence of carbon porosity on the regeneration efficiency of

Fenton oxidation has not been well studied.

(2) Even though EF has been proposed and investigated for the regeneration of carbon

materials, limited information is available for this technology about the influence of different

parameters, which limited the analysis of EF regeneration mechanism.

(3) The application of AC-PTFE electrode as both adsorbent for adsorption and cathode for

EF regeneration has not been investigated.

39
Chapter Three: Materials and Methods

This chapter contains the general methodology, materials, and procedures that have been

applied in this research, including preparation of materials and electrode, characterization

methods, and adsorption and regeneration procedures. More specific details are presented in the

corresponding chapters. Additionally, the errors associated with the main steps in this research

were analyzed.

3.1 Methodology

The main objective of this dissertation is the determination of critical factors affecting Fenton

oxidation and EF oxidation regeneration of carbon materials. The regeneration efficiency of

Fenton oxidation and EF oxidation with different operational parameters and adsorbent

properties were studied using MO, MB, and benzoic acid (BA) as model compounds.

Experiments were conducted through adsorbent preparation, initial adsorption, Fenton or EF

oxidation regeneration, and re-adsorption to estimate the regeneration efficiency. Figure 3-1

presents a flowchart showing the experimental strategy of this dissertation.

First, carbon materials were saturated with MO, and then regenerated by Fenton oxidation

with a focus on the effect of the carbon pore structure. Next, EF regeneration of the carbon

materials saturated with MO, MB, or BA under different operational conditions was explored.

The regeneration efficiency of EF and Fenton oxidation was compared to understand the

advantages of adding electrochemistry. The third part presents the feasibility of using AC-PTFE

as both an adsorbent for adsorption and a cathode during EF regeneration.

40
Figure 3-1. Experimental strategy used in this dissertation.

3.2 Materials

3.2.1 Methyl orange

Methyl orange (MO) was chosen as the model adsorbate for most of this research, because it

is a typical organic pollutant and a synthetic dye which is resistant to biodegradation [77, 81],

and it can be easily detected and measured by ultraviolet/visible (UV/Vis) spectrophotometry.

The structure of a MO molecule is shown in Figure 3-2. It is an anionic azo dye with the

molecular formulae of C14H14N3SO3Na and a molecular weight of 327.33 g/mol. The molecular

size was estimated to be 1.61 nm × 0.61 nm × 0.52 nm [81]. The pKa of MO is 3.4. Under different

pH conditions, the MO molecule can interact with protons or hydroxyl ions as shown in Figure

3-3 [88]. At pH < 3.1, MO will change to a quinoid structure and its solution is red; while at pH >

41
4.4, MO will remain as the azo structure, and its solution is yellow. In solution, MO anions tend

to combine with each other to generate dimers or oligomers. According to Kendrick et al., MO

dimerizes at concentrations below 1 mM (327 mg/L) and forms higher oligomers (~6) between 5

and 10 mM (1635-3270 mg/L) [161].

Figure 3-2. Molecular structure of MO.

Figure 3-3. Molecular structures of MO anion under acid and base solution.

3.2.2 Preparation of magnetic carbon materials

Powder carbon materials were used as the adsorbents in Chapter 4 and Chapter 5, because

they have faster adsorption kinetics due to less diffusion limitations. However, it was difficult to

separate them from a solution via simple settling, and the application of filtration could lead to

some carbon loss. To facilitate the separation of carbon materials after the adsorption and

regeneration processes, magnetic carbon materials loaded with γ-Fe2O3 were prepared. The

magnetic carbon materials were prepared according to a method reported previously [162, 163]

with slight modifications, as shown in Figure 3-4.

42
Figure 3-4. Procedure for the preparation of magnetic (γ-Fe2O3) carbon materials.

First, an appropriate amount of FeCl3 and FeSO4•7H2O were dissolved in 20 mL of deionized

water at 70 °C. The solution was maintained at this temperature with constant stirring on a hot

plate, while ~1.0 g of carbon was added to form a suspension, which was subsequently stirred for

30 min before the drop-wise addition of 60 mL of Na2CO3 solution (13.5 g/L). After 1 h, the

suspension was cooled to room temperature and aged for 24 h. The suspension was filtered and

washed with deionized water. The solid on the filter paper was collected and dried at 50 °C

overnight before storing.

3.2.3 Preparation of carbon-PTFE electrodes

Carbon-PTFE electrodes were applied as cathodes and/or adsorbents in Chapter 5 and Chapter

6. They were prepared following a procedure given in the literature [24, 164, 165], as shown in

Figure 3-5.

43
Figure 3-5. Procedure for the preparation of carbon-PTFE electrodes.

Briefly, ~0.1 g of carbon, an appropriate amount of PTFE preparation solution, and a suitable

amount of isopropanol were mixed in a glass bottle under ultrasonication for 20 min. The

mixture was then transferred to a water bath at 80 °C to evaporate the isopropanol until a dough-

like solid was formed. The solid was then rolled and cut into two sheets (4 cm in length and 1 cm

in width), which were then pressed together with a piece of stainless steel mesh in the middle

(width of 1 cm) using a force of 5 MPa for 10 min. The pressed assembly was then calcined in a

muffle furnace at 350 °C for 1 h to obtain the carbon-PTFE electrode.

3.3 Characterization of materials

3.3.1 Nitrogen adsorption-desorption isotherms

The pore structure properties of the samples were determined by nitrogen adsorption-

desorption isotherms at -196 °C. The isotherms were collected by using a physisorption

44
instrument (TriStar II Plus, Micromeritics Tristar Instrument). Before analysis, the samples were

degassed at 105 °C under vacuum (~10 Pa) overnight (>12 h).

The Brunauer–Emmett–Teller (BET) specific surface area was calculated according to

equation (3-1) [166] within the partial pressure of 0.02-0.3 using MicroActive (Micromertitics

Instrument Corporation) software ensuring that the constant C was > zero. The BET equation is

𝑝/𝑝0 1 𝐶−1
0
= + (𝑝/𝑝0 ) (3-1)
𝑉(1 − 𝑝/𝑝 ) 𝑉𝑚 𝐶 𝑉𝑚 𝐶

where V is the volume of adsorbed gas at relative pressure of p/p0, and Vm is monolayer

adsorption capacity. However, it is inappropriate to use BET method to analyze microporous

materials because it is difficult to separate the monolayer-multilayer adsorption and micropore

filling [167]. Therefore, besides BET surface area, a density functional theory (DFT) surface area

was also reported for some samples according to the two-dimensional non-local density

functional theory (2D-NLDFT) method [168] using the SAIEUS program from Micromertitics.

The total pore volume was taken from the N2 adsorption isotherm at partial pressures of 0.96-

0.97. The pore size distributions and micropore volumes were derived from the N2 adsorption

isotherms using 2D-NLDFT method built in the SAIEUS program. The carbon dioxide

adsorption isotherms of several samples were also determined at 0 °C, and the results were

summarized in Appendix E. Similar results of pore size distribution were obtained compared

with nitrogen adsorption. Therefore, in this research, all the pore structure properties were

reported based on the nitrogen adsorption-desorption isotherms.

45
3.3.2 X-ray diffraction

X-ray diffraction (XRD) is the diffraction of X-ray by the atoms in a crystal with regularly

arranged atoms. When the scattered X-rays from different parallel crystal planes are completely

in phase, i.e. in accordance with the Bragg equation (3-2), the X-ray intensity increases with a

corresponding peak in the XRD pattern [169]. The crystal structures of the prepared samples

were determined by an X-ray diffractometer (Multiflex, Rigaku, USA) from 10° to 90° 2-theta

with a scan rate of 2 °/min at 40 kV and 30 mA. The species of the loaded iron oxides on the

carbon were analyzed according to the XRD patterns by comparing with standard patterns in the

Powder Diffraction File in the MDI Jade 5.0 software. The particle size of the deposited iron

oxide was estimated with the Scherrer equation (3-3) [170].

𝜆 = 2𝑑 sin 𝜃 (3-2)

𝐾𝜆
𝐿= (3-3)
𝐵 cos 𝜃

where λ is the wavelength of the X-ray, d is the interplanar spacing, θ is the incident angle

(Bragg angle), L is the apparent crystallite size, K is the Scherrer constant, and B is the half-

width at full maximum of the 2θ peak.

3.3.3 Thermogravimetric analysis

Thermogravimetric analysis (TGA) is a thermal analysis method that measures the weight

change of a sample over time when the temperature changes under different atmospheres. TGA

has been applied for proximate analysis [171]. The ash content was determined by using a

thermogravimetric analyzer (SDT Q600, TA Instruments-Waters LLC) by heating the samples

(~6 mg) at 10 °C/min to 800 °C and holding at this temperature for 10 min in flowing air (50

46
mL/min) while the mass was monitored. The iron oxide loading of magnetic carbon materials

was calculated by subtracting the ash content of the corresponding carbon support.

3.3.4 Fourier-transform infrared spectroscopy

Fourier-transform infrared spectroscopy (FTIR) was used to characterize the functional

groups on the surface of the carbon materials. The energy required to excite the vibration modes

in organic molecules is in the infrared region [172, 173]. The FTIR spectroscopy of the carbon

materials was measured using a NICOLET iS50 FT-IR (Thermo Scientific, USA) in ATR

(attenuated total reflectance) mode. Before the test, ~2 wt % of carbon materials were diluted by

potassium bromide (KBr) and mixed in an agate mortar.

3.4 Analysis of solutions

3.4.1 MO solution and regeneration solution analysis

The concentration of MO solution as well as the UV-vis spectrum of solution after

regeneration was measured by using a UV-vis spectrometer (EVOLUTION 220, Thermo

scientific). Approximately 3 mL of solution was transferred to a 1 cm quartz cuvette, and then

the spectrum was measured within the wavelength range of 200-800 nm. The UV-Vis spectra of

the MO solution at pH 3 and pH 7 are shown in Figure 3-6. Calibration curves of the MO

solutions were obtained according to the absorbance at a wavelength of 464 nm for the solution

at pH 7 and a wavelength of 505 nm for the solution at pH 3 based on the UV-Vis spectra. As

shown in Figure 3-7, linear fits were obtained within the MO concentration ranges of 0-30 mg/L

for pH 3 solution and 0-40 mg/L for pH 7 solution. The UV-Vis spectra of solutions after

regeneration was also measured according to the same procedure except that the wavelength

range was changed to 190-800 nm.

47
Figure 3-6. The UV-Vis spectra of an aqueous MO solution at pH 3 and pH 7. The solution at pH
3 has a MO concentration of 12.6 mg/L, and the MO solution at pH 7 has a concentration of 18.8
mg/L.

Figure 3-7. The calibration curves for MO solutions at pH 3 and pH 7. In the calibration fits, A is
the absorbance and C is the MO concentration (mg/L).

48
3.4.2 H2O2 concentration analysis

As an important reagent for both Fenton oxidation and EF oxidation, H2O2 is necessary to

generate •OH radicals for organic pollutants decomposition. The H2O2 concentration was

quantified by a colorimetric method in which titanium sulfonate (TiOSO4) was used as an

indicator. The TiOSO4 reacts with H2O2 to produce a yellow pertitanic acid according to reaction

(3-4) [174].

𝑇𝑖 4+ + 𝐻2 𝑂2 + 2𝐻2 𝑂 → 𝐻2 𝑇𝑖𝑂4 + 4𝐻 + (3-4)

In this thesis, for H2O2 concentration analysis, 0.25 mL of TiOSO4 solution (1.9-2.1 %,

Sigma-Aldrich) was added into the 3-mL sample and tested by the UV-vis spectrometer

immediately. A typical UV-Vis spectrum of this solution is shown in Figure 3-8. The calibration

curve (Figure 3-9) was generated using various concentrations of a H2O2 solution (30 wt% in

H2O, Sigma-Aldrich) and recording the absorbance at 407 nm. Based on the H2O2 concentration,

different calibration curves were obtained at low concentration (0-100 mg/L) and high

concentration (100-160 mg/L) (Figure 3-9) because of the absorbance deviation from linearity at

higher concentrations.

49
Figure 3-8. The UV-Vis spectra of solutions with H2O2 (~150 mg/L) and TiOSO4 (0.25 mL) as
well as deionized water and TiOSO4 (0.25 mL).

Figure 3-9. The calibration curve for H2O2 concentration determined using TiOSO4 as an
indicator. In the calibration fits, A is the absorbance and C is the H2O2 concentration (mg/L).

50
3.4.3 Adsorbent extraction analysis

To extract the intermediates remaining on the carbon surface after regeneration, 5 mL of

methanol was mixed with ~8 mg of regenerated carbon. The mixture was put in a glass bottle

sealed with a screw and oscillated in a shaker (VWR Symphony 5000I Shaker, Henry Troemner

LLC) at 25 °C for 4 h. The Fe-loaded carbon separated from the solution using a magnet. The

extraction solution was analyzed in the UV-Vis spectrometer.

3.5 Adsorption experiments

3.5.1 Adsorption kinetics

All the adsorption experiments were carried out using a batch method. All the solutions used

in the adsorption experiments were prepared by dissolving an appropriate amount of adsorbate in

deionized water.

For the powder carbon samples, ~0.1 g of fresh carbon was put in 100 mL of MO solution

with different initial concentrations in a beaker. The concentrations were chosen to balance the

saturation of carbon materials and the accuracy of MO concentration measurement. The

suspension was mixed using a magnetic stirrer at a speed of 500 rpm at room temperature. At

different adsorption times, 0.5 mL of solution was taken out from the beaker to determine the

MO concentration.

For the carbon-PTFE electrodes, the adsorption kinetics were determined by putting the whole

electrode (i.e. with the stainless-steel mesh) in a glass bottle with 100 mL of MO solution at an

initial concentration of ~300 mg/L. The bottle was then shaken at 250 rpm at 25 °C. At different

times, 0.5 mL of solution were sampled and diluted to quantify the MO concentration with the

UV-Vis spectrometer.

51
The adsorbed amounts of MO at different times were calculated according to the following

equation (3-5) by considering the volume loss from sampling:

(𝐶𝑡,𝑛 − 𝐶𝑡,𝑛−1 )[𝑉0 − 0.5(𝑛 − 1)]


𝑞𝑡,𝑛 = 𝑞𝑡,𝑛−1 + (3-5)
𝑚

where qt,n and Ct,n are the adsorbed amount of MO (mg/g) and MO concentration (mg/mL),

respectively, at the adsorption time of nth sampling, V0 is the initial MO solution volume (mL),

and m is the carbon mass (g).

3.5.2 Adsorption isotherms

The adsorption isotherms of the powder carbon samples were measured in a batch mode by

fixing the initial MO concentration while changing the carbon amount. The initial MO

concentrations were different for different carbon materials. The initial MO concentrations and

carbon dosage were selected to obtain a fully developed isotherm for each carbon sample. An

appropriate amount of carbon was transferred into 20 mL of MO solution in a glass bottle. The

bottle was sealed with a screw to avoid evaporation and oscillated in the shaker at 250 rpm and

25 °C for 1 day to achieve equilibrium. After adsorption, the MO concentration was determined

by using the UV-Vis spectrometer. The MO adsorption capacity at different conditions was

calculated by using equation (3-6).

(𝐶0 − 𝐶𝑒 )𝑉0
𝑞𝑒 = (3-6)
𝑚

where qe is the equilibrium adsorption capacity (mg/g), C0 is the initial MO concentration

(mg/L), and Ce is the equilibrium MO concentration (mg/L).

52
3.6 Regeneration experiments

3.6.1 Fenton oxidation regeneration

The Fenton oxidation regeneration of the MO saturated carbon materials was carried out in a

beaker following the procedures shown in Figure 3-10. After the first adsorption, the MO

saturated carbon material was separated from the MO solution using a magnet. Then an

appropriate amount of H2O2 solution (0.3 wt%) was added as the regeneration agent. The initial

pH of the solution was adjusted to 3 by adding HNO3 solution. The iron oxides loaded on the

carbon surface acted as the catalysts for the heterogeneous Fenton oxidation. The mixture was

stirred by a magnetic stir at 500 rpm and room temperature (~20 °C). After regeneration for an

appropriate time, the carbon material was separated from the regeneration solution using a

magnet and utilized for a second adsorption under the same conditions as the first adsorption.

The regeneration efficiency was calculated according to equation (3-7), which has been widely

used in the literature [27, 98, 99, 175].

Figure 3-10. Experimental flow diagram of Fenton oxidation regeneration.


53
𝑞𝑟
𝑅𝑒𝑔𝑒𝑛𝑒𝑟𝑎𝑡𝑖𝑜𝑛 𝑒𝑓𝑓𝑖𝑐𝑖𝑒𝑛𝑐𝑦 (%) = × 100% (3-7)
𝑞0

where qr and q0 are the MO adsorption capacity (mg/g) of regenerated and fresh carbon material,

respectively. The adsorption conditions were chosen so that the regeneration efficiency would

not be overestimated as shown in Figure 3-11. As depicted in the Figure 3-11, at low equilibrium

concentration and/or high adsorbent dosage, the regeneration efficiency could be overestimated

[175].

Figure 3-11. Potential overestimation of regeneration efficiency under some adsorption


conditions.

3.6.2 EF regeneration using graphite felt cathode

EF regeneration experiments were conducted in a reactor shown schematically in Figure 3-12.

A three-electrode electrochemical cell was employed and controlled by a potentiostat (SI 1287,

Solartron). The cable connections for the potentiostat are shown in Figure 3-13. The RE2 cable

54
was connected to the WE cable and then attached to the cathode of the electrochemical cell. The

RE1 cable was connected to the reference electrode, and the CE cable was linked to the anode.

In this section, graphite felt was used as the cathode, a graphite rod or platinum (Pt) wire was

used as the anode, and Ag/AgCl was the reference electrode, respectively. The electrolyte

consisted of 0.1 M Na2SO4 (≥99.0%, Sigma-Aldrich) with the initial pH adjusted to 3 by diluted

H2SO4 solution. After MO adsorption, the saturated carbon materials were put into the EF reactor

for regeneration. The regeneration was then carried out by maintaining a constant cathodic

potential (vs Ag/AgCl) with the potentiostat. During regeneration, the suspension was stirred and

aerated continuously with an air flow of ~70 mL/min. After regeneration, the regenerated carbon

materials were used for a second adsorption, and the regeneration efficiency was calculated as

for the Fenton oxidation regeneration.

Figure 3-12. Schematic diagram of EF regeneration system. 1: cathode; 2: reference electrode; 3:


anode; 4: gas inlet; 5: electrolyte; 6: stirrer.

55
Figure 3-13. The cable connections of Solartron SI1287 potentiostate for a three-electrode
electrochemical cell shown in Figure 3-12.

3.6.3 EF regeneration-cathode as adsorbent

In these studies, carbon-PTFE electrode was employed as both an adsorbent for adsorption

and the cathode for EF regeneration. After adsorption, the carbon-PTFE electrode contaminated

with MO was assembled into the EF reactor (Figure 3-12) as the cathode. The other regeneration

conditions were the same as those in the EF regeneration by graphite felt cathode except the

catalyst. Herein, 0.05 mM of FeSO4 was added into the electrolyte as catalyst for the EF

regeneration. The carbon-PTFE electrode was then applied to a second adsorption again to

evaluate the regeneration efficiency.

3.7 Experimental error analysis

Experimental errors in this study were mainly from the adsorption process and regeneration

process. During the Fenton oxidation regeneration process, the initial H2O2 concentration could

vary in small range (< 1%) due to the error of micropipette when adding the 30% H2O2 into the

solution, and the reaction temperature could change within range of 19 °C to 22 °C because the

temperature of the regeneration process was not controlled. The preliminary result in Appendix-

A indicated that the small change (< 1%) in H2O2 concentration and the 3 °C change in
56
temperature had little effect on the regeneration efficiency (< 1%). In the EF regeneration, the

reaction temperature (19 – 22 °C), the electrolyte concentration (< 0.005%), the electrolyte

volume (< 1%), Fe2+ concentration (< 2%), cathodic potential (< 0.03%) may lead to small errors

(< 5%) in the final results, as indicated by the impacts of these factors on the regeneration

efficiency in Chapter 5.

In addition to the errors derived from the variation of regeneration conditions, the other main

contribution of errors was from the analysis of MO concentrations before and after adsorption for

both fresh and regenerated carbon materials. Because MO concentrations were determined by the

UV-Vis spectrometer, the reliability of this method was estimated according the calibration

curves built at different times. Figure 3-14 shows the data points used for the MO calibration

curves within a time range of ~2 years. Almost all the data points are located on the same line.

The slopes of the calibration curves ranged from 0.0718 to 0.0737 with an error of < 3%.

Figure 3-14. The UV-Vis absorbance of the MO standard solutions at 464 nm prepared at
different time. The numbers indicate the time of preparation in the form of YYYYMMDD.

57
In addition, the MO adsorption capacity of the Fe2O3/G5 carbon material was measured using

different calibration curves for several times as shown in Table 3-1. The deviation of adsorption

capacity was very small with an error of < 1%.

Table 3-1. The MO adsorption capacity of Fe2O3/G5 tested at different time.

Testing time MO adsorption capacity


(mg/g)
Aug 28, 2016 178.78
Sep 9, 2016 178.94
May 24, 2017 181.60
Jun 20, 2017 179.11
Jul 26, 2017 181.64
Jul 31, 2017 177.50
Oct 5, 2017 177.49
Dec 3, 2017 179.38
Dec 5, 2017 177.03
Average 179.05
Standard deviation 1.67 (n = 9)

The average (3-8) and standard deviation (3-9) were calculated as follows:

∑𝑛𝑖=1 𝑥𝑖
𝑥̅ = (3-8)
𝑛

𝑛
1
𝑠=√ ∑(𝑥𝑖 − 𝑥̅ )2 (3-9)
𝑛−1
𝑖=1

58
Chapter Four: Impact of Pore Size on Fenton Oxidation of Methyl Orange Adsorbed on
Magnetic Carbon Materials: Trade-off between Capacity and Regenerability

Fenton oxidation has been widely used for carbon regeneration but with unsatisfactory

efficiencies (typically < 60%). The mechanisms behind this low effectiveness have not been well

studied. In this chapter, Fenton oxidation regeneration of MO saturated carbon materials was

investigated with the focus on the impact of the pore structure of the adsorbents. More

specifically three carbon supports, with pore structure ranging from mainly microporous to half

microporous-half mesoporous to mainly mesoporous, were impregnated with γ-Fe2O3 to make

them magnetic for easy separation. The carbon samples were characterized before adsorption and

after regeneration. In addition, adsorption kinetics and isotherms were collected, and the WM

diffusion model and Freundlich isotherm model fit to the data. The adsorption capacity increased

with increasing microporosity while the regeneration efficiency increased with increasing

mesoporosity. Further experiments with varying regeneration and adsorption conditions

suggested that the regeneration process may be diffusion limited. The MO adsorbed in the

micropores was strongly adsorbed and difficult to remove unlike the MO adsorbed in the

mesopores. Thus, there was a trade-off between adsorption capacity and regeneration.

Most of the results in this chapter have been published in the journal Environmental Science

& Technology and reproduced with permission from [Xiao, Y., & Hill, J. M. (2017). Impact of

Pore Size on Fenton Oxidation of Methyl Orange Adsorbed on Magnetic Carbon Materials:

Trade-Off between Capacity and Regenerability. Environmental Science & Technology, 51(8),

4567-4575.] Copyright [2017] American Chemical Society.

59
4.1 Introduction

Adsorption technologies, especially those using carbon materials such as AC [176], CNTs

[177], three-dimensional graphene-based macrostructures [75], and biochar [72], have been

widely investigated for the treatment of water contaminated by organic pollutants. The

adsorption process concentrates the organic pollutants on the carbon matrix, which can then be

disposed or regenerated. Because of the potential of leaching organic pollutants during the

disposal process and high price of AC (>$1500/ton [15]), it may be more environmentally and

economically beneficial to regenerate the used carbon materials. Thus, this study focuses on

improving the regeneration process by studying the relationship between pore size and

regeneration efficiency of three carbon-based adsorbents with different physical structures.

The current commercial thermal regeneration processes require high temperatures (800-900

°C) and suffer from loss of carbon, high energy intensity, and high cost [16, 17]. Desorption-

based technologies may alleviate these problems [13, 16, 18]; the adsorbates are desorbed intact

by manipulating pH, solvent polarity, pressure, and temperature (lower than 300 °C). Many

adsorbates, however, will not desorb at the lower temperatures and so degradation-based AOP,

of which Fenton oxidation is the most common, are used to mineralize the adsorbates. Although

some novel electrochemical AOP reported over 90% regeneration efficiencies [23, 24, 30, 99,

111, 158, 178], the regeneration efficiencies of Fenton oxidation (without electrochemistry) are

generally less than ~60% [21, 25, 28, 98, 120], with exceptions for the treatment of volatile

organic compounds [97, 120].

Some improvements in regeneration efficiency for the Fenton oxidation process have been

achieved by introducing UV light [25], increasing the Fe content [21], using an acid pretreatment

60
[120], increasing temperature [120], and reducing particle size [28]. Despite the improvements,

the regeneration efficiencies were still less than 70% after one cycle. The loss of adsorption

capacity was attributed to (i) damage of the carbon matrix by the strongly oxidizing •OH, (ii)

blockage of pores by the oxidation products, and (iii) incomplete removal of the adsorbates [13].

The first two reasons are inconsistent with the often observed stable adsorption capacity after the

first adsorption-regeneration cycle [27]. That is, carbon deterioration and pore blockage should

continuously decrease the adsorption capacity with each subsequent regeneration cycle. The third

reason has been related to the slow diffusion of H2O2 within the AC [28]. Most adsorbents have

high (>1000 m2/g) surface areas to maximize adsorption capacity, but these high surface areas

can only be achieved with microporous materials, which will have diffusion limitations. Indeed,

the highest regeneration efficiencies (> 90%) were reported for mesoporous carbon materials

[26, 179] but a direct study of the role of pore structure has not been done.

In this paper, we propose that the pore structure of the carbon materials plays a critical role

during the Fenton oxidation regeneration process. To test this hypothesis, three samples with

different sized pores were cycled through adsorption and regeneration steps. The samples were

powders with particle sizes below 0.1 mm that were impregnated with iron oxide to facilitate

separation from solution. MO was chosen as the adsorbate because it is a typical organic

pollutant likely to be irreversibly adsorbed at the testing temperatures (20 °C), removing the

complication of desorption [77, 180, 181], and easily detected by ultraviolet/visible (UV/Vis)

spectrophotometry. Adsorption isotherms and kinetics were collected, the regeneration and

adsorption conditions varied, and the samples characterized at various stages throughout the

experiments.

61
4.2 Materials and methods

4.2.1 Materials

Mesoporous carbon (>99.95%, denoted as SMC) and activated charcoal (denoted as NORIT)

were purchased from Sigma-Aldrich (St Louis, MO, USA), while AC Colorsorb G5 (denoted as

G5) was obtained from Jacobi Carbon company (Columbus, OH, USA). These three carbon

samples were modified using ferrous sulfate (FeSO4•7H2O, ≥ 99.0%, Anachemia, Montreal, QC,

Canada), iron (III) chloride (FeCl3, 97%, Sigma-Aldrich), and sodium carbonate (Na2CO3, ≥

99.5%, EM Science, Gibbstown, NJ, USA). MO (C14H14N3NaO3S, ACS reagent grade, Ricca

Chemical Company, Arlington, TX, USA) was the model organic pollutant. Regeneration was

done with a H2O2 solution (30 wt% in water, Sigma-Aldrich) diluted to different concentrations.

4.2.2 Preparation and characterization of magnetic samples

The magnetic samples were prepared according to a method reported previously [162, 163]

with slight modifications. First, ~0.19 g of FeCl3 and ~0.66 g of FeSO4•7H2O were dissolved in

20 mL of water at 70 °C. The solution was maintained at this temperature with constant stirring

while ~1.0 g of carbon (SMC, G5, or NORIT) was added to form a suspension, which was

subsequently stirred for 30 min, before the drop-wise addition of 60 mL of Na2CO3 solution

(13.5 g/L). After 1 h, the suspension was cooled to room temperature and aged for 24 h. The

suspension was filtered and washed with deionized water. The solid on the filter paper was

collected and dried at 50 °C overnight before storing. Iron oxide without carbon addition was

also prepared by the same procedure. The solids were named as follows: Fe2O3, Fe2O3/SMC,

Fe2O3/G5 and Fe2O3/NORIT and each contained approximately 20 wt% Fe, which was a

sufficient loading to be able to easily magnetically separate them from solution.

62
The crystal structures of the prepared samples were determined by an X-ray diffractometer

(Multiflex, Rigaku, USA) with a scan rate of 2 °/min at 40 kV and 30 mA from 10° to 90° 2-

theta. The particle size of the deposited iron oxide was estimated with the Scherrer equation. The

physical properties of the samples were determined by N2 adsorption at -196 °C. The BET

surface area and total pore volume were determined from the N2 adsorption isotherms within

partial pressure ranges of 0.02-0.3 and 0.96-0.97, respectively. The pore size distributions and

micropore volumes were derived from the N2 adsorption isotherms using the two-dimensional

non-local density functional theory (2D-NLDFT) method. The metal loadings were determined

with TGA (SDT Q600, TA Instruments-Waters LLC, USA) by heating the samples (~ 6 mg) at

10 °C/min to 800 °C and holding at this temperature for 10 min in flowing air (50 mL/min) while

the mass was monitored.

4.2.3 Adsorption isotherms

MO adsorption isotherms were collected using batch adsorption with an incubator shaker

(VWR symphony 5000I, Henry Troemner LLC, USA). Several magnetic samples (8 mg) were

weighed and placed in separate glass vials. Then 20 mL of MO solutions with different

concentrations were added to the vials, which were then placed in the shaker and shaken at 250

rpm and 25 °C for 24 h. After adsorption, the magnetic samples were separated from the solution

with a magnet, and the MO concentration in the solution determined by a UV-Vis

spectrophotometer (EVOLUTION 220, Thermo scientific, USA). The experimental errors were

less than 3.0% at a 95% confidence level based on duplicate experiments.

63
4.2.4 Adsorption and Fenton oxidation regeneration

In a typical experiment, the first step consisted of mixing 0.1 g of fresh magnetic sample and

100 mL of MO solution in a beaker, and stirring the resulting suspension at 20 °C on a stir plate.

To obtain the adsorption kinetics, 0.5 mL of solution was withdrawn at various times, and its

MO concentration determined with the UV/Vis spectrophotometer. After 24 h of stirring, the

suspension was separated with a magnet. The solution layer was decanted from the beaker

leaving the saturated sample behind. Then 40 mL of deionized water was added to the beaker,

and the pH of the suspension was adjusted to 3 by adding a dilute HNO3 solution. Next, 0.4 mL

of 30 wt% H2O2 solution was added and the new suspension stirred at 20 °C for 8 h. This amount

of H2O2 was sufficient for the long regeneration time, as indicated by the H2O2 concentration

during the Fenton oxidation regeneration of Fe2O3/NORIT (Figure 4-1). The pseudo first order

H2O2 decomposition rate constant was calculated to be 1.8 × 10-4 min-1. Considering the dosage

of carbon in the solution, the pseudo second order decomposition rate constant was calculated to

be 9.8 × 10-4 L/(g min) (carbon based) or 4.9 × 10-3 L/(g min) (iron oxides based). This value

was higher than those reported in the literature for Fenton oxidation, which was ~2.0 × 10-4 L/(g

min) for Fe amended GAC [120], and it was within the range of H2O2 decomposition rate

constants by iron (III) oxide nanoparticles, which were 3-27 × 10-3 L/(g min) [182]. There was

still 92% of H2O2 left in the solution after 8 h regeneration, and 84% of added H2O2 would be

left in the solution after 16 h regeneration based on calculation.

64
Figure 4-1. The H2O2 concentration (■) during regeneration of Fe2O3/NORIT. The
Fe2O3/NORIT was regenerated at room temperature at initial pH of 3. ▲ and straight line is the
linear fitting of pseudo first order kinetics model.

The suspension was then magnetically separated, and the upper layer decanted. The remaining

solid was the regenerated magnetic sample. The second cycle involved returning to the first step

but using the regenerated sample rather than fresh sample. After regeneration, the prefix “R-”

was added to the sample name (e.g. R-Fe2O3/SMC).

4.3 Results and discussion

4.3.1 Characterization of samples

The XRD patterns (Figure 4-2) of the prepared samples had six peaks at 30.4°, 35.7°, 43.4°,

53.8°, 57.3°, and 62.9° 2-theta, corresponding to the (220), (311), (400), (422), (511), and (440),

respectively, planes of maghemite (γ-Fe2O3).

65
Figure 4-2. XRD patterns of prepared samples.

Maghemite is magnetic and the average particle sizes were 23 nm, 20 nm, and 18 nm on

Fe2O3/SMC, Fe2O3/G5, and Fe2O3/NORIT, respectively, as calculated by the Scherrer equation

using the MDI Jade 5.0 software. The particle size was similar to that obtained by Oliveira et al.

[162], for Fe2O3 supported on Aldrich Darco G60 AC. The peak at 26.2° 2-theta in the XRD

pattern of Fe2O3/SMC corresponds to the (002) plane of graphite (#08-0415). The patterns of the

samples supported on G5 and NORIT have a peak at 21° 2-theta that may be associated with

silica (#74-0201).

Table 4-1 contains the physical properties of the samples and their metal oxide loadings. The

BET surface areas, total pore volumes, and micropore volumes consistently increased in the

following order: Fe2O3/SMC < Fe2O3/G5 < Fe2O3/NORIT.

66
Table 4-1. Physical properties and metal oxide loadings of the samples.

Surface Total pore Micropore Fe2O3 Ash Fe2O3


Samples area volume volume crystal size content contenta
(cm2/g) (cm3/g) (cm3/g) (nm) (%) (%)
SMC 84 0.22 0.01 - 0.0 -
G5 853 0.54 0.31 - 3.0 -
NORIT 1440 0.69 0.57 - 2.4 -
Fe2O3/SMC 69 0.22 0.01 23 ± 3 22 22
Fe2O3/G5 749 0.49 0.27 20 ± 2 22 20
Fe2O3/NORIT 1110 0.56 0.42 18 ± 4 23 21
MO/Fe2O3/G5 254 0.24 0.04 ndc ndc ndc
R-Fe2O3/SMCb 63 0.23 0.01 ndc 21 21
b
R-Fe2O3/G5 289 0.26 0.09 nd 21 18
b
R-Fe2O3/NORIT 149 0.12 0.05 nd 16 14
a
iron oxide content was calculated by subtracting the ash content of the corresponding carbon
support; b magnetic samples were regenerated under typical conditions: 0.3 wt. % H2O2, MO:
H2O2 =1: 430 (molar ratio), initial pH of 3, and 8 h reaction at 20 °C; c not determined.

After the addition of Fe2O3, the surface areas and pore volumes decreased (except for the pore

volume of the Fe2O3/SMC sample). The majority of this decrease was a result of the change in

composition of the samples – the metal loadings were approximately 20 wt% and, thus, the

specific surface areas were expected to decrease to ~80% of the values for the bare carbon

supports if the added Fe2O3 only increased the mass and did not contribute to the porosity. The

fact that the total pore volumes of the SMC-supported samples were identical (0.22 cm3/g)

suggests that the added Fe2O3 was porous. Comparing the pore size distributions of Fe2O3/SMC

and SMC (Figure 4-3), there were additional pores in the size ranges of 10-20 nm and 30-40 nm

for the magnetic sample.

67
Figure 4-3. Pore size distributions of SMC and Fe2O3/SMC.

As the Fe2O3 crystal sizes (18-23 nm) and metal loadings were similar (~20-22%) on all

magnetic samples (Table 4-1 and Figure 4-4), the decreased pore volumes on the G5- and

NORIT-supported samples may have partially been a result of pores being blocked by the added

Fe2O3. Analysis of the XRD peaks only provides an average crystal size and not the distribution.

Figure 4-5 illustrates the pore size differences of the magnetic samples. Fe2O3/SMC has

mesopores with pores ranging in diameter from 3-20 nm and 28-50 nm, while Fe2O3/NORIT has

mainly micropores 1-2 nm in diameter. Fe2O3/G5 had a mixture of micropores and mesopores up

to ~20 nm. Both NORIT and G5 had some pores of diameter 28-50 nm but considerably fewer

than SMC.

68
(a) (b)

(c)

Figure 4-4. TGA of samples before and after iron oxide loading and regeneration. The carbons
were regenerated under typical conditions of 0.3 wt% H2O2 solution (MO: H2O2 = 1:430, molar
ratio) with an initial pH of 3 at 20 °C for 8 h. (a) Carbon samples, (b) Fresh magnetic samples,
and (c) Regenerated magnetic samples.

69
Figure 4-5. Pore size distributions for fresh (○) and regenerated (●) magnetic samples: (a)
Fe2O3/SMC, (b) Fe2O3/G5, and (c) Fe2O3/NORIT. The insets contain an enlarged view of the
region from 0-2 nm and in all cases the x-axis is the Pore width (nm).

The surface areas and pore volumes were lower after regeneration (again with the exception

of the pore volume of the SMC-supported sample, Table 4-1 and Figure 4-5). The decreases for

the Fe2O3/SMC sample were relatively small compared to the decreases for the other two

samples. Blockage of the smallest pores is visible in the pore size distributions in Figure 4-5 (see

insets in each graph). In addition, the measured ash content decreased after regeneration for the

G5- and NORIT-supported samples. These results are consistent with MO being trapped in the

micropores and increasing the combustible carbon content of the regenerated samples R-

Fe2O3/G5 and R-Fe2O3/NORIT (Figure 4-4).

4.3.2 Adsorption kinetics

The rate of MO adsorption was determined on the fresh and regenerated magnetic samples as

shown in Figure 4-6. Within ~30 min, 80% of the total MO adsorption capacity was reached on

the fresh Fe2O3/NORIT and Fe2O3/G5 samples, and these adsorption capacities were 362 mg/g

and 178 mg/g, respectively. After regeneration, 460 min and 230 min were required to reach

70
80% of the total capacities and these capacities were reduced to 30.7 mg/g and 64.0 mg/g,

respectively. In contrast, the adsorption on the Fe2O3/SMC sample was much faster, with almost

complete adsorption in 30 min on both the fresh and regenerated samples but the total capacity

was lower at 26.5 mg/g. These results confirmed that a mixing time of 1 day (1440 min) was

sufficient to reach equilibrium and obtain the adsorption isotherms.

Figure 4-6. Rate of MO adsorption at 20 °C for fresh (open symbols) and regenerated (filled
symbols) magnetic samples: Fe2O3/SMC (∆, ▲), Fe2O3/G5 (□, ■), and Fe2O3/NORIT (○, ●). The
regeneration conditions were: 0.3 wt% H2O2 solution (MO: H2O2 = 1:430, molar ratio) with an
initial pH of 3 at 20 °C for 8 h.

The WM diffusion model [51, 183] was applied to the data in Figure 4-6 and the resulting

model parameters are given in Table 4-2. The model fit was good in terms of correlation

coefficients (R2 >0.92). According to the WM adsorption kinetics model, if the adsorption

process is only controlled by an intraparticle diffusion process, a straight line through the origin

will be obtained by plotting 𝑞𝑡 versus 𝑡 0.5 [184]. For MO adsorption on all samples, these plots
71
were not linear. When the collected data was divided into three steps, governed by film diffusion,

mesopore diffusion, and micropore diffusion, consistent with previous dye adsorption studies

[183, 185, 186], the fits were linear in each step (Table 4-2). The diffusion rate constants

increased after regeneration for Fe2O3/SMC, while they decreased by 31-74% for Fe2O3/G5 and

64-97% for Fe2O3/NORIT, consistent with the previously discussed data (Table 4-1, Figure 4-5

and Figure 4-6).

Table 4-2. WM diffusion model fits for the adsorption kinetics of the fresh and regenerated
magnetic samples.

WM diffusion model
Samples
𝑘1 𝑘2 𝑘3
Fe2O3/SMC 11.8 (0.96) - -
R-Fe2O3/SMC 15.1 (1.0) - -
Fe2O3/G5 77.0 (1.0) 8.93 (0.92) 0.988 (0.95)
R-Fe2O3/G5 19.7 (1.0) 3.46 (0.95) 0.685 (0.98)
Fe2O3/NORIT 108 (1.0) 19.5 (0.96) 1.33 (0.92)
R-Fe2O3/NORIT 3.59 (1.0) 0.849 (0.96) 0.480 (1.0)
Note: 𝑘1 to 𝑘3 is the adsorption rate constant of WM diffusion model at different steps, numbers
in the brackets are the 𝑅 2 of linear fitting for WM diffusion model.

4.3.3 Adsorption isotherms

The adsorption isotherms for MO on the magnetic samples before and after regeneration are

shown in Figure 4-7 (note the change in the y-axis range between Figure 4-7a and Figure 4-7b).

Consistent with the characterization data described above, the adsorption capacities of

Fe2O3/NORIT and Fe2O3/G5 reduced significantly after regeneration (Figure 4-7b). In contrast,

the isotherms for Fe2O3/SMC as prepared and after regeneration were essentially the same.

72
(a) (b)

Figure 4-7. MO adsorption isotherms on magnetic samples (a) as prepared and (b) after
regeneration: Fe2O3/NORIT (○, ●), Fe2O3/G5 (□, ■) and Fe2O3/SMC (∆, ▲). The solid lines are
the fits to the Freundlich isotherm.

As used in many research publications [27, 98, 99, 175], the regeneration efficiency was

calculated according to equation (3-7). As seen in Figure 4-7, the isotherm shape can change

after regeneration, which complicates the use of equation (3-7) and may lead to an

overestimation of the regeneration efficiency. To avoid this issue, high initial MO concentrations

of 100 mg/L, 400 mg/L, and 600 mg/L were used with Fe2O3/SMC, Fe2O3/G5, and

Fe2O3/NORIT, respectively, according to the isotherms (Figure 4-7) [175]. Under these

adsorption conditions (Figure 4-6), the calculated regeneration efficiencies were 97%, 36%, and

8.5% for Fe2O3/SMC, Fe2O3/G5, and Fe2O3/NORIT, respectively. The regeneration efficiencies

decreased in the order of Fe2O3/SMC > Fe2O3/G5 > Fe2O3/NORIT - the reverse order of their

micropore volumes (Table 4-1).

73
To quantify the change in isotherm shape after regeneration, the Freundlich and Langmuir

models were applied to the data in Figure 4-7 and the resulting parameters are given in Table

4-3.

Table 4-3. The Freundlich isotherm and Langmuir isotherm parameters of MO adsorption for
different magnetic samples.

Samples Langmuir isotherm model Freundlich isotherm model


2
𝑄0 𝑏 𝑅𝐿 𝑅 𝐾𝐹 n 𝑅2
Fe2O3/SMC 28.9 0.28 0.034 0.858 21.1 16.7 0.918
R-Fe2O3/SMC 28.0 0.24 0.040 0.941 18.8 13.0 0.987
Fe2O3/G5 167 1.05 0.0024 0.849 114 13.9 0.998
R-Fe2O3/G5 70.3 0.091 0.027 0.917 32.8 7.69 0.986
Fe2O3/NORIT 352 0.94 0.0018 0.935 229 12.8 0.982
R-Fe2O3/NORIT 38.9 0.18 0.0091 0.706 26.7 15.4 0.860
2
Note: 𝑅 was obtained by nonlinear fitting using SigmaPlot 13.0 software.

The isotherm data was better fit by the Freundlich model according to the values of R2, which

were between 0.86 and 1.0 for the Freundlich model and between 0.71 and 0.94 for the Langmuir

model. The fits were significantly poorer for the R-Fe2O3/NORIT sample than for the other

samples. The values of the surface heterogeneity indicator (7.7 < n < 17, Table 4-3) indicated the

non-uniform nature of the adsorption sites on both fresh and regenerated magnetic samples [10],

which is common for adsorption on carbon substrates [7]. After regeneration, the value of n

decreased for the Fe2O3/SMC and Fe2O3/G5 samples, and increased for the Fe2O3/NORIT

sample. Although the fit was poorer with the Langmuir model, the change in the separation

factor (RL) calculated for this model [10] was consistent with adsorption being less favourable

after regeneration for all samples but more so for the G5- and NORIT-supported samples than

for the SMC-supported samples. The similar adsorption capacities and isotherm parameters of

SMC samples before and after regeneration indicated that the Fenton oxidation process had little

74
effect on the carbon surface properties, which is consistent with the literature that stated little or

no reduction adsorption capacity resulted from oxidation of the carbon surface [28, 121]. Thus,

the sites on which adsorption was the strongest on the G5 and NORIT materials were not

regenerated. Based on the characterization data, these sites are located in the micropores in

which the molecules would have more interaction with the carbon substrate, and hence, a

stronger adsorption.

4.3.4 Effect of loading Fe

Table 4-4 shows the adsorption capacity of the fresh and regenerated pure carbon materials

and magnetic carbon materials. ~20% of iron loading decreased the adsorption capacity of the

carbon materials notably, but the difference on adsorption capacities based on carbon weight was

much smaller. Meanwhile, although the mass based adsorption capacities were significantly

different among the magnetic carbons, the BET surface area normalized adsorption capacities

were similar with slight decrease in the order of Fe2O3/SMC > Fe2O3/NORIT > Fe2O3/G5. The

differences on the BET surface area normalized adsorption capacity might be owing to the

distinction on the pore size distributions of these carbons. According to the MO molecular size of

1.61 nm × 0.61 nm × 0.52 nm [81], MO cannot enter slit pores less than 0.52 nm and cylindrical

pores less than 0.80 nm, where Fe2O3/G5 had most of its miropores. On the other hand, assuming

monolayer adsorption, 69.3%, 50.2%, and 59.0% of the surface of Fe2O3/SMC, Fe2O3/G5, and

Fe2O3/NORIT was occupied by MO molecules after adsorption. The high occupation of

Fe2O3/SMC surface indicates the high accessibility for MO, and the left unoccupied surface

might be owing to the repulsive and steric hindrance effect of neighboring adsorbed molecules

[176].

75
Table 4-4. Effect of iron loading on the adsorption capacity of fresh and regenerated carbon
samples and magnetic samples.

Adsorption capacity (mg/g) Normalized to surface area (mg/m2)


Carbon
Fresh Regenerated Fresh Regenerated
SMC 33.9 31.9 0.404 0.380 a
G5 260 42.6 0.305 0.0499 a
NORIT 476 20.1 0.330 0.0139 a
c c
Fe2O3/SMC 26.5/33.8 25.8/32.9 0.384 0.374 a/0.410 b
Fe2O3/G5 178/223 c 55.7/69.8 c 0.278 0.0744 a/0.221 b
c c
Fe2O3/NORIT 362/458 30.7/38.8 0.327 0.0278 a/0.206 b
Note: a based on the BET surface area of unused carbons; b based on the BET surface area of
regenerated carbons; c based on the carbon mass in the magnetic materials.

Table 4-4 indicates that ~20% of iron oxides loading increased the regeneration efficiencies of

oxidation process, which was attributed to the catalytic effect of iron species on carbon surface

[21, 27, 28, 120]. The influence of iron loading on regeneration efficiency for Fe2O3/G5 was

more significant than its for Fe2O3/SMC and Fe2O3/NORIT. ~20% of iron loading increased the

regeneration efficiency of G5 from 16.4% to 31.3%, while similar result was also obtained by

loading 0.56% of iron which might be owing to better iron dispersion [28, 120]. The regeneration

efficiency of SMC by H2O2 was 94.1%. It was reported that graphite and AC could act as

catalyst for H2O2 decomposition to generate •OH with the participation of basic functional

groups or delocalized π-electrons [31, 187, 188]. The catalytic effect of graphite might account

for the high MO removal from SMC carbon surface.

4.3.5 Effect of adsorption and regeneration conditions

The regeneration and adsorption conditions were varied to study their influence on the MO

capacity and regeneration efficiencies. As the adsorption was fastest on the Fe2O3/SMC sample,

the regeneration time was reduced from 8 h to 4 h and to 1 h. As shown in Figure 4-8, reducing

the time by a factor of two had a relatively small impact on the adsorption capacity while

76
reducing the time by a factor of eight had a much larger impact – capacities of 25.8 mg/g, 24.1

mg/g and 13.6 mg/g, respectively. To confirm that the impact of desorption was minor, a

regeneration experiment with only deionized water (no H2O2) was performed (Figure 4-9). After

desorption, less than 5% of adsorbed MO was desorbed into the water for all the carbon samples.

Figure 4-8. Impact of regeneration conditions on Fe2O3/SMC adsorption capacity. Regeneration


conditions of 40 mL of 0.3 wt% H2O2 solution at initial pH of 3 with different time, and
desorption done with 100 mL of deionized water for 5 h.

77
Figure 4-9. Desorption kinetics of MO saturated carbon materials. ○ Fe2O3/NORIT; □ Fe2O3/G5;
and ∆ Fe2O3/SMC. After desorption, the amount of MO desorbed in water only accounted for
4.0%, 2.9%, and 4.8% of total adsorbed MO on Fe2O3/NORIT, Fe2O3/G5, and Fe2O3/SMC.

The regeneration time, initial H2O2 concentration, and volume of H2O2 solution (reported as a

ratio of H2O2 to MO) were increased separately for Fe2O3/G5 regeneration and all these factors

improved the regeneration efficiencies as shown in Figure 4-10. A similar increase in adsorption

capacity (56 mg/g to ~76 mg/g) was achieved by increasing the regeneration time from 8 to 16 h

or the initial H2O2 concentration from 0.3 to 1.5 wt.%. Increasing the amount of available H2O2

(by increasing the volume of solution while maintaining the initial H2O2 concentration) resulted

in a smaller increase in adsorption capacity - 55.7 mg/g to 64.0 mg/g. In addition, the influence

of shaking speed on the regeneration efficiency was investigated (Figure 4-11). There were no

differences in the adsorption capacities of regenerated Fe2O3/G5 with (250 rpm) or without

78
shaking suggesting that the Fenton oxidation regeneration of MO saturated carbon was not

controlled by diffusion in the bulk solution.

Figure 4-10. Impact of time, H2O2 concentration, and H2O2 amount during regeneration of
Fe2O3/G5. Baseline regeneration conditions of 40 mL of 0.3 wt.% H2O2 solution (H2O2: MO =
70: 1, molar ratio) at initial pH of 3 for 8 h; the H2O2:MO molar ratio was increased by
increasing the H2O2 solution volume while maintaining the initial H2O2 concentration.

79
Figure 4-11. Regeneration kinetics of Fe2O3/G5 with 250 rpm shaking (○) and without shaking
(●). Regeneration conditions: initial pH of 3, ~3.0 % H2O2, Fe2O3/G5: water=1:1 (mg: mL), at 25
°C.

Finally, the MO concentration was varied with the Fe2O3/NORIT sample, which had the

highest initial (i.e., before regeneration) adsorption capacity. Increasing the initial MO

concentrations from 500 mg/L to 520 mg/L to 600 mg/L increased the adsorption capacities from

342 mg/g to 351 mg/g to 362 mg/g, respectively, but also decreased the adsorption capacities

after regeneration from 63.7 mg/g to 50.7 mg/g to 30.7 mg/g, respectively (Figure 4-12). Note,

the regeneration conditions were varied within this set of experiments. A longer regeneration

time (20 h) was used for the second experiment (MO initial concentration of 520 mg/L), and a

larger volume of H2O2 solution (540 mL) was used for the third experiment (MO initial

concentration of 600 mg/L). The application of more rigorous regeneration conditions, which

were effective for Fe2O3/SMC and Fe2O3/G5, were not able to compensate for the additional MO

80
adsorption on the Fe2O3/NORIT sample. Extrapolating the adsorption isotherm for

Fe2O3/NORIT (Figure 4-7) to higher equilibrium concentrations, this sample should have

additional adsorption capacities of ~40 mg/g, 39 mg/g, and 22 mg/g (i.e., equilibrium capacities

at 440 mg/L, 470 mg/L, and 570 mg/L). Thus, even without regeneration, the sample would have

adsorbed more MO on a second exposure to the MO solutions, complicating the calculation of a

regeneration efficiency and possibly explaining some of the discrepancies in the literature.

Figure 4-12. Impact of initial MO concentration on adsorption capacity of Fe2O3/NORIT. For


experiments with 500 mg/L MO initially, regeneration conditions were 40 mL of 0.3 wt% H2O2
solution at initial pH of 3 for 8 h; for experiments with 520 mg/L MO initially, regeneration
conditions were 40 mL of 0.3 wt% H2O2 solution at initial pH of 3 for 20 h; and for experiments
with 600 mg/L MO initially, regeneration conditions were 540 mL of 0.3 wt% H2O2 solution at
initial pH of 3 for 8 h.

4.3.6 Effect of particle size

The particle size of all the three carbons was less than 100 μm. According to Kan and Huling

[120], the calculated Thiele-modulus of 100 μm porous spherical particles was 0.13, indicating

81
the Fenton oxidation regeneration was not limited by H2O2 diffusion even for the largest

particles.

Additional experiments on the influence of particle size on Fenton oxidation regeneration

efficiency were carried out (Figure 4-13).

Figure 4-13. MO adsorption kinetics of fresh and regenerated Fe2O3/NORIT with different
carbon support sizes. <20 μm (● fresh, ○ regenerated); 63-90 μm (▲ fresh, ∆ regenerated); and
125-150 μm (■ fresh, □ regenerated).

There were some differences in the regeneration efficiencies - 17.0% for <20 μm vs 7.23% for

125-150 μm - but the higher external surface area (149 m2/g for <20 μm vs 131 m2/g for 125-150

μm) for the smaller particle sizes accounts for some of this difference. Similarly, the initial

adsorption amount at ~5 min was higher for carbon particle sizes < 20 μm. Nonetheless, the

experimental results for these three samples followed the same trend in the relationship between

82
regeneration efficiency and microporosity, indicating that the regeneration efficiency was mainly

controlled by the pore structure instead of the particle size.

4.3.7 Multiple adsorption-regeneration cycles

The Fe2O3/SMC and Fe2O3/G5 samples were tested for stability over multiple

adsorption/regeneration cycles and the results are shown in Figure 4-14. The adsorption capacity

of Fe2O3/SMC was 27 mg/g initially and then 24 mg/g and 23 mg/g in cycles 2 and 3,

respectively (Figure 4-14a). The regeneration between the third and fourth cycles was improved

by leaving the sample overnight after the 4 h regeneration and solution decanting. The solution

remaining in the pores continued the Fenton oxidation overnight removing more of the MO and

resulting in higher adsorption capacities of 32 mg/g after the fourth and fifth cycles. After the

fifth cycle, the original regeneration method was employed and the adsorption capacity in the

sixth cycle was 25 mg/g, which corresponds to a regeneration efficiency of 93%. The Fe2O3/G5

sample, which had more micropores than the Fe2O3/SMC sample, had a more significant

decrease in adsorption capacity in the first two cycles from 97 mg/g to 78 mg/g, with a gradual

decrease in capacity between cycles 3 (73 mg/g) and 6 (65 mg/g) (Figure 4-14b).

83
(a) (b)

Figure 4-14. The MO adsorption capacity of (a) Fe2O3/SMC and (b) Fe2O3/G5 during 6
consecutive adsorption-regeneration cycles. ~0.1 g of saturated sample was regenerated by 40
mL of 0.3 wt. % H2O2 solution with initial pH of 3 at 20 °C for 4 h (Fe2O3/SMC) and 8 h
(Fe2O3/G5). Before 4th and 5th adsorption of Fe2O3/SMC, the regenerated sample was left in the
beaker (after decanting the solution) for overnight. Adsorption experiments were carried out with
100 mL MO solution of 100 mg/L for Fe2O3/SMC and 50 mL MO solution of 200 mg/L for
Fe2O3/G5.

4.3.8 Relationship between regeneration efficiency and pore structure

To better understand the effect of pore structure properties on the Fenton oxidation process,

regeneration efficiency was plotted versus microporosity (%) and the WM diffusion rate

constants (Figure 4-15).

84
(a) (b)

(c) (d)

Figure 4-15. Relationship between regeneration efficiency and pore structure: (a) micoporosity;
(b)-(d) intraparticle diffusion rate constants.

The regeneration efficiency was inversely proportional to microporosity (Figure 4-15a) - the

higher the fresh sample microporosity, the less effective the regeneration. The relationship

between regeneration efficiency and the parameters of the WM interparticle diffusion model
85
were varied (Figure 4-15b, c, and d). A linear relationship between regeneration efficiency and

𝑘𝑑𝑖𝑓𝑓,2 (mesopore diffusion) was observed (R2 of 0.982), and there appeared to be a positive

trend (R2 of 0.75) with 𝑘𝑑𝑖𝑓𝑓,1 (film diffusion), but no trend with 𝑘𝑑𝑖𝑓𝑓,3 (micropore diffusion).

These results are consistent with the MO binding strongly in the micropores and not being

removed during the regeneration – an explanation that is in line with the studies that reported

regeneration efficiencies of less than 60% for microporous AC regenerated with Fenton

oxidation [21, 25, 28, 98, 120, 189]. As shown in Figure 4-5, the micropores were blocked after

regeneration.

Several studies stated that electrostatic interactions governed the adsorption process for MO

especially when metals or metal oxides were involved [27, 87, 88], but these interactions could

not explain the high MO adsorption capacity at high pH conditions when the adsorbent surface

was negatively charged [27, 78]. Other mechanisms including hydrogen bonding, pore filling,

and van der Waals interactions were also proposed for MO adsorption [78, 81]. Dispersion forces

could dominate the interaction as MO and the surface groups on AC are dissimilar [190]. Due to

the anisotropy of dispersion force as well as repulsive steric forces, the MO molecule would

adsorb with the orientation of closest distance to the carbon surface [190]. In addition, for

solution concentrations below ~300 mg/L, MO anions tend to generate dimers [161]. Adsorption

of a MO dimer could result in complete blockage of a micropore of ~1.6 nm in diameter

according to the MO molecular size (1.61 nm × 0.61 nm × 0.52 nm) [81], and leave the

benzenesulfonate (-SO3-) or dimethylanilinium (-N+(CH3)2) side exposed. This orientation is

illustrated in Figure 4-16.

86
Figure 4-16. Schematic of MO adsorption in two pores of different diameters, and subsequent
reaction with H2O2.

In the Fenton oxidation process, no ferric ions were detected in the solutions after 8 h

regeneration of the samples, which was consistent with literature reports that only a small

amount of Fe (<1%~3%) leached from samples [21, 27, 191]. This result indicated that surface

reactions, instead of homogeneous reactions in solution, mainly contributed to the Fenton

oxidation regeneration [21]. On the sample surface, H2O2 can react with both Fe2O3 and the

carbon [118, 187] to generate •OH, which have short life time (~3 μs) [192]. For the micropores

blocked by the adsorbed MO, •OH can only react with the exposed benzenesulfonate or

dimethylanilinium - rate constants of 4.7 × 105 L/(mol s) (HSO4-→SO4•-) [193] and 3.5 × 107

L/(mol s) (CH3NH3+→•CH2NH3+) [194], respectively, have been reported for these reactions. In

larger pores, •OH can access other sites on the MO to react - a rate constant of 2.0 × 1010 L/(mol

s) [195] has been reported for reactions between •OH and the benzene rings in MO. The latter

reaction rate is several orders of magnitude higher than the reaction rates with the end groups,

87
consistent with MO in mesopores being removed much faster than MO adsorbed in micropores.

The regeneration kinetics of Fe2O3/G5 were measured (Figure 4-11). The adsorption capacity of

the regenerated carbon increased to 42 mg/g in the first 30 min and then to ~62 mg/g over the

next 7 hours. These results are consistent with MO oxidation reaction being the rate-limiting step

in the Fenton regeneration of the studied carbon materials.

4.4 Conclusions

The results of this research illustrate the trade-off between high adsorption capacity and high

regeneration efficiency. As the pore size decreased, more MO was adsorbed but the adsorbed

MO was not easily removed to regenerate the adsorbent. The economics will dictate how many

times an adsorbent should be used to make a process feasible. As shown in Figure 4-14, the

SMC-supported adsorbent had a higher regeneration efficiency. Over 6 cycles, however, the total

amount of MO removed from solution was significantly higher on the G5-supported adsorbent

(450 mg versus 163 mg). Although MO was used in this study, the results are applicable to other

organic pollutants, and the optimal pore structure for a particular contaminant can be determined

and the adsorbent tailored accordingly.

88
Chapter Five: Mechanistic Insights for the Electro-Fenton Regeneration of Methyl Orange
Saturated Carbon Materials

As indicated in Chapter 4, the Fenton oxidation regeneration of carbon materials saturated

with MO was significantly affected by the pore structures of the adsorbents. The Fenton

oxidation regeneration of microporous materials was not effective. To improve the regeneration

efficiency of MO saturated microporous carbon materials, EF regeneration method was applied

in this chapter. Carbon materials loaded with magnetic γ-Fe2O3 were saturated with MO and then

regenerated for 8 h by controlling the cathodic potential using a reference electrode. The

influence of different factors, including cathodic potential, carbon pore structure, cathode type,

anode type, and presence of air flow, on the regeneration efficiency and pore structure recovery

was systematically investigated to better understand the regeneration mechanism. Regeneration

was improved with a higher cathodic potential, better contact with the cathode, and carbon

materials containing larger pores. The addition of an electrochemical potential improved the

regeneration of adsorption sites within micropores. These results support electro-desorption

being the dominant pathway for EF regeneration. Cathodic reduction and •OH oxidation may

also contribute a minor part to regeneration through mineralization of the adsorbed and desorbed

organic pollutants. The majority of results in this chapter have been submitted to the journal of

Water Research.

5.1 Introduction

Adsorption onto AC is an effective method to remove various organic pollutants from water

systems over a wide range of concentrations [176]. After saturation with the contaminants, the

adsorbents often must be regenerated and reused both for environmental and economic

considerations [13, 16, 91, 196]. Various regeneration methods have been developed including

89
thermal [16], biological [13], and AOP [13, 196] approaches. Electrochemical regeneration

methods, in particular, are promising because of the high regeneration efficiencies (70-100%)

obtained and the ease of operation [17, 30]. Higher adsorption capacity materials, such as AC,

are regenerated with cathodic regeneration because for the same carbon material under similar

conditions, cathodic regeneration was 5-10% more efficient than anodic regeneration [30, 106,

197].

To further improve the regeneration for strongly bound adsorbents, Fenton oxidation in

combination with electrochemistry - i.e., electro-Fenton (EF) regeneration - was proposed and

investigated [23, 29, 158, 159, 198]. In this process, oxygen and protons combine to form H2O2

at the cathode at a rate dependent on the cathode type, oxygen flux, and cathodic potential [29,

128, 144]. Typically graphite felt [152, 199, 200] or carbon-polytetrafluorethylene (PTFE) [151,

201] was used as cathode because of the high H2O2 production rate, with high mineralization

efficiencies (> 90%) reported [152, 199, 202]. H2O2 generation was improved by the presence of

air (or pure oxygen), which would further improve the removal efficiency of organic pollutants

from water [133, 164, 200, 201, 203], for instance the orange II removal efficiency increased

from 60% to 90% by increasing oxygen flow rate from 0 to 8 mL/min [133]. On the other hand,

different anodes were used in the EF process to improve the removal of TOC in water, because

only appropriate anodes could degrade the generated Fe(III)-oxalato complexes [151] which

were resistant to the oxidation of bulk •OH [29]. For instance, Pt had much less oxidation power

than BDD and graphite anode which resulted in lower mineralization (20% less) [151] and

slower kinetics (1.75-fold lower) [153]. Although these factors were crucial to the EF oxidation

process, their impacts on the EF regeneration were not investigated.

90
Additionally, the mechanism of EF regeneration was not clear. Regeneration of the adsorbent

and mineralization (or degradation) of the pollutants are related but potentially separate

processes. For example, the pollutant may desorb, regenerating the adsorbent, but not necessarily

be mineralized. Some authors suggest that oxidation is the main cause for regeneration [23, 198,

204] while others suggest that the electro-desorption (i.e., desorption in the presence of an

electric field) has the major contribution to the regeneration [17, 105, 205]. It can be challenging

to unambiguously separate the desorption and oxidation in EF regeneration. Nevertheless,

Bañuelos et al. [23] suggested that the increase (15% to 70%) in regeneration efficiency obtained

in the presence of iron (FeSO4 loaded ion-exchange resin) implied that oxidation was the main

contributor to EF regeneration of toluene on AC. Additionally, peroxide species related oxidation

was ascribed to the increasing porosity of carbon materials during electrochemical treatment

[204].

Despite the importance of the oxidation process [29], in many cathodic regeneration studies

no H2O2 was produced and yet high regeneration efficiencies (>80%-~100%) were reported [30,

105, 109, 205]. During regeneration, a rapid increase in the TOC content of the water within the

first few hours was always observed, which indicated the desorption of organic pollutants and

regeneration of the carbon [24, 160, 206]. Meanwhile, Trellu et al. reported that remaining

adsorbed TOC on the carbon sponge was similar, i.e. similar regeneration, after 9 h EF reaction

for both Pt and more powerful BDD anodes [160], indicating the ineffectiveness of anodic

oxidation on EF regeneration. These results implied the importance of electro-desorption for the

cathodic regeneration of carbon materials saturated with organic pollutants [105, 205].

91
To better understand the regeneration of carbon adsorbents using an EF oxidation process,

three carbon materials with different pore structures were saturated with MO and then

regenerated in an electrochemical cell at room temperature, in this study. The impacts of

cathodic potential, electrode type, air addition, and pore size on the regeneration were

investigated. The results were compared to those obtained using only Fenton oxidation and

related to the physical properties of the carbon materials.

5.2 Materials and methods

5.2.1 Electrodes

Graphite felt (AvCarb G100, FuelCellStore) was used as received and cut into strips of 1.5 cm

× 10 cm. Carbon black (CB)-PTFE cathodes were prepared following a procedure given in the

literature [24, 164, 165]. Briefly, ~0.1 g of CB (>99.9%, Alfa Aesar), an appropriate amount of

an aqueous PTFE solution (60 wt%, Sigma-Aldrich), and 3 mL of isopropanol (≥99.5%, Sigma-

Aldrich) were mixed in a bottle with ultrasonication for 20 min. The mixture was then

transferred to a water bath at 80 °C to evaporate the isopropanol until a dough-like solid was

formed. The solid was then rolled to a thickness of ~0.2 mm and cut into strips (1 cm × 4 cm).

Two strips were pressed together with a piece of stainless steel mesh in the middle (width of 1

cm) using a force of 2000 N for 10 min. The pressed assembly was then calcined in a muffle

furnace at 350 °C for 1 h to obtain the carbon-PTFE cathode. A platinum wire (0.25 mm in

diameter, 99.9%, Sigma-Aldrich) or a graphite rod (6.35 mm in diameter, Pine Research

Instrumentation Inc.) was used as the anode, and a Ag/AgCl electrode (gel reference electrode,

+199 mV vs NHE, Pine Research Instrumentation Inc.) was used as the reference electrode.

92
5.2.2 H2O2 generation

The H2O2 generation was measured in an undivided three-electrode system in a batch mode

(Figure 3-12), where graphite felt or CB-PTFE was the cathode with a platinum wire anode and

Ag/AgCl reference electrode. The H2O2 generation experiments were run in potentiostatic mode

with applied potentials between the cathode and the reference electrode being controlled by a

potentiostat (SI 1287, Solartron). The 100-mL electrolyte consisted of 0.05 M Na2SO4 (≥99.0%,

Sigma-Aldrich) with the initial pH adjusted to 3 by the addition of 0.1 M H2SO4 solution. Before

H2O2 electrogeneration, air was bubbled through the solution for 10 min. During the experiment,

3 mL of electrolyte was taken out for analysis every 20 min and replenished with fresh

electrolyte.

For H2O2 concentration analysis, 0.25 mL of TiOSO4 solution (1.9-2.1 %, Sigma-Aldrich)

was added into the 3-mL sample and tested by the UV-vis spectrometer (EVOLUTION 220,

Thermo scientific) immediately. The calibration curve was generated using various

concentrations of a H2O2 solution (30 wt% in H2O, Sigma-Aldrich), and the absorbance at 407

nm was recorded.

5.2.3 Carbon material preparation, adsorption, and Electro-Fenton regeneration

Iron species attached to the cathode or suspended in the solution act as catalysts for the EF

[23, 149, 207] and Fenton oxidation [124] with the similar effect as adding iron salts into the

electrolyte [29]. Thus, the iron oxides added to the carbon materials catalyzed the reaction as

well as facilitated magnetic separation and recovery of the solids from the solutions. The

magnetic carbon materials were prepared for the experiments using the same method as reported

previously [196]. Briefly, three carbon materials with different pore structures, NORIT (activated

93
charcoal, Sigma-Aldrich), G5 (AC, Jacobi Carbon Company), and SMC (mesoporous carbon,

>99.95%, Sigma-Aldrich), were suspended in a combined ferrous sulfate (≥ 99.0%, Anachemia)

and iron (III) chloride (97%, Sigma-Aldrich) aqueous solution. The iron species were

precipitated onto the carbon surface by adding Na2CO3 (≥ 99.5%, EM Science) solution to obtain

the magnetic carbon materials, which were named as Fe2O3/NORIT, Fe2O3/G5, and Fe2O3/SMC,

respectively. The samples contained approximately 20 wt% iron oxides.

For the adsorption experiments, 0.1 g of each sample was mixed with 100 mL of the MO

(C14H14N3NaO3S, Ricca Chemical Company) solutions with concentrations of 100 mg/L, 400

mg/L, or 600 mg/L, respectively. The concentrations were chosen according to the results in

Chapter 4 to guarantee saturation of the adsorbents before regeneration [196]. The mixtures were

then continuously shaken at 250 rpm and 25 °C for 24 h (VWR Symphony 5000I Shaker, Henry

Troemner LLC). The carbon materials were magnetically separated from the solution before

analysis and/or regeneration.

Regeneration experiments were carried out in the electrochemical reactor (Figure 3-12),

which contained 0.1 g of saturated carbon material and 100 mL of 0.05 M Na2SO4 solution at an

initial pH of 3. The slurry was continuously stirred with air bubbling under a cathodic

polarization of -0.8 V (Ag/AgCl) for 8 h, unless otherwise specified. The majority of the

experiments were carried out with a Pt wire as anode but a graphite rod was also used in a few

experiments for comparison. Graphite felt and CB-PTFE were used as cathodes. A few

experiments were conducted with no air bubbling. In one set of experiments, the potential was

varied between -0.5 and -3.0 V. After regeneration, a second adsorption was performed under the

same conditions as the first adsorption to determine the regeneration efficiency.

94
5.2.4 Characterization of carbon materials

Samples were degassed at 105 °C under vacuum (~10 Pa) for overnight (>12 h) before

nitrogen adsorption at -196 °C (TriStar II Plus, Micromeritics Tristar Instrument). The total pore

volume was determined from the nitrogen adsorption isotherms within a partial pressure range of

0.96-0.97. The specific surface area, micropore volume, mesopore volume, and pore size

distribution were derived from the adsorption-desorption isotherms using the two-dimensional

non-local density functional theory (2D-NLDFT) method.

5.3 Results and discussion

5.3.1 Effect of cathodic potential

The H2O2 production with a graphite felt cathode under different applied potentials was

examined and compared to that generated with the CB-PTFE cathode at -0.8 V (Figure 5-1).

With the latter cathode, the accumulated H2O2 concentration reached 83.3 mg/L after 120 min,

which was similar to that reported by Yatagai et al. (~2.4 mM, ~81.6 mg/L) [146], and

comparable to the values reported when air, instead of pure oxygen, was used [129, 164, 165,

201]. Less H2O2 was generated using the graphite felt cathode with the highest H2O2

concentration of 43.2 mg/L obtained at an applied potential of -0.8 V (vs Ag/AgCl), which was

consistent with the cyclic voltammetry curve of graphite felt (Figure 5-2). This value was similar

to that reported by Zhou et al. (~50 mg/L) under similar conditions [200], and of the same

magnitude of other studies with slightly different conditions [133, 138]. As indicated in these

investigations [133, 200], this H2O2 concentration (~50 mg/L) was sufficient to remove organic

pollutants from wastewater, which was confirmed in this research for the fast decolorization of

MO from an aqueous solution (Figure 5-3). Less H2O2 was produced at voltages higher and

lower than -0.8 V (vs Ag/AgCl, Figure 5-1). Additionally, the current efficiency for H2O2

95
generation decreased with increasing cathodic potential from -0.5 V to -3.0 V with a reasonable

current efficiency of 28% at -0.8 V at 20 min (Figure 5-4). The lower current efficiency at higher

cathodic potential indicated the occurrence of more side reactions, such as hydrogen evolution,

resulting in less H2O2 generated. Therefore, -0.8 V was used for most of the EF regeneration

experiments in this study.

Figure 5-1. H2O2 concentration as a function of time generated with CB-PTFE (carbon black-
PTFE) electrode and GF (graphite felt) at different cathodic potential. Cathodic potential was
relative to Ag/AgCl reference electrode. The cell contained 100 mL of 0.05 M Na2SO4 solution
with an initial pH of 3 and continuous air flow, at room temperature (~20 °C). Lines are only a
guide for the eye.

96
Figure 5-2. Cyclic voltammetry (CV) curve of graphite felt. 100 mL of 0.05 M Na2SO4 solution,
initial pH of 3, at room temperature (~20 °C).

(a) (b)

Figure 5-3. MO removal kinetics by EF oxidation process. 100 mL of 0.05 M Na2SO4 solution
with initial MO concentration of ~30 mg/L, initial pH of 3, at room temperature (~20 °C), at
cathodic potential of -0.8 V (vs Ag/AgCl) using graphite felt as the cathode.

97
Figure 5-4. Current efficiency of H2O2 electrochemical generation by graphite felt at different
cathodic potentials.

The effect of the cathodic potential was also examined for the regeneration of the carbon

materials, and some results for the Fe2O3/G5 sample are shown in Figure 5-5. Both the

adsorption capacity and regeneration efficiency increased with increasing cathodic potential.

Specifically, the adsorption capacity increased from 94 mg/g to 143 mg/g when the applied

cathodic potential increased from -0.5 V to -3.0 V (vs Ag/AgCl). Within the cathodic potential

range of -0.5 V to -1.6 V, the regeneration efficiency increased almost linearly with increasing

cathodic potential as expected given that the electronic repulsion force is proportional to the

cathodic potential (further discussion below). The regeneration efficiency at -1.6 V (vs Ag/AgCl)

with a current of ~140 mA (Figure 5-6) was lower than the efficiencies reported in the studies of

Bañuelos et al. [23, 158] and Hannah et al. [159], where regeneration efficiencies of ~100% were

achieved using similar conditions. In these studies, however, the carbon materials were not

saturated before regeneration [158, 159], which could have led to an overestimation of

98
regeneration efficiency [175]. The regeneration efficiency of 74% (Figure 5-5) was similar to

that reported for the cathodic regeneration of AC cloth saturated with bentazone at -1.5 V (vs

SCE) (~75%) [105], and phenol at 0.2 A (70%-80%) [205].

Figure 5-5. Effect of cathodic potential on the adsorption capacity of regenerated Fe2O3/G5 and
regeneration efficiency. ~0.1 g of MO saturated Fe2O3/G5 (178 mg/g) was regenerated in 100
mL 0.05 M Na2SO4 solution at initial pH of 3 at room temperature (~20 °C) with continuous air
flow. Error bars show the standard deviation of duplicate experiments.

99
Figure 5-6. The potentiostatic polarization curves of graphite felt during H2O2 generation and
regeneration at different cathodic potentials (vs Ag/AgCl).

The applied charge during the 8 h regeneration period, which indicated the energy

consumption of the process, was calculated at different applied potentials according to the

potentiostatic polarization curves (Figure 5-6). As shown in Table 5-1, the regeneration

efficiency increased from 52 % to 74 % as the applied charge increased from 650 C to 5014 C;

this trend has also been observed previously in the literature [104].

Table 5-1. The influence of applied charge on regeneration efficiency under different EF
conditions.

Regeneration conditions Applied charge (C) Regeneration efficiency (%)


a
-0.5 V 650.0 52 ± 3
-0.8 V a 1169 59 ± 3
a
-1.2 V 2671 65 ± 2
-1.6 V a 5014 74 ± 3
a
-3.0 V 21950 80 ± 4
Regeneration by CB-PTFE cathode a, b 775 38 ± 7
b, c
G5-CB-PTFE cathode regeneration 274 89
G5-PTFE cathode regeneration b, c 276 83
a b
Note: Fe2O3/G5 was regenerated using graphite felt cathode; regeneration carried out at -0.8 V
(vs Ag/AgCl) cathodic potential; c the material acted as both adsorbent in adsorption and cathode
in regeneration.

100
Beyond an applied potential of -1.6 V, the increase in regeneration efficiency was minimal

(i.e., from 74 % to 80 % when the applied potential increased from -1.6 V to -3.0 V (vs

Ag/AgCl), indicating that the applied charge was not a limiting factor in the EF regeneration.

This was validated by the regeneration kinetics of Fe2O3/G5 at cathodic potential of -0.8 V as

depicted in Figure 5-7. The regeneration mainly occurred in the first 30 min with an adsorption

capacity of 90 mg/g. The adsorption capacity only increased slightly to 104 mg/g by increasing

regeneration time to 2 h while applied charge increased to 3 times. Furthermore, even though the

applied change increased to 5 times, the regeneration efficiencies at 2 h and 8 h were essentially

the same (58 ± 3% vs 59 ± 3%).

Figure 5-7. The adsorption capacity (■) of regenerated Fe2O3/G5 and applied charge (▲) during
EF regeneration at different time. Regeneration was carried at cathodic potential of -0.8 V.

101
5.3.2 Effect of carbon pore structure-comparison with Fenton oxidation regeneration

The physical characteristics of the adsorbents are shown in Table 5-2. Note, both the surface

areas determined by BET and DFT methods are reported because most previous studies reported

BET surface areas. As has been established for microporous materials, the surface area

determined by the BET equation is larger than that determined by DFT methods because the

BET equation does not account for the increased density of adsorbed nitrogen within micropores

[167].

Table 5-2. Physical properties of adsorbents as determined by nitrogen adsorption.

Total pore Micropore


BET surface DFT surface
Adsorbent volume volume
area (m2/g) area (m2/g)
(cm3/g) (cm3/g)
Fe2O3/NORIT
Fresh 1110 801 0.56 0.42
Regenerated, Fenton 149 109 0.12 0.05
a
Regenerated, -0.8 V 431 292 0.23 0.16
Regenerated, -3.0 V a 705 434 0.35 0.27
Fe2O3/G5
Fresh 749 639 0.49 0.27
MO saturated 254 184 0.24 0.07
Regenerated, Fenton 289 234 0.26 0.09
Regenerated, -0.5 V a 414 306 0.32 0.13
a
Regenerated, -0.8 V 436 355 0.32 0.15
Regenerated, -1.2 V a 479 421 0.36 0.17
a
Regenerated, -1.6 V 576 466 0.42 0.19
Regenerated, -3.0 V a 562 413 0.43 0.18
a b
Regenerated, -0.8 V , no air 526 448 0.37 0.19
Regenerated, -0.8 V a, MB c 300 214 0.26 0.09
Fe2O3/SMC
Fresh 69 70 0.22 0.01
Regenerated, Fenton 63 67 0.23 0.01
a
Regenerated, -0.8 V 72 68 0.23 0.01
a b
Note: EF regeneration at different cathodic potentials (vs Ag/AgCl); air was not bubbled
through the electrolyte; c regeneration of MB saturated Fe2O3/G5.

102
The Fe2O3/SMC sample had the lowest porosity and this porosity was not significantly

changed after Fenton or EF regeneration (at a cathodic potential of -0.8 V (vs Ag/AgCl)). In

contrast, the surface area, total pore volume, and micropore volume decreased significantly for

the more porous Fe2O3/NORIT and Fe2O3/G5 samples after regeneration (Table 5-2), suggesting

that these samples were only partially regenerated.

More porosity was recovered after EF than Fenton regeneration. For example, for

Fe2O3/NORIT, the DFT specific surface area, total pore volume, and micropore volume were

109 cm2/g, 0.12 cm3/g, and 0.05 cm3/g, respectively, after Fenton regeneration, compared to 293

cm2/g, 0.23 cm3/g, and 0.16 cm3/g after EF regeneration. The pore size distributions of Fe2O3/G5

regenerated by Fenton and EF oxidation are compared to a fresh sample in Figure 5-8. Within

the mesopore range, there were only slight differences between the samples. On the other hand,

within the micropore range, the differences were significant. The micropore peak intensity for

the fresh sample was twice as large as the peaks in the micropore region for the sample after EF

regeneration and over four times as large as that after Fenton regeneration. These peaks occurred

at pore widths of 0.88 nm for the fresh sample and 0.92 nm for the regenerated samples. Thus,

EF regeneration was significantly more successful at recovering the micropores less than 1 nm

than Fenton oxidation method, although not all micropores were recovered.

103
Figure 5-8. 2D-NLDFT pore size distributions of fresh and regenerated Fe2O3/G5. Insert is the
pore size distribution within the mesoporous range (2-50 nm). ▲ fresh; ■ regenerated by EF; ●
regenerated by Fenton oxidation.

The adsorption capacities of the samples after EF regeneration were measured. The initial MO

adsorption capacities were 359 mg/g, 178 mg/g, and 27.8 mg/g for the Fe2O3/NORIT, Fe2O3/G5,

and Fe2O3/SMC samples, respectively. After regeneration with EF oxidation (-0.8 V vs

Ag/AgCl), the adsorption capacity was essentially unchanged for Fe2O3/SMC (26.5 mg/g), but

decreased to 135 mg/g and 106 mg/g for Fe2O3/NORIT and Fe2O3/G5, respectively. This trend is

similar to that observed when regenerating these materials with a Fenton oxidation [196] – that

is, the larger the pore size, the easier the regeneration. The addition of an applied potential

improved the regeneration as shown in Figure 5-9, in which the regeneration efficiency is plotted

as a function of the microporosity of the sample.

104
Figure 5-9. Regeneration efficiency as a function of adsorbent microporosity at different applied
potentials for the Fenton and EF (-0.8 V and -3 V) oxidation of carbon adsorbents saturated with
MO. During Fenton oxidation regeneration, ~3 wt% H2O2 was added.

The Fe2O3/NORIT sample had 75% microporosity and the regeneration efficiency increased

from 8% after Fenton regeneration to 38% and 64% with an applied potential of -0.8 V and -3.0

V, respectively. The Fe2O3/G5 sample had 55% microporosity and for the same change in

applied potentials, the regeneration efficiency changed from 38% to 59% and 80%, respectively.

These changes in the adsorption capacity and regeneration efficiency were consistent with the

variation of the pore structure properties among different carbon materials as explained above.

As the cathodic potential increased, the differences among the regeneration efficiencies of

different carbon materials became smaller (Figure 5-9), suggesting that an increased potential

overcame the stronger bonding in the smaller pores. Additionally, although in the Fenton

oxidation experiments the initial H2O2 concentration was ~3000 mg/L compared to 43.3 mg/L in

the current EF regeneration experiments, lower regeneration efficiencies were obtained for

105
Fe2O3/NORIT and Fe2O3/G5. In addition, a regeneration efficiency of 57 ± 7% was obtained for

MO saturated G5 (without iron) after an EF regeneration (Figure 5-10). These results suggested

that desorption of the adsorbates, rather than the H2O2 concentration, was rate-limiting for the

regeneration of the carbon adsorbents.

Figure 5-10. Adsorption capacity and regeneration efficiency (■) of different adsorbates on
different carbon materials. Regeneration was carried out at cathodic potential of -0.8 V.

The relationship between regeneration efficiency and the change in specific physical

properties after regeneration was further investigated (Figure 5-11). Regeneration efficiency was

positively and linearly correlated with the percentage of micropore volume (R2=0.968) and DFT

surface area (R2=0.946) recovered, consistent with the recovery of micropores being critical for

the regeneration of MO saturated carbon materials [205]. The regeneration efficiency was also

positively related to the recovery of total pore volume and mesopore volume but with lower

correlation coefficients (0.897 and 0.439, respectively). Thus, although it is easier to regenerate

the larger pores, these pores have lower adsorption capacities [124, 196].
106
(a) (b)

(c) (d)

Figure 5-11. The relationship between regeneration efficiency and the recovery of carbon pore
structure properties.

5.3.3 Effect of electrodes

Two different cathodes, GF and CB-PTFE, were used for the regeneration of Fe2O3/G5. After

regeneration, the adsorption capacities were 106 mg/g and 67 mg/g with corresponding

regeneration efficiencies of 59% and 38%, respectively (Table 5-3). The total applied charges

were 1169 C and 775 C with the GF and CB-PTFE cathodes (Table 5-1), respectively. The lower

107
applied charge does not fully account for the lower regeneration efficiency with the CB-PTFE

cathode because, as discussed above, the current is not limiting in these systems.

Table 5-3. Effect of cathode type on EF regeneration of Fe2O3/G5 saturated with MO. A cathodic
potential of -0.8 V (vs Ag/AgCl) was used for each experiment.

Cathode Initial MO adsorption MO adsorption capacity after Regeneration


type capacity (mg/g) regeneration (mg/g) efficiency (%)
Graphite
178 106 59
felt a
CB-PTFE a 178 67 38
b c c
G5-PTFE 40 33 83
G5-CB-
53 c 47 c 89
PTFE b
Note: a saturated adsorbent suspended in the electrolyte; b saturated adsorbents as cathode, and
the regeneration was carried out in the presence of 1 mM FeSO4 with intial pH of 3 at room
temperature (~20 °C); c adsorption was carried out in ~100 mg/L MO solution for 1 day at 25 °C,
and capacities were calculated based on the G5 mass in the electrode.

During the EF regeneration with a GF cathode, the electrolyte was turbid initially with the

black Fe2O3/G5 particles suspended in the solution, but after 8 h of regeneration became

transparent (Figure 5-12). Analysis of the cathode after the regeneration confirmed that the

carbon particles had deposited on the porous GF cathode (Figure 5-13). This deposition was not

observed with the non-porous CB-PTFE cathode. The increased contact with the GF cathode

compared to the CB-PTFE cathode could account for the improved regeneration efficiency.

Several published studies have reported high regeneration efficiencies (>90%-100%) when the

saturated carbon materials contacted, or acted as, the electrodes [24, 105, 109, 204-206, 208,

209].

108
(a) (b)

Figure 5-12. Reactor pictures before (a) and after (b) EF regeneration using graphite felt as
cathode.

(a) (b)

Figure 5-13. Pictures of graphite felt before (a) and after (b) regeneration.

109
To follow-up on this idea, several carbon-PTFE electrodes were prepared, exposed to MO,

and regenerated with 1 mM of FeSO4 in solution (Table 5-3). More specifically, the prepared

G5-PTFE or G5-CB-PTFE electrodes were put in MO solutions of ~100 mg/L, dried overnight

under ambient conditions, and then connected as the cathodes in the electrochemical

regeneration system (Figure 3-12). Although the G5-PTFE and G5-CB-PTFE electrodes had

differences in composition, carbon weight, and initial adsorption capacity, they both had high

regeneration efficiencies (82% and 89%, respectively), which were the highest efficiencies

obtained, although the lowest total charge was applied (Table 5-1). These efficiencies were

comparable or even superior to those reported in the literature under similar electrochemical

conditions [104, 105, 198] but as mentioned above, may not be comparable to the other

regeneration efficiencies reported in this study because the electrodes were not saturated with

MO. Nevertheless, the results suggest that improved regeneration is obtained when the adsorbent

is in direct contact with the electrode.

The regeneration efficiencies obtained with different anodes with and without air flow are

shown in Table 5-4. Within error, the results were the same for the NORIT and G5 based

adsorbents with either a graphite rod or Pt wire as anode. The regeneration efficiency of

Fe2O3/SMC, however, was lower when the anode was a graphite rod. The decreased adsorption

capacity may be due to the anodic oxidation and dissociation of the graphite rod [153] into debris

that would contribute to the mass obtained after regeneration. As the SMC support had the

lowest adsorption capacity, this sample would be most sensitive to contamination affecting the

mass.

110
The effect of shutting off the air flow was different for all three adsorbents - there was a

negligible change for the Fe2O3/SMC sample, an improvement for the Fe2O3/G5 sample, and a

detraction for the Fe2O3/NORIT sample. Generally the presence of air (or oxygen) improved the

H2O2 generation, and in turn, the removal efficiency of organic pollutants from water [133, 164,

200, 201, 203]. The regeneration efficiency was not improved by using an anode with higher

oxidation power. Besides, presence of air flow even decreased the regeneration efficiency. These

results indicated that the regeneration was not controlled by the oxidation process, which is

consistent with the regeneration process being electro-desorption controlled.

Table 5-4. Effect of anodes and air flow on MO adsorption capacity of regenerated carbon and
regeneration efficiency.

Carbon Anode Air flow Capacity (mg/g) Regeneration efficiency (%)


Fe2O3/NORIT Graphite rod Yes 133 ± 23 37 ± 6
Fe2O3/NORIT Pt wire Yes 135 ± 8 38 ± 2
Fe2O3/NORIT Pt wire No 108 ± 17 30 ± 5
Fe2O3/G5 Graphite rod Yes 113 ± 12 64 ± 7
Fe2O3/G5 Pt wire Yes 106 ± 5 59 ± 3
Fe2O3/G5 Pt wire No 142 ± 6 80 ± 4
Fe2O3/SMC Graphite rod Yes 19 ± 3 70 ± 12
Fe2O3/SMC Pt wire Yes 26 ± 2 95 ± 9
Fe2O3/SMC Pt wire No 24 ± 1 87 ± 9

5.3.4 Mineralization of MO

UV-visible spectra of the electrolyte during EF regeneration of Fe2O3/NORIT at -3.0 V (vs

Ag/AgCl) were obtained (Figure 5-14). Initially the electrolyte solution had little absorbance

consistent with the MO being fully adsorbed on the carbon sample. After 0.5 h of regeneration,

however, a peak developed at a wavelength of 246 nm, which may be consistent with the

cathodic reduction of the azo structure in MO through reactions (5-1) and (5-2) [159, 210-212].

111
As regeneration continued, the peaks decreased in intensity. This behavior is consistent with

intermediates formed by MO reduction, possibly negatively charged sulfanilate [213, 214],

desorbing from the carbon and then being oxidized in the electrolyte.

𝑅1 − 𝑁 = 𝑁 − 𝑅2 + 2𝑒 − + 2𝐻 + → 𝑅1 − 𝑁𝐻 − 𝑁𝐻 − 𝑅2 (5-1)

𝑅1 − 𝑁𝐻 − 𝑁𝐻 − 𝑅2 + 2𝑒 − + 2𝐻 + → 𝑅1 − 𝑁𝐻2 + 𝑁𝐻2 − 𝑅2 (5-2)

Figure 5-14. The UV-vis spectra of the electrolyte solution during the EF regeneration of
Fe2O3/NORIT. The spectrum of an MO solution is included for comparison. Regeneration
conditions: 100 mL 0.05 M Na2SO4 solution at initial pH of 3 at room temperature (~20 °C) with
continuous air flow at cathodic potential of -3.0 V (vs Ag/AgCl).

Further evidence of the reduction and oxidation reaction of MO during regeneration was

demonstrated by extraction of the adsorbed species from a regenerated carbon sample (Figure

5-15). For the MO in methanol, it had maximum adsorption at 426 nm which was associated

with azo double bond [210]. However, after regeneration, the peak was shifted to 395 nm, which

could be the partial reduced azo bond or N, N-dimethyl-4-nitroaniline [215, 216]. In addition, a

112
shoulder peak at 246 nm, which was assigned to sulfanilic acid [213], was also detected from the

methanol extraction solution of regenerated carbons. Meanwhile, the peak at ~207 nm, which

was associated with the phenolic compounds [217] indicating oxidation of benzene ring, was

developed during EF regeneration. The presence of these peaks implied the incomplete

regeneration of carbon materials, but the increase in cathodic potential reduced the remaining

organic chemicals in regenerated carbons which was consistent with the higher regeneration

efficiency.

Figure 5-15. The UV-visible spectra of methanol solution after extraction of EF regenerated
carbon. ~8 mg regenerated carbon was extracted by 5 mL methanol at 25 °C for 4 h.

5.3.5 Regeneration and mineralization mechanisms

As explained above, the increasing of cathodic potential during EF regeneration process

increased the regeneration efficiency (Figure 5-5) and micropore volume recovery (Table 5-2) as

well as decreased remaining organic chemicals in the regenerated carbons (Figure 5-15). These

results could be illustrated by the electro-desorption mechanism, which was also concluded in
113
other investigations on cathodic regeneration [24, 105, 205]. Negative charge would build up on

the cathode surface during cathodic polarization [205]. The negatively charged MO anions (or

intermediates) could be repulsed by the electrostatic force and attracted by the solvent. These

two forces can compete with the dispersion force between the MO anions and the carbon surface

to determine whether adsorption or desorption occurred. The interaction energy of a molecule

with a surface can be estimated according to the following equations (5-3), (5-4), (5-5), and (5-6)

[190, 218],

𝐸𝑠𝑡𝑎𝑡𝑖𝑐 = −𝑞𝜑 (5-3)

𝐸𝑑𝑖𝑠𝑝𝑒𝑟𝑠𝑖𝑜𝑛 = −𝛽𝛼𝜑 2 /2𝑟 2 (5-4)

𝑞2 1
𝐸𝑠𝑜𝑙𝑣𝑎𝑡𝑖𝑜𝑛 = (1 − ) (5-5)
8𝜋𝜀0 𝑎 𝜀

𝐸𝑡𝑜𝑡𝑎𝑙 = 𝐸𝑠𝑡𝑎𝑡𝑖𝑐 + 𝐸𝑑𝑖𝑠𝑝𝑒𝑟𝑠𝑖𝑜𝑛 + 𝐸𝑠𝑜𝑙𝑣𝑎𝑡𝑖𝑜𝑛 (5-6)

where 𝐸𝑠𝑡𝑎𝑡𝑖𝑐 , 𝐸𝑑𝑖𝑠𝑝𝑒𝑟𝑠𝑖𝑜𝑛 , and 𝐸𝑠𝑜𝑙𝑣𝑎𝑡𝑖𝑜𝑛 are the interaction energy (J) owing to electrostatic

interaction, dispersion, and solvation, respectively; 𝑞 is the charge of a molecule which was -1.6

× 10-19 C for the MO anion, 𝜑 is the potential (V) of the carbon surface, 𝛽 is the enhancement

factor in the pores (ranging from 1-2), 𝛼 is the electric polarizability (C2m2/J), 𝑟 is the distance

(m) between the centers of MO and carbon, 𝜀0 is the dielectric permittivity of vacuum which is

8.854 × 10-19 C2/(J m), 𝑎 is the radius (m) of the MO anion, and 𝜀 is the dielectric constant of

water which is 80.4. According to the equation, the electrostatic interaction could be attractive or

repulsive depending on the polarity of the charge and the potential, while dispersion is always

attractive and solvation energy is always positive (in terms of adsorption system). The total

interaction energy is a parabolic function of the carbon surface potential (𝜑). This explained the

non-linear increase in the regeneration efficiency with respect to the cathodic potential (Figure
114
5-5). All the potentials tested in this study were located in the increasing interval of the

interaction energy function, and the electrostatic repulsion was much higher than the dispersion

interaction. However, according to the equations, further increase in the cathodic potential may

result in the decrease in the regeneration efficiency, which was observed for the cathodic

regeneration of acid orange 7 [109]. In fact, the application of a potential would result in the

increase in the adsorption capacity for m-cresol [219], perfluorooctanoic acid and

perfluorooctane sulfonate [220], phenol [221], and aniline [222], which indicated the improved

dispersion interaction and/or dipole-dipole interaction in these situation. Consequently, instead of

regeneration, the adsorption capacity would increase if no repulsive interaction exists.

On the other hand, the importance of attaching carbon material to the cathode (Table 5-3) was

also consistent with the electro-desorption mechanism. On the surface of an electrode, most of

the potential drop originated from the double layer of the electrode [223], hence, the potential

was the largest near the cathode surface. Because all the tested potentials were within the

increasing interval of the interaction energy function as indicated above, the electrostatic

repulsion was more profound and the interaction energy was higher when the carbon was

attached to the cathode. As a consequence, a larger regeneration efficiency was obtained for the

attached carbon materials compared with the suspended ones.

In addition, the significance of pore structure on the regeneration efficiency can also be

explained by the relationship among electrostatic repulsion, solvation, and dispersion interaction

during EF regeneration process. First, the adsorption of MO molecules in micropores would

decrease the solvation energy because of the inaccessibility of the water molecules. On the other

hand, dispersion interaction between MO anions and the carbon surface was stronger in

115
micropores than in mesopores and on surfaces owing to the enhancement of the pores [190]. As a

result, the interaction energy was decreased and thus the regeneration efficiency was decreased

(Figure 5-9) when the same potential was applied on a micropores carbon material.

To test the importance of electro-desorption during EF regeneration process, Fe2O3/G5

saturated with 245 mg/g of MB, which had positive charge, was also regenerated at a cathodic

potential of -0.8 V (vs Ag/AgCl) (Figure 5-10). The regenerated Fe2O3/G5 only had adsorption

capacity of 88 mg/g with corresponding regeneration efficiency of 36%, which was much lower

than regeneration of MO. This value was similar to the regeneration efficiency of Fenton

oxidation in our previous study for MO saturated Fe2O3/G5. The lower regeneration efficiency

resulting from positive charged adsorbate indicated that electro-desorption played a crucial role

in the EF regeneration. In addition, the EF regeneration of the BA saturated Fe2O3/G5 was also

carried out (Figure 5-10) with a regeneration efficiency of 70%. BA is a weak acid having a pKa

of 4.2 [224], and it will mainly exist in the form of benzoate ions in the cathodic area because of

the high local pH as shown in Figure 5-16 [17]. The benzoate ion had -1 charge which was the

same as MO anion, so they had same electrostatic interaction energy at the same potential.

However, the benzoate ion had a smaller radius which resulted in higher solvation energy, and it

had a smaller electric polarizability because of the smaller molecular size [190] which led to

smaller dispersion interaction. Therefore, a higher regeneration efficiency was obtained for the

BA saturated carbon material compared with the MO saturated ones during EF regeneration.

116
Figure 5-16. The fraction of BA and benzoate at different pH. Benzoate is the main species in
solution at pH over 8.

However, there is a debate on the origin of the electro-desorption force [17, 205]. Some

researchers indicated the importance of cathode polarity [205], while others emphasized the

significance of local pH increasing [17]. The local pH can be estimated according to the Nernst

equation (5-7) by assuming the hydrogen evolution reaction (5-8) reaches equilibrium.

𝑅𝑇 𝑎𝑂𝑥
𝐸 = 𝐸𝜃 + ln (5-7)
𝑧𝐹 𝑎𝑅𝑒𝑑

1
𝐻 + + 𝑒 − → 𝐻2 (5-8)
2

where Eθ is the standard reduction potential at the temperature of interest T, R is the universal gas

constant, z is the number of electrons transferred, F is the Faraday constant, aOx and aRed are the

activities of oxidant and reductant. At cathodic potential of -0.8 V (vs Ag/AgCl), the pH on the

surface of cathode was estimated to be 10.2.

117
In order to differentiate these two desorption mechanisms, MO saturated Fe2O3/G5 was

desorbed in a pH 14 solution for 8 h, which was more basic than the local pH on cathode during

EF regeneration at cathodic potential of -0.8 V (vs Ag/AgCl). The MO adsorption capacity of

Fe2O3/G5 after desorption was only 54 mg/g with corresponding regeneration efficiency of 30%

which was much lower than EF regeneration. The result indicated that the pH induced desorption

only partially contributed to the electro-desorption.

On the other hand, during the EF regeneration process, the presence of air flow (Table 5-4)

and higher H2O2 generation rate (Figure 5-17) did not increase the regeneration efficiency. The

irrelevant or even negative influence of H2O2 generation rate on regeneration efficiency (Figure

5-17) implied the regeneration was not dominated by the H2O2 related oxidation process. On the

other hand, switching anodes between Pt wire and graphite rod did not change the regeneration

efficiency (Table 5-4), which indicated that the anodic oxidation had small influence on the

regeneration process. In addition, similar regeneration efficiency of 57 ± 7% for MO saturated

G5 (without iron loading) by EF also indicated the little influence of oxidation process on

regeneration. Nevertheless, the decrease in UV-visible absorbance of the electrolyte (Figure

5-14) and the presence of phenolic compounds in regenerated carbon (Figure 5-15) indicated that

oxidation process also played an important role in the EF regeneration, which was mineralization

of the desorbed organic pollutants.

118
Figure 5-17. Relationship between H2O2 generation rate (derived from the initial 20 min) and
regeneration efficiency of Fe2O3/G5 after EF regeneration.

According to results obtained from this research, a mechanism was proposed for the EF

regeneration of MO saturated carbon materials (Figure 5-18). Three different reactions can occur

during EF regeneration. The first one is the enhanced desorption due to the presence of electric

field. During EF regeneration, the saturated carbon surface contacts with the cathode to be

negatively charged due to cathodic polarization. Besides, consumption of protons on the cathode

can result in local pH increase which also adds negative charge to the carbon surface. The

negative charged MO anion was repulsed from the carbon surface due to the electrostatic

interaction. This is the main pathway for EF regeneration of MO saturated carbon. Second, the

•OH generated from the in-situ produced Fenton reagents (FeII and H2O2) can oxidize the

desorbed or adsorbed MO. Finally, MO can be reduced electrochemically to generate aromatic

amines. The last two reactions may contribute to the EF regeneration via generation of

intermediates which can be easily desorbed.


119
Figure 5-18. Proposed scheme for the EF regeneration of carbon materials.

5.4 Conclusions

This study investigated the EF regeneration of carbon material saturated with MO. Within the

studied factors, cathodic potential, cathode type, and pore structure of carbon material had

significant influence on the regeneration efficiency, while the anode type and presence of air

flow had limited effect on regeneration efficiency. Higher regeneration efficiency was obtained

with higher cathodic potential, better contact with cathode, and larger pore structure. In

comparison to Fenton oxidation regeneration, EF regeneration had higher regeneration efficiency

because it was capable of regenerating adsorption sites within micropores < 1 nm. Based on the

obtained results, the EF regeneration process was dominated by the electro-desorption of the

adsorbed organic pollutants, which was followed by the further mineralization of the desorbed

organic chemicals by reduction and oxidation.

120
Chapter Six: Investigation of the Impact of Hydrophilicity on Activated Carbon-
Polytetrafluoroethylene Electrode Adsorption and Electro-Fenton Regeneration

As indicated in Chapter 5, the electro-desorption of MO was the main pathway for

regeneration during EF process, and attaching carbon materials to the cathode can improve the

regeneration efficiency. Therefore, it is beneficial to apply the carbon material as both an

adsorbent for adsorption and the cathode for EF regeneration. However, most of the cathodes

that have been used for EF process were mainly graphite felt and carbon-PTFE electrode. They

usually had low adsorption capacity for removing organic pollutants from water. Therefore, in

this chapter, AC-PTFE electrodes with high specific areas and high adsorption capacities were

prepared and applied for MO adsorption and EF regeneration.

More specifically, G5 AC was mixed with different amounts of PTFE to obtain electrodes

with different G5-PTFE ratios, and the effect of G5-PTFE ratio on the MO adsorption and EF

regeneration was investigated. The addition of PTFE in the electrode had small effect on the pore

structure of the carbon material based on the G5 weight, but it had significant influence on the

hydrophilicity (or wettability) of the electrodes. The MO adsorption kinetics data indicated that

the hydrophilicity of the electrode had a linear relationship with the MO diffusion rate constant,

i.e., the higher the hydrophilicity, the higher the adsorption rate. This result indicated that the

hydrophilicity of electrodes controlled the MO adsorption process. The higher hydrophilicity

also resulted in higher regeneration efficiency of the EF regeneration process. In order to obtain a

satisfactory adsorption rate (within hours) and regeneration efficiency (> ~70%), the contact

angle of electrode with water needs to be lower than 90°.

121
6.1 Introduction

Carbon adsorption is one of the most effective and efficient methods to remove organic

pollutants from water. A high adsorption capacity is possible for these materials because of their

tunable pore structure and surface functional groups. After adsorption, the carbon material needs

to be regenerated to reduce costs and possible secondary pollution. Numerous investigations

have studied the regeneration of carbon materials by using technologies including thermal

regeneration [16, 225, 226], solvent extraction and desorption [13, 177], biological regeneration

[100], and electrochemical regeneration [17, 30, 99, 107]. As an AOP, EF oxidation regenerates

the adsorbent and decomposes the adsorbed pollutants simultaneously [23].

As a water treatment technology, EF oxidation has been widely investigated for the treatment

of all kinds of organic pollutants in water [20, 29, 128, 153, 160, 227]. For example, Trellu et al.

investigated the EF oxidation of humic acids in water using a carbon sponge as the cathode

[160]. According to this research, >80% of the initial humic acids in water were first adsorbed

and then decomposed, regenerating the carbon sponge [160]. Roth et al. used a CNT-based

electrode for the adsorption of Acid Red 14, and the electrode was then regenerated in an EF

process with regeneration efficiency of 97-100% [159]. AC saturated with toluene and Orange II

was also regenerated (> 95%) with EF oxidation [23, 158].

As indicated by the results in Chapter 5, the main mechanism of EF regeneration was electro-

desorption, hence, attaching the contaminated carbon materials to the cathode resulted in better

regeneration. In general, higher regeneration efficiencies were reported when carbon materials

had better contact with the electrodes or acted as the electrodes [24, 105, 109, 204-206, 208,

122
209]. Therefore, it is beneficial to use the carbon materials as both the adsorbents and then the

cathode for EF regeneration.

Carbon materials were frequently used as the cathode for the EF process because they have

high activity for the two-electron ORR to produce H2O2 and low catalytic activity for the

hydrogen evolution reaction and H2O2 decomposition reaction [133]. Graphite/carbon felt and

GDEs made from carbon materials and PTFE have been applied as the cathode during the EF

oxidation process [29, 129, 153, 202]. As indicated in some investigations, however, most of the

conventional cathodes were not suitable for adsorption of organic pollutants because of their low

adsorption capacity (< 10 mg/g) [228, 229].

Recently, AC-PTFE cathodes were prepared and applied for ORRs in microbial fuel cells

[230, 231]. Because the AC had a higher specific surface area, a higher ORR activity was

observed [231]. The application of AC based cathodes for EF oxidation of organic pollutants in

water was reported with reasonable removal rates (>90% in 1h) [232, 233]. The application of

AC-PTFE electrodes for adsorption and EF regeneration has not been reported in the literature.

In addition, the incorporation of PTFE in an electrode reduced the hydrophilicity of the electrode

[33], which may result in a lower water and water-soluble species transportation within the

electrode [234], and impact the adsorption and EF regeneration process.

In this chapter, the application of AC-PTFE electrodes as both the adsorbent during

adsorption and the cathode during EF regeneration was proposed and studied. More specifically,

the impact of the electrode hydrophilicity on the adsorption and regeneration process was

investigated. AC-PTFE electrodes with different AC to PTFE ratios were prepared, and their

hydrophilicities were measured. The electrodes were then tested for MO adsorption followed by

123
EF regeneration. The relationships between the electrode hydrophilicity and MO adsorption

kinetics and regeneration efficiency were then analyzed to determine the appropriate

hydrophilicity.

6.2 Materials and methods

6.2.1 Preparation of AC-PTFE electrodes

As-received ColorSorb G5 (denoted as G5) from Jacobi Carbon company was sieved, and

particles < 90 μm were collected for the AC-PTFE electrode preparation. An appropriate amount

of PTFE preparation solution (60 wt% dispersion in water, Sigma-Aldrich) was added into a

glass vial, and then ~5 mL of isopropanol (>99.5%, Sigma-Aldrich) and ~0.1 g of G5 carbon

were added to the vial which was then sonicated for 20 min. The obtained slurry was put in a

water bath at 80 °C to evaporate the isopropanol until a dough-like material was formed. After

that, the material was rolled into two sheets (1 cm in width and 4 cm in length), and they were

then pressed together with a stainless-steel mesh (40 mesh) in the middle under 2000 N for 10

min. The obtained assembly was then transferred into a muffle furnace and calcined at 350 °C

for 1 h. The electrodes with different mass ratios of G5: PTFE were prepared, and the electrodes

with G5: PTFE ratios of 7: 1, 5: 1, 3: 1, 2: 1, and 1: 1 were named as G5-PTFE-7-1, G5-PTFE-5-

1, G5-PTFE-3-1, G5-PTFE-2-1, and G5-PTFE-1-1, respectively. The highest G5: PTFE ratio

was chosen as 7: 1 because this was the highest ratio that the prepared electrode can bind

together using the current preparation method. The as-prepared electrodes had a carbon layer of

~0.2 mm on each side.

6.2.2 Characterization of AC-PTFE electrodes

The nitrogen adsorption-desorption isotherms of the electrode materials were measured at -

196 °C (TriStar II Plus, Micromeritics Tristar Instrument). Before analysis, the materials were
124
degassed at 150 °C under a vacuum of < 10 Pa overnight (>12 h). The BET surface area and total

pore volume of the electrode materials were obtained based on the nitrogen adsorption isotherms

within the partial pressure ranges of 0.05-0.3 and 0.97-0.99, respectively. The pore size

distribution and DFT surface area were derived from the whole nitrogen adsorption-desorption

isotherm using a two-dimensional nonlocal density functional theory (2D-NLDFT) method.

Contact angles were measured using the sessile drop method by dropping 7 μL of deionized

water on the electrodes. The contact angles were analyzed by using ImageJ software. The surface

morphology of the electrode material was analyzed by a scanning electron microscope (SEM,

Phenom Pro, PhenomWorld) at an accelerate voltage of 15 kV. The surface elemental

distribution was mapped by using energy-dispersive X-ray spectroscopy (EDX). The functional

groups of the G5-PTFE electrode material were determined with Fourier-transform infrared

spectroscopy (FTIR, Nicolet iS50, Thermo Fisher Scientific,). For the FTIR analysis ~2 wt % of

electrode material was diluted in potassium bromide (KBr) and ground in a mortar pestle before

analysis. The FTIR spectrum was obtained using an attenuated total reflection (ATR) accessory

from 4000 cm-1 to 400 cm-1.

6.2.3 Adsorption and regeneration

The fresh and regenerated G5-PTFE electrodes were used for MO (C14H14N3NaO3S, Ricca

Chemical Company) adsorption under the same conditions. In a typical run, the electrode was

put in 100 mL of ~300 mg/L MO solution and oscillated in a shaker (VWR Symphony 5000I

Shaker, Henry Troemner LLC) at 25 °C and 250 rpm for 2 days. During the adsorption process,

~0.5 mL of MO solution was taken out from the system, diluted to appropriate concentration,

and then analyzed by a UV-Vis spectrometer (EVOLUTION 220, Thermo scientific). The

concentration of MO solution at different times was determined to obtain the adsorption kinetics
125
of the electrodes. The adsorption capacity was calculated based on the G5 content in the

electrode.

After adsorption, the MO contaminated G5-PTFE electrode was installed in the EF reactor

(Figure 3-12) as a cathode. Platinum (Pt) wire (10 cm in length and 0.25 mm in diameter, Sigma-

Aldrich) and an Ag/AgCl electrode (gel reference electrode, +199 mV vs NHE, Pine Research

Instrumentation Inc.) acted as the anode and reference electrode, respectively. The regeneration

experiment was conducted in 100 mL of 0.05 M Na2SO4 (≥99.0%, Sigma-Aldrich) solution at

initial pH of 3. Approximately 1 mM of FeSO4 (≥ 99.0%, Anachemia) was added into the

electrolyte as the catalyst for the EF reaction. The experiments were run under potentiostatic

mode at a cathodic potential of -0.8 V (vs Ag/AgCl), controlling by a potentiostate (SI 1287,

Solartron), for 8 h with continuous stirring and ~70 mL/min air flow. After regeneration, the G5-

PTFE electrodes were used for a second adsorption experiment under the same conditions as the

first adsorption. The electrochemical H2O2 generation of the G5-PTFE electrode was carried out

under the same conditions as the EF regeneration except that no FeSO4 catalyst was added in the

system. The concentration of H2O2 at different reaction times was determined by the UV-Vis

spectrometer after mixing ~3 mL of sample with 0.25 mL of TiOSO4 solution (1.9-2.1 %,

Sigma-Aldrich).

6.2.4 Regeneration cycles

Two G5-PTFE-7-1 electrodes were applied for several adsorption-regeneration cycles with or

without changing the electrolyte for each regeneration cycle. For the first electrode, the

adsorption experiments continued for 2 days, and the regeneration was conducted by supplying

126
fresh electrolyte for each cycle. For the other electrode, the adsorption experiments only lasted

for 8 h, and the electrolyte was stored and reused for the next regeneration process.

6.3 Results and discussion

6.3.1 Characterization of AC-PTFE electrodes

The pore structure of the G5-PTFE electrode materials was characterized by the nitrogen

adsorption and desorption isotherms (Figure 6-1a). All the G5-PTFE electrode materials had

similarly shaped nitrogen adsorption-desorption isotherms, but the nitrogen adsorption capacity

decreased with increasing PTFE content. Correspondingly, the pore size distributions of these

materials (Figure 6-1b) were similar with micropores centered at ~1.0 nm and some mesopores

at ~3.5 nm, ~10 nm, and ~15 nm. On the other hand, the peak of pore size distribution at ~1.0 nm

shifted to lower pore width, which indicated the blockage of some pores with a higher PTFE

content.

(a) (b)

Figure 6-1. The nitrogen adsorption and desorption isotherms (a) and pore size distribution (b) of
G5-PTFE electrode materials. ◄ G5; ♦ G5-PTFE-7-1; ▼ G5-PTFE-5-1; ▲ G5-PTFE-3-1; ●
G5-PTFE-2-1; ■ G5-PTFE-1-1. Note, the insert contains the pore size distribution for pores less
than 2 nm.

127
Table 6-1 contains the surface area, pore volume and contact angles for the G5-PTFE

electrode materials. The BET surface area, DFT surface area, micropore volume, and total pore

volume decreased with a lower G5 content. Normalizing these properties by the percent G5 in

the electrode material, all the G5-PTFE materials had lower values for these properties (except

the DFT surface area for G5-PTFE-2-1), among which the G5-PTFE-1-1 had the lowest values.

This indicated that the G5-PTFE electrodes preparation process can block some pores (up to 15%

based on total pore volume) in the G5 AC. Except that, all the G5-PTFE electrodes had similar

pore structure properties based on the G5 percentage in the electrodes.

Table 6-1. The pore structure properties of the G5-PTFE electrode materials.

AC-PTFE G5 BET DFT Micropore Total Contact


electrodes percentage surface area surface area volume volume angle
(%) (m2/g) (m2/g) (cm3/g) (cm3/g) (°)
(a) (b) (a) (b) (a) (b) (a) (b)
G5-PTFE-1-1 50.0 393 786 328 656 0.14 0.28 0.25 0.50 121
G5-PTFE-2-1 66.3 565 852 474 715 0.20 0.30 0.36 0.54 118
G5-PTFE-3-1 75.0 638 851 498 664 0.22 0.29 0.40 0.53 88
G5-PTFE-5-1 83.2 708 851 564 678 0.25 0.30 0.45 0.54 86
G5-PTFE-7-1 87.5 770 880 580 663 0.26 0.30 0.49 0.56 47
G5 100 918 704 0.31 0.59 -
Note: (a) based on the total mass of G5-PTFE; (b) based on the mass of G5 in the G5-PTFE.

Additionally, the hydrophilicity of the electrodes was also measured by using the sessile drop

method (Figure 6-2). The contact angle of water on G5-PTFE-1-1 was similar to the value (120-

126°) of pure PTFE material as reported in the literature [235]. This electrode as well as G5-

PTFE-2-1 were hydrophobic because the contact angles were larger than 90° [236]. On the other

hand, the contact angle of G5-PTFE-7-1 was 47°, which was similar to the values reported for

AC [237] and AC-PTFE (10:1 weight ratio) material [33]. The contact angles were higher than

those reported by Dong et al., and the possible explanation was that the electrodes in this
128
research were calcined at a higher temperature for longer times [33]. The contact angles

decreased as the PTFE content decreased consistent with PTFE lowering the affinity of the

electrodes to water.

G5-PTFE-1-1 G5-PTFE-2-1 G5-PTFE-3-1

G5-PTFE-5-1 G5-PTFE-7-1

Figure 6-2. Measurement of contact angle with deionized water for electrodes with different G5
to PTFE ratios.

Figure 6-3 shows the SEM and EDX images of the G5-PTFE electrode materials. The SEM

images show that the materials consisted of angular particles roughly ~10-40 μm in length and

irregular structures attaching to some of the particles. The EDX images indicate that the irregular

structures were composed of fluorine containing materials, i.e. PTFE. The PTFE was distributed

all over the G5 particles with some aggregates, especially for the G5-PTFE-1-1 electrode

material.

129
G5-PTFE-1-1 G5-PTFE-1-1, Florine

G5-PTFE-7-1 G5-PTFE-7-1, Florine

Figure 6-3. The SEM and EDX images of fresh G5-PTFE-1-1 and G5-PTFE-7-1.

Figure 6-4 shows the FTIR spectra of the G5-PTFE electrode materials after preparation and

the fresh G5 material. No peaks were observed from wavenumbers of 4000 cm-1 to 2250 cm-1 for

all materials, indicating a lack of hydrogen related functional groups such as C-H, O-H, and N-H

[172]. The peaks at 2150 cm-1 ~2000 cm-1, and 1560 cm-1 may be assigned to C≡C, C=C, and

benzene ring C=C stretching [63, 172], separately, which were the basic structures of AC G5.
130
Consequently, the intensity of these peaks decreased with an increase in PTFE content. The

broad peak within 960-1270 cm-1 might be ascribed to the C-C bond [172]. The peak at 1150 cm-
1
corresponded to the C-F stretching vibration from the PTFE molecules [172, 238]. The absence

of peaks from 1600-1900 cm-1 and sharp peaks at ~1000 cm-1 indicated a lack of oxygen

functional groups, C=O and C-O [172, 239], on G5 and the prepared G5-PTFE materials. The

FTIR spectra implied that the electrode preparation method did not add functional groups to the

surface of the G5 AC even though the mixture was calcined at 350 °C for 1 h.

Figure 6-4. The FTIR spectrum of G5-PTFE electrode materials. The materials were diluted to 2
wt% in KBr before analysis.

6.3.2 Hydrogen peroxide generation

The kinetics of electrochemical generation of H2O2 by different G5-PTFE cathodes were

measured (Figure 6-5). The curves reached or were approaching plateau values, which varied

between ~14 mg/L for G5-PTFE-7-1 and >75 mg/L for G5-PTFE-1-1. At longer times,

131
decomposition of H2O2 on the anode may occur [200]. The H2O2 concentration at 2 h decreased

with the decrease in PTFE content and had similar values when the G5: PTFE ratios were over 3.

The highest H2O2 concentration of 79 mg/L was obtained with a G5: PTFE ratio of 1:1 after 2 h

reaction at -0.8 V cathodic potential. This value was similar to the accumulated H2O2

concentration (~75 mg/L) produced by a commercialized GDE under similar conditions [146].

As shown in Figure 6-5b, a lower concentration of H2O2 (< 20 mg/L) was generated when the

electrodes became hydrophilic. The low H2O2 concentration produced by cathodes with higher

carbon-PTFE ratio was also observed by Zhou et al. for a graphite-PTFE cathode [164]. The

authors ascribed cathode flooding and low gas diffusion as the reasons for lower H2O2

production on hydrophilic cathodes [164].

(a) (b)

Figure 6-5. The H2O2 production kinetics (a) of different AC-PTFE electrodes at applied
cathodic potential of -0.8 V (vs Ag/AgCl) and the relationship between H2O2 concentration at 2 h
and hydrophilicity (represented by cosθ). Experiments were carried out in 100 mL of 0.05 M
Na2SO4 electrolyte at pH 3 under room temperature (~20 °C).

6.3.3 Electro-Fenton oxidation of MO

The G5-PTFE electrodes were also applied as cathodes for EF oxidation of MO in solution.

100 mL solution with ~25 mg/L MO, 0.05 M Na2SO4, and 1 mM FeSO4 at initial pH of 3 was
132
transferred into the EF reactor (Figure 3-12). A cathodic potential of -0.8 V was kept for 2 h, and

samples were taken out at different reaction times to measure the MO concentration. The

degradation kinetics of MO by using different G5-PTFE cathodes are depicted in Figure 6-6. All

the cathodes removed > 95% of MO after 2 h reaction, and the kinetics were fitted by a pseudo

first-order model (Table 6-2).

Figure 6-6. EF oxidation kinetics of MO with different G5-PTFE cathodes. Reaction conditions:
initial pH of 3, 0.05 M Na2SO4 as electrolyte, 1 mM FeSO4 as catalyst, air flow rate of ~70
mL/min, at room temperature (~20 °C).

Table 6-2. Pseudo first-order rate constant of MO removal by EF using different G5-PTFE
cathodes.

Cathode First-order rate constant (k1)


G5-PTFE-1-1 0.059
G5-PTFE-2-1 0.098
G5-PTFE-3-1 0.028
G5-PTFE-5-1 0.035
G5-PTFE-7-1 0.025

133
Among all the cathodes, the G5-PTFE-2-1 had the highest rate constant of 0.098 min-1, which

was similar to those in the literature under similar conditions for PTFE based cathodes [164,

165]. The G5-PTFE-7-1 had the lowest rate constant of 0.025 min-1, which had similar removal

rate of MO using an air diffusion AC packed electrode at applied current of 50 mA [232]. The

UV-Vis spectra of the electrolyte during EF oxidation (Figure 6-7) shows the decrease in the

absorbance at ~500 nm and shifted to lower wavelengths, which indicated the degradation of the

azo structure in the MO molecules. Additionally, the absorbance within the wavelength range of

200-350 nm increased due to the generation of intermediates during MO decomposition.

Comparing the spectra among different cathodes, more intermediates were produced for the G5-

PTFE-1-1 and G5-PTFE-2-1 cathodes than the other cathodes. The low concentration of

intermediates during EF oxidation using the G5-PTFE-3-1, G5-PTFE-5-1, and G5-PTFE-7-1

cathodes implied that adsorption partially contributed to removal of the MO during the EF

oxidation. The inconsistent trend of pseudo first-order rate constants with G5 to PTFE ratios may

be due to the different H2O2 generation, Fe2+ regeneration, and adsorption rate.

G5-PTFE-1-1 G5-PTFE-2-1

134
G5-PTFE-3-1 G5-PTFE-5-1

G5-PTFE-7-1

Figure 6-7. The UV-Vis spectra of electrolyte during MO EF oxidation using different G5-PTFE
cathodes. The initial MO concentration was ~25 mg/L. EF oxidation was carried out in 100 mL
of 0.05 M Na2SO4 solution at pH of 3 with continuous stirring and air bubbling.

6.3.4 MO adsorption before and after electro-Fenton regeneration

The adsorption kinetics of the G5-PTFE electrodes before and after EF regeneration are

shown in Figure 6-8. With the increase in G5 to PTFE ratio, the MO adsorption rate and the

adsorbed MO increased for both fresh and regenerated G5-PTFE cathodes. After 2 days

adsorption, G5-PTFE-7-1 had the highest adsorbed MO of 176 mg/g (normalized by the carbon

mass in the cathode), which was 76% of the saturation adsorption capacity (240 mg/g) of G5 AC.

135
G5-PTFE-1-1 had the lowest MO adsorption of 23 mg/g after 2 days. The low adsorption

capacities of G5-PTFE-1-1 and G5-PTFE-2-1 indicated that they were not suitable as adsorbents

for MO removal from water. Reasonable adsorption capacity was obtained (> ~100 mg/g) after

two days when the G5: PTFE ratio was larger than 3, i.e. when the electrode was hydrophilic

(contact angle < 90 °).

(a) (b)

(c) (d)

Figure 6-8. MO adsorption kinetics of fresh (a) and regenerated (b) G5-PTFE electrodes. The
lines represent WM model fitting. Adsorption kinetics of fresh (c) and regenerated (d) G5-PTFE-
7-1 were also fitted to a two-step WM model as well. Adsorption conditions: 100 mL of ~300
mg/L MO solution, 25 °C, and 250 rpm, for maximum 2 days. The adsorbed amount was
normalized by the mass of carbon in the electrodes. Regeneration was carried out in 100 mL 0.05

136
M Na2SO4 solution in presence of 0.1 mM Fe2+ for 8 h under room temperature (~20 °C) by
controlling the cathodic potential of -0.8 V (vs Ag/AgCl).

The Weber-Morris diffusion model was applied to the adsorption kinetics in Figure 6-8. The

kinetics data obtained over 2 days adsorption was well fitted by the WM model for most of the

electrodes. The data obtained with the G5-PTFE-7-1 electrode was not as well fitted by the

model. However, it was better fitted to a two-step WM model (Figure 6-8c and d). This result

indicated that the MO adsorption was controlled by the interparticle diffusion in the first stage

(0-8 h) and then was limited by the intraparticle diffusion at longer time (8-48 h) for the G5-

PTFE-7-1 electrode [186]. The MO adsorption for all the other electrodes, however, was within

the interparticle diffusion control range over 2 days adsorption. The WM interparticle diffusion

rate constant increased with the decrease in PTFE content in the electrode. As indicated in the

SEM images (Figure 6-3), PTFE was located on the out surface of the carbon particles, which

can affect the transportation of water and water soluble species among the carbon particles [33].

The interparticle diffusion rate constants for the different electrodes increased linearly with the

hydrophilicity of the electrodes (represented by the cosine of the contact angle) as shown in

Figure 6-9. As nonvolatile and soluble species, MO molecules diffused and transported within

the pores together with water molecules. However, the penetration of water into a packed powder

sample is controlled by capillary forces and has a linear relationship with the square root of time

(√𝑡) as indicated by the Washburn equation [234], which has similar expression as WM model.

𝑚 = 𝐶√cos 𝜃 𝑡 (6-1)

where 𝑚 is the mass of a penetrating liquid into the packed powder sample at time 𝑡, cos 𝜃

represents the wettability of the packed powder sample, and 𝐶 is a constant related to packing

parameters of the powder sample. This implied that the penetration or diffusion of water into the
137
electrodes, which is determined by the electrode hydrophilicity, controlled the adsorption of MO

molecules from water. Presumably, water cannot penetrate a hydrophobic material according to

the Washburn equation [234], but it has been reported that a hydrophobic surface can still adsorb

water because of the defects at the atomic/nanometer scales [240]. Nevertheless, the

hydrophilicity of the electrodes can still affect the MO adsorption in macroscopic range as shown

in Figure 6-9.

Figure 6-9. The relationship between WM diffusion rate constant and G5-PTFE electrodes
hydrophilicity (represented by cosθ). Solid lines are linear fitting of the data points.

After regeneration, the MO adsorption capacity decreased (Figure 6-8), however, as depicted

in Figure 6-10, the regeneration efficiency increased with the increasing electrode hydrophilicity.

The highest regeneration efficiency of 81% was obtained with the G5-PTFE-7-1 electrode, while

G5-PTFE-1-1 had the lowest regeneration efficiency of 49%. The regeneration efficiency

increased by 21% from 49% to 70% when the contact angle with water for the electrode

decreased from 121° to 88°. After the electrode became hydrophilic, however, the regeneration
138
efficiency only increased by 11% from 70% to 81% when contact angle decreased from 88° to

47°.

Figure 6-10. The relationship between EF regeneration efficiency and G5-PTFE electrode
hydrophilicity. The MO adsorption experiments for both fresh and regenerated electrodes were
carried out at 25 °C for 2 days.

On the other hand, the regeneration efficiency might also be affected by the electrical

conductivity of the G5-PTFE electrodes due to different G5: PTFE ratios. According to the

electro-desorption mechanism, less potential drop would be applied for the desorption of ions if

the electrode was less conductive [241, 242]. However, as indicated in the literature, the

conductivity of the carbon-PTFE electrodes was similar (< 50% variation) if the carbon content

was within the range of 50-80% [241-243], which was the case for this study. As a consequence,

the double layer capacitances, which represents the adsorption and desorption of ions, were

similar (< 10%) among those electrodes [241, 242]. Therefore, the higher electrical conductivity

139
of the electrodes only partially (< 1%, based on the adsorption capacity of regenerated electrode)

contributed to the higher regeneration efficiency.

6.3.5 Effect of adsorption time

To investigate the impact of adsorption time on the regeneration, G5-PTFE-1-1 and G5-

PTFE-3-1 were used for MO adsorption over longer times (6 days), and G5-PTFE-7-1 was used

for MO adsorption for shorter times (8 h). As depicted in Figure 6-11a, the G5-PTFE-1-1 and

G5-PTFE-3-1 electrodes continued to adsorb MO from the water solution after 2 days. After 6

days, the adsorption capacity of G5-PTFE-1-1 and G5-PTFE-3-1 increased by 22.4 mg/g and

49.4 mg/g to 52.4 mg/g and 152.4 mg/g, respectively. In contrast, the MO adsorption capacity of

G5-PTFE-7-1 reached 97.2 mg/g after only 8 h adsorption (Figure 6-11b). These results further

demonstrated that higher hydrophilicity improved the rate of MO adsorption.

(a) (b)

Figure 6-11. MO adsorption kinetics G5-PTFE electrodes for different adsorption times. The
solid lines are the nonlinear fitting of WM model. The electrodes were put into 100 mL of ~300
mg/L MO solution in a shaker at 25 °C and 250 rpm for 6 days (a) and 8 h (b). The adsorbed
amount was normalized by the mass of carbon in the electrodes. The sample name with the
prefix “R-” was the regenerated carbon.

140
The adsorption kinetics were also fitted by the WM model (Figure 6-11). The diffusion rate

constants for G5-PTFE-1-1 and G5-PTFE-3-1 were 3.9 mg/ (L h0.5) and 13.4 mg/ (L h0.5),

respectively. These rates were similar to those obtained from the 2-day adsorption experiments,

indicating the diffusion limitation was constant within the 6 days adsorption for these two

electrodes. The diffusion rate constant for G5-PTFE-7-1 was 35.1 mg/ (L h0.5) over 8 h compared

to 28.2 mg/ (L h0.5) for adsorption over 48 h. The diffusion limitations may have increased over

time because the intraparticle diffusion (mesopore or micropore diffusion) may be dominant at

longer adsorption time [183, 185, 186], leading to a lower diffusion rate constant.

After EF regeneration, the adsorption capacities were consistently lower than the fresh

electrodes (Figure 6-11) regardless of the length of the adsorption experiment. For G5-PTFE-1-1,

changing the adsorption time from 2 days to 6 days, the regeneration efficiencies essentially

remained the same (49 ± 9% vs 47.5%). In contrast, the regeneration efficiency for G5-PTFE-7-1

increased slightly from 81 ± 2% to 84 ± 2% when the adsorption time decreased from 2 days to 8

h. The regeneration efficiency of G5-PTFE-3-1 after EF regeneration increased from 70 ± 4% to

80% when the adsorption time increased from 2 to 6 days. Although more MO was loaded on the

G5-PTFE electrodes by increasing adsorption time, the regeneration efficiency did not decrease

after the same EF regeneration process. The slight impact of initial MO loading on the

regeneration efficiency was also observed in other research [104]. The results indicated that the

irreversibly adsorbed MO (adsorbed at longer time) can also be regenerated by EF oxidation

under the typical conditions.

141
6.3.6 Regeneration cycles

As indicated above, the hydrophilicity, MO adsorption rate, and regeneration efficiency of the

G5-PTFE-7-1 electrode were the highest among all the prepared electrodes. Therefore, this

electrode was employed for several adsorption-regeneration cycles with adsorption over 2 days

and EF regeneration. As shown in Figure 6-12, the MO adsorption capacity after 2 days

adsorption decreased from 176 mg/g in the first adsorption to 143 mg/g in the second adsorption

and then gradually decreased to 112 mg/g in the fifth adsorption. The calculated regeneration

efficiencies were 81%, 74%, 71%, and 64%, respectively.

Figure 6-12. MO adsorption kinetics of G5-PTFE-7-1 in different adsorption-regeneration


cycles. Fresh electrolyte was supplied for each regeneration process.

The continuous decrease in adsorption capacity and regeneration efficiency indicated that

some organic pollutants remained on the electrode after each regeneration process. This was

testified by the regeneration of saturated electrode in the fourth regeneration cycle. Before the

fourth regeneration, the electrode after 2-day adsorption (fourth adsorption) was used for re-
142
adsorption of MO under same conditions for another 4 days to be saturated. 53.8 mg/g more of

MO was adsorbed during this process with a total adsorption capacity of 178.0 mg/g. However,

in the fifth adsorption, the saturated MO adsorption capacity was only 165.7 mg/g, indicating the

remaining of organic pollutants. The adsorption kinetics (23-26 mg/(g h0.5)) and adsorption

capacities (65-72 mg/g) of the regenerated electrodes, however, were similar within the first 8 h,

and the main differences occurred at longer times. These results demonstrated that the remaining

organic pollutants were mainly located at places that is more difficult for MO to reach. Despite

the decrease in regeneration efficiency, after these adsorption and regeneration processes, the

total adsorbed MO was 740 mg/g based on the G5 weight mass, which was 3.1 times of the

saturation adsorption capacity of G5. During these processes, only 1829 C of applied charge was

consumed (Table 6-3), which was much lower than those reported in the literature [24, 104].

Table 6-3. The applied charge for the regeneration of G5-PTFE-7-1 for each cycle when fresh
electrolyte was supplied for each cycle. MO was adsorbed at 25 °C for 2 days. Regeneration was
carried out in 100 mL of 0.05 M Na2SO4 at pH 3 and room temperature.

Regeneration Regeneration conditions Applied Regeneration


cycle number Potential (V) Time (h) charge (C) efficiency (%)
1 -0.8 8 351 81
2 -0.8 8 482 74
3 -0.8 8 497 71
4 -0.8 8 499 64

Instead of supplying fresh electrolyte for each regeneration cycle, the electrolyte was reused

without any adjustment for the G5-PTFE-7-1 electrode (Figure 6-13). During these processes,

adsorption was carried out for 8 h rather than 2 days. As depicted in Figure 6-13, the MO

adsorption capacity decreased gradually from 82 mg/g in the second adsorption cycle to 66 mg/g

in the fifth adsorption cycle. The corresponding regeneration efficiencies of the first four
143
regeneration cycles were 84%, 77%, 72%, and 68%, respectively. These values were similar to

those obtained when fresh electrolyte was supplied for each regeneration, implying that the reuse

of electrolyte did not affect the regeneration process significantly.

Figure 6-13. Adsorption-regeneration cycles of G5-PTFE-7-1 without changing electrolyte.


Adsorption was carried out for 8 h. Unless otherwise indicated, the regeneration was carried out
at a cathodic potential of -0.8 V for 8 h. The electrolyte was reused for every regeneration cycle
without any adjustment.

The regeneration conditions were varied to recover more adsorption sites. As indicated in

Figure 6-13, the adsorption capacity was not improved by increasing the cathodic potential to 1.6

V and regenerating for 4 h in the fifth and sixth regeneration cycles. On the contrary, the

adsorption capacity decreased significantly to 42 mg/g. After these two regenerations, some red

materials were deposited on the cathode surface (Figure 6-14), which may be the precipitation of

Fe(OH)3 due to the high local pH at high cathodic potential [29, 146, 244, 245]. These iron

144
species deposits could block the pores and result in a lower adsorption capacity because the iron

hydroxide precipitate had a low azo dye adsorption capacity [245].

(a)

(b)

Figure 6-14. Deposition of iron species (red) on the G5-PTFE-7-1 cathode after the sixth (a) and
eighth (b) regeneration cycles.

The adsorption capacity was recovered to 64 mg/g after the seventh regeneration cycle,

however, under the typical conditions (i.e. -0.8 V for 8 h). The MO adsorption capacity

decreased again after the eighth regeneration cycle under a cathodic potential of -3.0 V for 0.5 h

because of iron precipitation (Figure 6-14). Further regeneration under the cathodic potential of -

0.8 V did not recover the adsorption capacity of the G5-PTFE-7-1 electrodes in the ninth, tenth,

and eleventh regeneration cycles, even with longer regeneration time. The UV-Vis spectra of the

electrolyte after each regeneration cycle are shown in Figure 6-15, and the UV-Vis absorbance

increased continuously in the first four regeneration cycles, indicating the accumulation of

desorbed organic pollutants especially under the typical regeneration conditions. This result

implied that the EF oxidation under the conditions used was not sufficient to degrade all the

desorbed organic pollutants. Meanwhile, even though the regeneration efficiency was not

improved, an increase in the cathodic potential and regeneration time reduced the remaining

organic pollutants in the solution as indicated by the lower UV-Vis absorbance of the electrolyte

145
after the fifth, sixth, and eleventh regeneration cycle (Figure 6-15b, d). Nonetheless, adsorption

capacity of > 40 mg/g was maintained after 12 adsorption-regeneration cycles without changing

the electrolyte.

(a) (b)

(c) (d)

Figure 6-15. The UV-Vis spectra of the electrolyte after each regeneration of G5-PTFE-7-1 when
the electrolyte was reused in each cycle. MO was adsorbed at 25 °C for 8 h. Regeneration was
carried out in 100 mL of 0.05 M Na2SO4 at pH 3 and room temperature.

After 11 adsorption cycles, the total removed MO by the G5-PTFE-7-1 cathode was 688

mg/g, which was ~3 times the saturation adsorption capacity of MO on G5 carbon. To achieve

this, only 5673 C of charge (Table 6-4) was applied in the process and only one batch of

146
electrolyte was used. The applied charge was similar to those reported in the literature for just

one cathodic regeneration cycle with regeneration efficiency of ~75% [104, 205], which

indicated the lower energy consumption of this technology. As an advantage, the electrode was

one piece of material so that it could be easily transferred between the adsorption process and the

regeneration process. A reasonable regeneration efficiency in several regeneration cycles, a low

energy consumption, and the ease of operation make the adsorption-EF regeneration technology

using AC-PTFE electrode feasible for practical applications.

Table 6-4. The applied charge for the regeneration of G5-PTFE-7-1 for each cycle when the
electrolyte was reused. MO was adsorbed at 25 °C for 8 h. The first regeneration was carried out
in 100 mL of 0.05 M Na2SO4 at pH 3 and room temperature, and the following regenerations
were carried out without adjusting the electrolyte.

Regeneration cycle Regeneration conditions Applied charge Regeneration


number Potential (V) Time (h) (C) efficiency (%)
1 -0.8 8 355 84.5
2 -0.8 8 419 76.8
3 -0.8 8 477 71.8
4 -0.8 8 532 68.3
5 -1.6 4 952 61.5
6 -1.6 4 996 43.6
7 -0.8 8 669 65.4
8 -3.0 0.5 409 49.3
9 -0.8 2 174 40.8
10 -0.8 8 690 46.4
11 -0.8 16 1102 43.2

6.4 Conclusions

This chapter tested the application of G5-PTFE materials as both the adsorbent for MO

adsorption and the cathode for EF regeneration. Although, introducing of PTFE did not

significantly change the pore structure and surface chemistry of the G5 AC, the increase in PTFE

content in the electrodes decreased their hydrophilicity. The MO adsorption rate constant of the
147
G5-PTFE electrodes decreased linearly with decreasing hydrophilicity. Higher hydrophilicity

also led to higher EF regeneration efficiency at a cathodic potential of -0.8 V (vs Ag/AgCl) for 8

h. The highest regeneration efficiency of 81% was obtained by using the G5-PTFE-7-1 electrode.

The hydrophilicity of the electrodes controlled the adsorption and regeneration processes by

affecting the diffusion rate of the water-soluble organic pollutants. Overall, the electrode should

be hydrophilic, i.e. contact angle of water < 90°, to obtain better adsorption capacities and

regeneration efficiencies. In addition, the initial adsorption capacity had small effect on the

regeneration efficiency, and the possible reason was that irreversible adsorbed MO at longer time

can also be regenerated under EF regeneration. The G5-PTFE-7-1 electrode could be regenerated

for 11 cycles without changing the electrolyte.

148
Chapter Seven: Conclusions and Recommendations

7.1 Conclusions

This thesis focused on the regeneration of carbon materials saturated with organic pollutants

by using Fenton oxidation based technologies, including heterogeneous Fenton oxidation and EF

oxidation. The effects of carbon material properties as well as regeneration parameters were

determined.

The pore structure of the carbon materials had a crucial influence on the adsorption and

heterogeneous Fenton oxidation regeneration process. The prepared magnetic carbon materials,

Fe2O3/SMC, Fe2O3/G5, and Fe2O3/NORIT, had similar iron oxide loading (~20 wt%) and iron

oxide particle size (~20 nm) but different pore size distribution. The microporosities (percentage

of micropore volume) of Fe2O3/SMC, Fe2O3/G5, and Fe2O3/NORIT were 4.5%, 55%, and 75%,

respectively, and their surface areas increased accordingly. The MO adsorption capacity of the

carbon materials increased with higher surface area. After Fenton oxidation regeneration, the

MO adsorption capacities decreased for Fe2O3/G5 and Fe2O3/NORIT indicating incomplete

regeneration for them, whereas essentially complete regeneration was achieved on the

mesoporous carbon material (Fe2O3/SMC). The MO adsorption kinetics and isotherms of the

fresh and regenerated carbon materials indicated that the MO was less favorably adsorbed after

Fenton oxidation. In addition, the regeneration efficiency was inversely proportional to

microporosity - the higher the fresh sample microporosity, the less effective the Fenton oxidation

regeneration. The regeneration efficiency of Fenton oxidation was also linearly related to the

mesopore diffusion rate constant derived from the WM model. These results were consistent

with the MO binding strongly in the micropores and not being removed during the Fenton

149
oxidation regeneration. Thus, there is a trade-off between adsorption capacity (high on

microporous material) and regeneration efficiency (high on mesoporous material).

Given the understanding that Fenton oxidation was ineffective in regenerating microporous

materials, the EF method was proposed to regenerate carbon materials saturated with MO using

graphite felt as the cathode. The EF regeneration was affected by the EF parameters and the

carbon materials pore structure. The regeneration efficiency of EF regeneration was improved

with a higher cathodic potential, better contact with the cathode, and larger pores with the carbon

adsorbent. The application of a more oxidative anode, the presence of air flow, and loading of

iron oxide, however, did not consistently improve the regeneration efficiency. In addition, the

regeneration kinetics study indicated that most regeneration of the carbon materials happened in

the first 30 min. On the other hand, compared with the Fenton oxidation regeneration process,

the regeneration efficiency for Fe2O3/G5 and Fe2O3/NORIT was improved with the presence of a

cathodic potential. The regeneration efficiency was linearly correlated to the recovery percentage

of micropore volume – the higher the micropore volume recovery, the higher the regeneration

efficiency. This result indicated that EF regeneration was capable of regenerating adsorption

sites within micropores. In addition, MB (cation dye) saturated carbon was also regenerated by

EF oxidation with lower regeneration efficiency compared with MO saturated ones. The results

were consistent with electro-desorption, originating from cathodic polarization and/or pH

induced desorption as the main pathway for the EF regeneration. Nevertheless, the reduction and

oxidation of MO was observed, which may also contribute to the regeneration and final

mineralization of the organic pollutants.

150
As demonstrated by the EF regeneration experiments using a graphite felt cathode, higher

regeneration efficiency was observed when the adsorbent had better contact with the cathode.

Based on this, G5-PTFE electrodes with different G5: PTFE ratios were used as both the

adsorbent during adsorption and the cathode during EF regeneration. The G5-PTFE preparation

process did not introduce new functional groups to the G5 and had small effect on the pore

structure (based on G5). However, the hydrophilicity of the electrodes decreased with increasing

PTFE content. The adsorption kinetics of the fresh and regenerated G5-PTFE electrodes were

well fitted by the WM model. The WM diffusion rate constant had a linear relationship with the

electrode hydrophilicity - the higher the hydrophilicity, the higher the diffusion rate constant. In

addition, the regeneration efficiency also increased with the increase in hydrophilicity. The

diffusion of MO and other compounds was controlled by the hydrophilicity of the electrodes and

thus resulted in the lower adsorption kinetics and lower regeneration efficiency. On the other

hand, increasing the initial MO loading with longer adsorption time did not affect the

regeneration efficiency, which indicated that the irreversibly adsorbed MO can also be

regenerated by the EF oxidation. The G5-PTFE-7-1 electrode with the highest hydrophilicity was

applied for 11 adsorption-regeneration cycles with or without changing the electrolyte for each

regeneration process. Over these experiments, more than 3 times the amount of MO was

adsorbed compared with a single use of G5, i.e. 5 adsorption cycles (2-day) by changing

electrolyte and in 10 adsorption cycles (8-hour) without changing electrolyte.

Overall, the Fenton oxidation and EF oxidation are capable of regenerating carbon materials.

The carbon materials can be used in different configurations. A mesoporous carbon that can be

readily separated from a solution can be placed directly in solution, both Fenton oxidation and

EF oxidation using the graphite felt cathode can be applied for the regeneration. If a separable
151
carbon material possesses mainly micropores, EF oxidation using the graphite felt cathode can be

used for regeneration. However, if the carbon material is not separable, it can be incorporated in

a hydrophilic AC-PTFE electrode and applied for adsorption. Then, EF oxidation using the AC-

PTFE electrode as the cathode can be applied for regeneration.

7.2 Recommendations for future research

Considering the limitations of this thesis, other research and experiments should be conducted

to get a comprehensive understanding the AOP regeneration technologies and use them in a

practical project for wastewater treatment.

(1) In this thesis, the solution after regeneration was only analyzed by the UV-Vis

spectrometer. Only limited information can be obtained from this method. In order to understand

the regeneration process more deeply, other analysis techniques, including high-performance

liquid chromatography (HPLC), liquid chromatography-mass spectrometry (LC-MS), and gas

chromatography-mass spectrometry (GC-MS) are proposed to determine the intermediates

during the regeneration process.

(2) Except the pore structure properties, the regenerated carbon materials were not tested for

other properties in this research. In particular, the point of zero charge of the adsorbents is of

great interest for understanding the mechanism of the second adsorption process.

(3) Most of the experiments were conducted using MO as the model organic pollutant. Other

organic pollutants should be applied to investigate the impact of some adsorbate properties on

the regeneration process. The emerging organic pollutants, such as perfluorooctanesulfonic acid,

perfluorooctanoic acid, pharmaceuticals and personal care products, endocrine disrupting

chemicals, can be used to evaluate the performance of the regeneration process.


152
(4) On the other hand, the impact of other organic pollutants, natural organic matter, and

inorganic ions on the adsorption and regeneration processes would be necessary for the

application of this technology in wastewater treatment. The influence of adsorption process

through either a batch or a flow-through mode on the adsorption and regeneration processes

should also be investigated. Experiments should also be carried out with real wastewater to test

the feasibility of the adsorption-regeneration process in a practical project.

(5) As concluded in this thesis, the application of an AC-PTFE electrode as both an adsorbent

for adsorption and the cathode for EF regeneration was controlled by the hydrophilicity. The

only parameter that was studied in this thesis was the AC: PTFE ratio. Other options, such as the

addition of zeolites or other metal oxides, could be applied to tune the wettability of the

electrode.

(6) During the EF regeneration of carbon materials in this thesis, the anode process was not

studied. However, as indicated in the literature [22, 99, 178], anodic oxidation process could also

be used for the regeneration. Therefore, application of the AC-PTFE electrodes as both the

cathode and the anode during the regeneration process could be beneficial to the regeneration by

saving energy consumption.

(7) Except for the Fenton oxidation and the EF oxidation, other AOP can also be investigated

and applied for the regeneration of carbon saturated with organic pollutants. Photocatalysis and

photoelectron-Fenton would be recommended. In the photocatalysis process, the regeneration is

possible to achieve without adding any chemicals or external power except the photocatalyst.

Some nitrogen-doped carbon materials were reported to be effective photocatalyst even under

visible light irradiation [246]. It is possible to use these materials as adsorbent which can be

153
regenerated under solar light. On the other hand, the irradiation of UV light during EF can

improve the decomposition of Fe(III)-carboxylate complexes [129, 130], which may be

beneficial to the regeneration of carbon materials during EF regeneration.

154
References

1. Pliego, G., J.A. Zazo, P. Garcia-Muñoz, M. Munoz, J.A. Casas, and J.J. Rodriguez,
Trends in the Intensification of the Fenton Process for Wastewater Treatment: An
Overview. Critical Reviews in Environmental Science and Technology, 2015. 45(24): p.
2611-2692.

2. Martínez-Huitle, C.A. and E. Brillas, Decontamination of wastewaters containing


synthetic organic dyes by electrochemical methods: A general review. Applied Catalysis
B: Environmental, 2009. 87(3–4): p. 105-145.

3. Brillas, E. and C.A. Martínez-Huitle, Decontamination of wastewaters containing


synthetic organic dyes by electrochemical methods. An updated review. Applied Catalysis
B: Environmental, 2015. 166–167: p. 603-643.

4. Malato, S., P. Fernández-Ibáñez, M.I. Maldonado, J. Blanco, and W. Gernjak,


Decontamination and disinfection of water by solar photocatalysis: Recent overview and
trends. Catalysis Today, 2009. 147(1): p. 1-59.

5. Chong, M.N., B. Jin, C.W.K. Chow, and C. Saint, Recent developments in photocatalytic
water treatment technology: A review. Water Research, 2010. 44(10): p. 2997-3027.

6. Adewuyi, Y.G., Sonochemistry in Environmental Remediation. 1. Combinative and


Hybrid Sonophotochemical Oxidation Processes for the Treatment of Pollutants in
Water. Environmental Science & Technology, 2005. 39(10): p. 3409-3420.

7. Pan, B. and B. Xing, Adsorption Mechanisms of Organic Chemicals on Carbon


Nanotubes. Environmental Science & Technology, 2008. 42(24): p. 9005-9013.

8. Lin, S.-H. and R.-S. Juang, Adsorption of phenol and its derivatives from water using
synthetic resins and low-cost natural adsorbents: A review. Journal of Environmental
Management, 2009. 90(3): p. 1336-1349.

9. Gupta, V.K. and Suhas, Application of low-cost adsorbents for dye removal – A review.
Journal of Environmental Management, 2009. 90(8): p. 2313-2342.

10. Foo, K.Y. and B.H. Hameed, Insights into the modeling of adsorption isotherm systems.
Chemical Engineering Journal, 2010. 156(1): p. 2-10.

11. Fallou, H., N. Cimetière, S. Giraudet, D. Wolbert, and P. Le Cloirec, Adsorption of


pharmaceuticals onto activated carbon fiber cloths – Modeling and extrapolation of
adsorption isotherms at very low concentrations. Journal of Environmental Management,
2016. 166: p. 544-555.

12. Yin, C.Y., M.K. Aroua, and W.M.A.W. Daud, Review of modifications of activated
carbon for enhancing contaminant uptakes from aqueous solutions. Separation and
Purification Technology, 2007. 52(3): p. 403-415.
155
13. Salvador, F., N. Martin-Sanchez, R. Sanchez-Hernandez, M.J. Sanchez-Montero, and C.
Izquierdo, Regeneration of carbonaceous adsorbents. Part II: Chemical, Microbiological
and Vacuum Regeneration. Microporous and Mesoporous Materials, 2015. 202: p. 277-
296.

14. USEPA, Wastewater Technology Fact Sheet: Granular Activated Carbon Adsorption and
Regeneration. 2000, U.S. EPA: Washington, D.C.

15. Inyang, M. and E. Dickenson, The potential role of biochar in the removal of organic and
microbial contaminants from potable and reuse water: A review. Chemosphere, 2015.
134: p. 232-240.

16. Salvador, F., N. Martin-Sanchez, R. Sanchez-Hernandez, M.J. Sanchez-Montero, and C.


Izquierdo, Regeneration of carbonaceous adsorbents. Part I: Thermal Regeneration.
Microporous and Mesoporous Materials, 2015. 202: p. 259-276.

17. Narbaitz, R.M. and J. McEwen, Electrochemical regeneration of field spent GAC from
two water treatment plants. Water Research, 2012. 46(15): p. 4852-4860.

18. Zanella, O., I.C. Tessaro, and L.A. Féris, Desorption- and Decomposition-Based
Techniques for the Regeneration of Activated Carbon. Chemical Engineering &
Technology, 2014. 37(9): p. 1447-1459.

19. Andreozzi, R., V. Caprio, A. Insola, and R. Marotta, Advanced oxidation processes
(AOP) for water purification and recovery. Catalysis Today, 1999. 53(1): p. 51-59.

20. Moreira, F.C., R.A.R. Boaventura, E. Brillas, and V.J.P. Vilar, Electrochemical advanced
oxidation processes: A review on their application to synthetic and real wastewaters.
Applied Catalysis B: Environmental, 2017. 202: p. 217-261.

21. Huling, S.G., P.K. Jones, and T.R. Lee, Iron Optimization for Fenton-Driven Oxidation
of MTBE-Spent Granular Activated Carbon. Environmental Science & Technology,
2007. 41(11): p. 4090-4096.

22. Sharif, F., L.R. Gagnon, S. Mulmi, and E.P.L. Roberts, Electrochemical regeneration of
a reduced graphene oxide/magnetite composite adsorbent loaded with methylene blue.
Water Research, 2017. 114(Supplement C): p. 237-245.

23. Bañuelos, J.A., F.J. Rodríguez, J. Manríquez Rocha, E. Bustos, A. Rodríguez, J.C. Cruz,
L.G. Arriaga, and L.A. Godínez, Novel Electro-Fenton Approach for Regeneration of
Activated Carbon. Environmental Science & Technology, 2013. 47(14): p. 7927-7933.

24. Zhan, J., H. Wang, X. Pan, J. Wang, G. Yu, S. Deng, J. Huang, B. Wang, and Y. Wang,
Simultaneous regeneration of p-nitrophenol-saturated activated carbon fiber and
mineralization of desorbed pollutants by electro-peroxone process. Carbon, 2016. 101: p.
399-408.

156
25. Muranaka, C.T., C. Julcour, A.-M. Wilhelm, H. Delmas, and C.A.O. Nascimento,
Regeneration of Activated Carbon by (Photo)-Fenton Oxidation. Industrial &
Engineering Chemistry Research, 2010. 49(3): p. 989-995.

26. Qin, Y., M. Long, B. Tan, and B. Zhou, RhB Adsorption Performance of Magnetic
Adsorbent Fe3O4/RGO Composite and Its Regeneration through A Fenton-like Reaction.
Nano-Micro Letters, 2014. 6(2): p. 125-135.

27. Do, M.H., N.H. Phan, T.D. Nguyen, T.T.S. Pham, V.K. Nguyen, T.T.T. Vu, and T.K.P.
Nguyen, Activated carbon/Fe3O4 nanoparticle composite: Fabrication, methyl orange
removal and regeneration by hydrogen peroxide. Chemosphere, 2011. 85(8): p. 1269-
1276.

28. Huling, S.G., E. Kan, and C. Wingo, Fenton-driven regeneration of MTBE-spent


granular activated carbon—Effects of particle size and iron amendment procedures.
Applied Catalysis B: Environmental, 2009. 89(3–4): p. 651-658.

29. Brillas, E., I. Sirés, and M.A. Oturan, Electro-Fenton Process and Related
Electrochemical Technologies Based on Fenton’s Reaction Chemistry. Chemical
Reviews, 2009. 109(12): p. 6570-6631.

30. Narbaitz, R.M. and J. Cen, Electrochemical regeneration of granular activated carbon.
Water Research, 1994. 28(8): p. 1771-1778.

31. Ribeiro, R.S., A.M.T. Silva, J.L. Figueiredo, J.L. Faria, and H.T. Gomes, Catalytic wet
peroxide oxidation: a route towards the application of hybrid magnetic carbon
nanocomposites for the degradation of organic pollutants. A review. Applied Catalysis B:
Environmental, 2016. 187: p. 428-460.

32. Huang, Q., S. Deng, D. Shan, Y. Wang, B. Wang, J. Huang, and G. Yu, Enhanced
adsorption of diclofenac sodium on the carbon nanotubes-polytetrafluorethylene
electrode and subsequent degradation by electro-peroxone treatment. Journal of Colloid
and Interface Science, 2017. 488(Supplement C): p. 142-148.

33. Dong, H., H. Yu, H. Yu, N. Gao, and X. Wang, Enhanced performance of activated
carbon–polytetrafluoroethylene air-cathode by avoidance of sintering on catalyst layer in
microbial fuel cells. Journal of Power Sources, 2013. 232: p. 132-138.

34. Brack, W., V. Dulio, M. Ågerstrand, I. Allan, R. Altenburger, M. Brinkmann, D. Bunke,


R.M. Burgess, I. Cousins, B.I. Escher, F.J. Hernández, L.M. Hewitt, K. Hilscherová, J.
Hollender, H. Hollert, R. Kase, B. Klauer, C. Lindim, D.L. Herráez, C. Miège, J. Munthe,
S. O'Toole, L. Posthuma, H. Rüdel, R.B. Schäfer, M. Sengl, F. Smedes, D. van de Meent,
P.J. van den Brink, J. van Gils, A.P. van Wezel, A.D. Vethaak, E. Vermeirssen, P.C. von
der Ohe, and B. Vrana, Towards the review of the European Union Water Framework
Directive: Recommendations for more efficient assessment and management of chemical
contamination in European surface water resources. Science of The Total Environment,
2017. 576(Supplement C): p. 720-737.
157
35. Ahad, J.M.E., H. Pakdel, M.M. Savard, A.I. Calderhead, P.R. Gammon, A. Rivera, K.M.
Peru, and J.V. Headley, Characterization and Quantification of Mining-Related
“Naphthenic Acids” in Groundwater near a Major Oil Sands Tailings Pond.
Environmental Science & Technology, 2013. 47(10): p. 5023-5030.

36. Yamashita, N., K. Kannan, S. Taniyasu, Y. Horii, G. Petrick, and T. Gamo, A global
survey of perfluorinated acids in oceans. Marine Pollution Bulletin, 2005. 51(8): p. 658-
668.

37. Chen, M., K. Ohman, C. Metcalfe, M.G. Ikonomou, P.L. Amatya, and J. Wilson,
Pharmaceuticals and endocrine disruptors in wastewater treatment effluents and in the
water supply system of Calgary, Alberta, Canada. Water Quality Research Journal of
Canada, 2006. 41(4): p. 351-364.

38. Kolpin, D.W., E.T. Furlong, M.T. Meyer, E.M. Thurman, S.D. Zaugg, L.B. Barber, and
H.T. Buxton, Pharmaceuticals, hormones, and other organic wastewater contaminants in
US streams, 1999− 2000: A national reconnaissance. Environmental science &
technology, 2002. 36(6): p. 1202-1211.

39. USEPA. Persistent Organic Pollutants: A Global Issue, A Global Response. [cited 2017
October 27]; Available from: https://www.epa.gov/international-cooperation/persistent-
organic-pollutants-global-issue-global-response.

40. Clemente, J.S. and P.M. Fedorak, A review of the occurrence, analyses, toxicity, and
biodegradation of naphthenic acids. Chemosphere, 2005. 60(5): p. 585-600.

41. Terzaghi, E., E. Zanardini, C. Morosini, G. Raspa, S. Borin, F. Mapelli, L. Vergani, and
A. Di Guardo, Rhizoremediation half-lives of PCBs: Role of congener composition,
organic carbon forms, bioavailability, microbial activity, plant species and soil
conditions, on the prediction of fate and persistence in soil. Science of The Total
Environment, 2018. 612: p. 544-560.

42. Kannan, K., J. Koistinen, K. Beckmen, T. Evans, J.F. Gorzelany, K.J. Hansen, P.D.
Jones, E. Helle, M. Nyman, and J.P. Giesy, Accumulation of perfluorooctane sulfonate in
marine mammals. Environmental science & technology, 2001. 35(8): p. 1593-1598.

43. Lofrano, G., S. Meriç, G.E. Zengin, and D. Orhon, Chemical and biological treatment
technologies for leather tannery chemicals and wastewaters: A review. Science of The
Total Environment, 2013. 461-462: p. 265-281.

44. de Morais, J.L. and P.P. Zamora, Use of advanced oxidation processes to improve the
biodegradability of mature landfill leachates. Journal of Hazardous Materials, 2005.
123(1): p. 181-186.

45. IUPAC, Compendium of Chemical Terminology. 2nd ed. 1997, Oxford: Blackwell
Scientific Publications.

158
46. Langmuir, I., The Constitution and Fundmental Properties of Solids and Liquids. Part I.
Solids. Journal of the American Chemical Society, 1916. 38(11): p. 2221-2295.

47. Freundlich, H., Over the adsorption in solution. The Journal of Physical Chemistry, 1906.
57: p. 1100-1107.

48. Qiu, H., L. Lv, B.-c. Pan, Q.-j. Zhang, W.-m. Zhang, and Q.-x. Zhang, Critical review in
adsorption kinetic models. Journal of Zhejiang University SCIENCE A, 2009. 10(5): p.
716-724.

49. Largergren, S., Zur theorie der sogenannten adsorption geloster stoffe. Kungliga Svenska
Vetenskapsakademiens Handlingar, 1898. 24: p. 1-39.

50. Ho, Y.S. and G. McKay, A Comparison of Chemisorption Kinetic Models Applied to
Pollutant Removal on Various Sorbents. Process Safety and Environmental Protection,
1998. 76(4): p. 332-340.

51. Weber, W.J. and J.C. Morris, Kinetics of adsorption on carbon from solution. Journal of
the Sanitary Engineering Division, 1963. 89(2): p. 31-60.

52. Smith, S.C. and D.F. Rodrigues, Carbon-based nanomaterials for removal of chemical
and biological contaminants from water: A review of mechanisms and applications.
Carbon, 2015. 91: p. 122-143.

53. Mohan, D., A. Sarswat, Y.S. Ok, and C.U. Pittman Jr, Organic and inorganic
contaminants removal from water with biochar, a renewable, low cost and sustainable
adsorbent – A critical review. Bioresource Technology, 2014. 160: p. 191-202.

54. Rossner, A., S.A. Snyder, and D.R.U. Knappe, Removal of emerging contaminants of
concern by alternative adsorbents. Water Research, 2009. 43(15): p. 3787-3796.

55. Dias, J.M., M.C.M. Alvim-Ferraz, M.F. Almeida, J. Rivera-Utrilla, and M. Sánchez-Polo,
Waste materials for activated carbon preparation and its use in aqueous-phase
treatment: A review. Journal of Environmental Management, 2007. 85(4): p. 833-846.

56. Dąbrowski, A., P. Podkościelny, Z. Hubicki, and M. Barczak, Adsorption of phenolic


compounds by activated carbon—a critical review. Chemosphere, 2005. 58(8): p. 1049-
1070.

57. Wang, S. and Y. Peng, Natural zeolites as effective adsorbents in water and wastewater
treatment. Chemical Engineering Journal, 2010. 156(1): p. 11-24.

58. Ali, I., M. Asim, and T.A. Khan, Low cost adsorbents for the removal of organic
pollutants from wastewater. Journal of Environmental Management, 2012. 113: p. 170-
183.

159
59. Adeyemo, A.A., I.O. Adeoye, and O.S. Bello, Metal organic frameworks as adsorbents
for dye adsorption: overview, prospects and future challenges. Toxicological &
Environmental Chemistry, 2012. 94(10): p. 1846-1863.

60. Çeçen, F. and Ö. Aktas, Activated carbon for water and wastewater treatment:
Integration of adsorption and biological treatment. 2011: John Wiley & Sons.

61. Marsh, H. and F.R. Reinoso, Activated carbon. 2006: Elsevier.

62. Virla, L.D., V. Montes, J. Wu, S.F. Ketep, and J.M. Hill, Synthesis of porous carbon from
petroleum coke using steam, potassium and sodium: Combining treatments to create
mesoporosity. Microporous and Mesoporous Materials, 2016. 234: p. 239-247.

63. Shin, S., J. Jang, S.H. Yoon, and I. Mochida, A study on the effect of heat treatment on
functional groups of pitch based activated carbon fiber using FTIR. Carbon, 1997.
35(12): p. 1739-1743.

64. Toles, C.A., W.E. Marshall, and M.M. Johns, Surface functional groups on acid-
activated nutshell carbons. Carbon, 1999. 37(8): p. 1207-1214.

65. Kalderis, D., S. Bethanis, P. Paraskeva, and E. Diamadopoulos, Production of activated


carbon from bagasse and rice husk by a single-stage chemical activation method at low
retention times. Bioresource Technology, 2008. 99(15): p. 6809-6816.

66. Al-Degs, Y.S., M.I. El-Barghouthi, A.H. El-Sheikh, and G.M. Walker, Effect of solution
pH, ionic strength, and temperature on adsorption behavior of reactive dyes on activated
carbon. Dyes and Pigments, 2008. 77(1): p. 16-23.

67. Radovic, L.R., I.F. Silva, J.I. Ume, J.A. Menéndez, C.A.L.Y. Leon, and A.W. Scaroni,
An experimental and theoretical study of the adsorption of aromatics possessing
electron-withdrawing and electron-donating functional groups by chemically modified
activated carbons. Carbon, 1997. 35(9): p. 1339-1348.

68. Samarghandi, M., M. Hadi, S. Moayedi, and F.B. Askari, Two-parameter isotherms of
methyl orange sorption by pinecone derived activated carbon. Journal of Environmental
Health Science & Engineering, 2009. 6(4): p. 285-294.

69. Yu, H., T. Wang, W. Dai, X. Li, X. Hu, and N. Ma, Single and bicomponent anionic dyes
adsorption equilibrium studies on magnolia-leaf-based porous carbons. RSC Advances,
2015. 5(79): p. 63970-63977.

70. EBC. European Biochar Certificate - Guidelines for a Sustainable Production of


Biochar. 2012 [cited 2017 December 1]; Version 6.3E of 14th August 2017:[Available
from: http://www.europeanbiochar.org/en/download.

71. Lehmann, J. and S. Joseph, Biochar for environmental management: science, technology
and implementation. 2009: Routledge.

160
72. Ahmad, M., A.U. Rajapaksha, J.E. Lim, M. Zhang, N. Bolan, D. Mohan, M. Vithanage,
S.S. Lee, and Y.S. Ok, Biochar as a sorbent for contaminant management in soil and
water: A review. Chemosphere, 2014. 99: p. 19-33.

73. Rajapaksha, A.U., S.S. Chen, D.C.W. Tsang, M. Zhang, M. Vithanage, S. Mandal, B.
Gao, N.S. Bolan, and Y.S. Ok, Engineered/designer biochar for contaminant
removal/immobilization from soil and water: Potential and implication of biochar
modification. Chemosphere, 2016. 148(Supplement C): p. 276-291.

74. Bhuiyan, T.I., J.K. Tak, S. Sessarego, D. Harfield, and J.M. Hill, Adsorption of acid-
extractable organics from oil sands process-affected water onto biomass-based biochar:
Metal content matters. Chemosphere, 2017. 168(Supplement C): p. 1337-1344.

75. Shen, Y., Q. Fang, and B. Chen, Environmental Applications of Three-Dimensional


Graphene-Based Macrostructures: Adsorption, Transformation, and Detection.
Environmental Science & Technology, 2015. 49(1): p. 67-84.

76. Yang, K. and B. Xing, Adsorption of Organic Compounds by Carbon Nanomaterials in


Aqueous Phase: Polanyi Theory and Its Application. Chemical Reviews, 2010. 110(10):
p. 5989-6008.

77. Yao, Y., H. Bing, X. Feifei, and C. Xiaofeng, Equilibrium and kinetic studies of methyl
orange adsorption on multiwalled carbon nanotubes. Chemical Engineering Journal,
2011. 170(1): p. 82-89.

78. León, G., F. García, B. Miguel, and J. Bayo, Equilibrium, kinetic and thermodynamic
studies of methyl orange removal by adsorption onto granular activated carbon.
Desalination and Water Treatment, 2016. 57(36): p. 17104-17117.

79. Ji, L., F. Liu, Z. Xu, S. Zheng, and D. Zhu, Adsorption of Pharmaceutical Antibiotics on
Template-Synthesized Ordered Micro- and Mesoporous Carbons. Environmental Science
& Technology, 2010. 44(8): p. 3116-3122.

80. Wang, J., Z. Shi, J. Fan, Y. Ge, J. Yin, and G. Hu, Self-assembly of graphene into three-
dimensional structures promoted by natural phenolic acids. Journal of Materials
Chemistry, 2012. 22(42): p. 22459-22466.

81. Ma, J., F. Yu, L. Zhou, L. Jin, M. Yang, J. Luan, Y. Tang, H. Fan, Z. Yuan, and J. Chen,
Enhanced Adsorptive Removal of Methyl Orange and Methylene Blue from Aqueous
Solution by Alkali-Activated Multiwalled Carbon Nanotubes. ACS Applied Materials &
Interfaces, 2012. 4(11): p. 5749-5760.

82. Moreno-Castilla, C., Adsorption of organic molecules from aqueous solutions on carbon
materials. Carbon, 2004. 42(1): p. 83-94.

161
83. Wang, Y., Q. Gao, Q. You, G. Liao, H. Xia, and D. Wang, Porous polyimide framework:
A novel versatile adsorbent for highly efficient removals of azo dye and antibiotic.
Reactive and Functional Polymers, 2016. 103: p. 9-16.

84. Tajizadegan, H., O. Torabi, A. Heidary, M.H. Golabgir, and A. Jamshidi, Study of methyl
orange adsorption properties on ZnO–Al2O3 nanocomposite adsorbent particles.
Desalination and Water Treatment, 2016. 57(26): p. 12324-12334.

85. Mokhtari, P., M. Ghaedi, K. Dashtian, M.R. Rahimi, and M.K. Purkait, Removal of
methyl orange by copper sulfide nanoparticles loaded activated carbon: Kinetic and
isotherm investigation. Journal of Molecular Liquids, 2016. 219: p. 299-305.

86. Liu, Y., Y. Tian, C. Luo, G. Cui, and S. Yan, One-pot preparation of a MnO2-graphene-
carbon nanotube hybrid material for the removal of methyl orange from aqueous
solutions. New Journal of Chemistry, 2015. 39(7): p. 5484-5492.

87. Hassanzadeh-Tabrizi, S.A., M.M. Motlagh, and S. Salahshour, Synthesis of ZnO/CuO


nanocomposite immobilized on γ-Al2O3 and application for removal of methyl orange.
Applied Surface Science, 2016. 384: p. 237-243.

88. Cheah, W., S. Hosseini, M.A. Khan, T.G. Chuah, and T.S.Y. Choong, Acid modified
carbon coated monolith for methyl orange adsorption. Chemical Engineering Journal,
2013. 215–216: p. 747-754.

89. Worch, E., Adsorption technology in water treatment: fundamentals, processes, and
modeling. 2012: Walter de Gruyter.

90. Ahmed, M.B., J.L. Zhou, H.H. Ngo, and W. Guo, Adsorptive removal of antibiotics from
water and wastewater: Progress and challenges. Science of The Total Environment,
2015. 532: p. 112-126.

91. Omorogie, M.O., J.O. Babalola, and E.I. Unuabonah, Regeneration strategies for spent
solid matrices used in adsorption of organic pollutants from surface water: a critical
review. Desalination and Water Treatment, 2016. 57(2): p. 518-544.

92. Weng, C.-H. and M.-C. Hsu, Regeneration of granular activated carbon by an
electrochemical process. Separation and Purification Technology, 2008. 64(2): p. 227-
236.

93. Guo, Y. and E. Du, The Effects of Thermal Regeneration Conditions and Inorganic
Compounds on the Characteristics of Activated Carbon Used in Power Plant. Energy
Procedia, 2012. 17, Part A: p. 444-449.

94. Klavarioti, M., D. Mantzavinos, and D. Kassinos, Removal of residual pharmaceuticals


from aqueous systems by advanced oxidation processes. Environment International,
2009. 35(2): p. 402-417.

162
95. Ledesma, B., S. Román, E. Sabio, and A. Álvarez-Murillo, Improvement of spent
activated carbon regeneration by wet oxidation processes. The Journal of Supercritical
Fluids, 2015. 104: p. 94-103.

96. Pera-Titus, M., V. Garcı́a-Molina, M.A. Baños, J. Giménez, and S. Esplugas,


Degradation of chlorophenols by means of advanced oxidation processes: a general
review. Applied Catalysis B: Environmental, 2004. 47(4): p. 219-256.

97. Toledo, L.C., A.C.B. Silva, R. Augusti, and R.M. Lago, Application of Fenton’s reagent
to regenerate activated carbon saturated with organochloro compounds. Chemosphere,
2003. 50(8): p. 1049-1054.

98. Chen, Q., H. Liu, Z. Yang, and D. Tan, Regeneration performance of spent granular
activated carbon for tertiary treatment of dyeing wastewater by Fenton reagent and
hydrogen peroxide. Journal of Material Cycles and Waste Management, 2017. 19(1): p.
256-264.

99. Brown, N.W., E.P.L. Roberts, A. Chasiotis, T. Cherdron, and N. Sanghrajka, Atrazine
removal using adsorption and electrochemical regeneration. Water Research, 2004.
38(13): p. 3067-3074.

100. Aktaş, Ö. and F. Çeçen, Bioregeneration of activated carbon: A review. International


Biodeterioration & Biodegradation, 2007. 59(4): p. 257-272.

101. Ledding, W.E., Wet oxidative reactivation of spent active carbon, U. Patent, Editor.
1961: US.

102. Shende, R.V. and V.V. Mahajani, Wet oxidative regeneration of activated carbon loaded
with reactive dye. Waste Management, 2002. 22(1): p. 73-83.

103. Lim, T.-T., P.-S. Yap, M. Srinivasan, and A.G. Fane, TiO2/AC Composites for
Synergistic Adsorption-Photocatalysis Processes: Present Challenges and Further
Developments for Water Treatment and Reclamation. Critical Reviews in Environmental
Science and Technology, 2011. 41(13): p. 1173-1230.

104. Narbaitz, R.M. and A. Karimi‐Jashni, Electrochemical regeneration of granular


activated carbons loaded with phenol and natural organic matter. Environmental
Technology, 2009. 30(1): p. 27-36.

105. Ania, C.O. and F. Béguin, Electrochemical Regeneration of Activated Carbon Cloth
Exhausted with Bentazone. Environmental Science & Technology, 2008. 42(12): p. 4500-
4506.

106. Zhang, H., Regeneration of exhausted activated carbon by electrochemical method.


Chemical Engineering Journal, 2002. 85(1): p. 81-85.

163
107. Mohammed, F.M., E.P.L. Roberts, A. Hill, A.K. Campen, and N.W. Brown, Continuous
water treatment by adsorption and electrochemical regeneration. Water Research, 2011.
45(10): p. 3065-3074.

108. Radjenovic, J. and D.L. Sedlak, Challenges and Opportunities for Electrochemical
Processes as Next-Generation Technologies for the Treatment of Contaminated Water.
Environmental Science & Technology, 2015. 49(19): p. 11292-11302.

109. Han, Y., X. Quan, X. Ruan, and W. Zhang, Integrated electrochemically enhanced
adsorption with electrochemical regeneration for removal of acid orange 7 using
activated carbon fibers. Separation and Purification Technology, 2008. 59(1): p. 43-49.

110. Hussain, S.N., E.P.L. Roberts, H.M.A. Asghar, A.K. Campen, and N.W. Brown,
Oxidation of phenol and the adsorption of breakdown products using a graphite
adsorbent with electrochemical regeneration. Electrochimica Acta, 2013. 92: p. 20-30.

111. Hussain, S.N., H.M.A. Asghar, A.K. Campen, N.W. Brown, and E.P.L. Roberts,
Breakdown products formed due to oxidation of adsorbed phenol by electrochemical
regeneration of a graphite adsorbent. Electrochimica Acta, 2013. 110: p. 550-559.

112. Liu, Y., J. Xie, C.N. Ong, C.D. Vecitis, and Z. Zhou, Electrochemical wastewater
treatment with carbon nanotube filters coupled with in situ generated H2O2.
Environmental Science: Water Research & Technology, 2015.

113. Munoz, M., Z.M. de Pedro, J.A. Casas, and J.J. Rodriguez, Preparation of magnetite-
based catalysts and their application in heterogeneous Fenton oxidation – A review.
Applied Catalysis B: Environmental, 2015. 176-177(Supplement C): p. 249-265.

114. Centi, G., S. Perathoner, T. Torre, and M.G. Verduna, Catalytic wet oxidation with H2O2
of carboxylic acids on homogeneous and heterogeneous Fenton-type catalysts. Catalysis
Today, 2000. 55(1): p. 61-69.

115. Pignatello, J.J., E. Oliveros, and A. MacKay, Advanced Oxidation Processes for Organic
Contaminant Destruction Based on the Fenton Reaction and Related Chemistry. Critical
Reviews in Environmental Science and Technology, 2006. 36(1): p. 1-84.

116. Ribeiro, R.S., A.M.T. Silva, J.L. Figueiredo, J.L. Faria, and H.T. Gomes, Catalytic wet
peroxide oxidation: a route towards the application of hybrid magnetic carbon
nanocomposites for the degradation of organic pollutants. A review. Applied Catalysis B:
Environmental, 2016. 187(Supplement C): p. 428-460.

117. Lin, S.-S. and M.D. Gurol, Catalytic Decomposition of Hydrogen Peroxide on Iron
Oxide: Kinetics, Mechanism, and Implications. Environmental Science & Technology,
1998. 32(10): p. 1417-1423.

164
118. Kwan, W.P. and B.M. Voelker, Rates of Hydroxyl Radical Generation and Organic
Compound Oxidation in Mineral-Catalyzed Fenton-like Systems. Environmental Science
& Technology, 2003. 37(6): p. 1150-1158.

119. Luo, W., L. Zhu, N. Wang, H. Tang, M. Cao, and Y. She, Efficient Removal of Organic
Pollutants with Magnetic Nanoscaled BiFeO3 as a Reusable Heterogeneous Fenton-Like
Catalyst. Environmental Science & Technology, 2010. 44(5): p. 1786-1791.

120. Kan, E. and S.G. Huling, Effects of Temperature and Acidic Pre-Treatment on Fenton-
Driven Oxidation of MTBE-Spent Granular Activated Carbon. Environmental Science &
Technology, 2009. 43(5): p. 1493-1499.

121. Huling, S.G., P.K. Jones, W.P. Ela, and R.G. Arnold, Fenton-driven chemical
regeneration of MTBE-spent GAC. Water Research, 2005. 39(10): p. 2145-2153.

122. Costa, R.C.C., M.F.F. Lelis, L.C.A. Oliveira, J.D. Fabris, J.D. Ardisson, R.R.V.A. Rios,
C.N. Silva, and R.M. Lago, Novel active heterogeneous Fenton system based on
Fe3−xMxO4 (Fe, Co, Mn, Ni): The role of M2+ species on the reactivity towards H2O2
reactions. Journal of Hazardous Materials, 2006. 129(1): p. 171-178.

123. Sarasidis, V.C., K.V. Plakas, and A.J. Karabelas, Novel water-purification hybrid
processes involving in-situ regenerated activated carbon, membrane separation and
advanced oxidation. Chemical Engineering Journal, 2017. 328: p. 1153-1163.

124. Li, B., J. Ma, L. Zhou, and Y. Qiu, Magnetic microsphere to remove tetracycline from
water: Adsorption, H2O2 oxidation and regeneration. Chemical Engineering Journal,
2017. 330(Supplement C): p. 191-201.

125. Sirés, I., E. Brillas, M. Oturan, M. Rodrigo, and M. Panizza, Electrochemical advanced
oxidation processes: today and tomorrow. A review. Environmental Science and
Pollution Research, 2014. 21(14): p. 8336-8367.

126. Brillas, E., R.M. Bastida, E. Llosa, and J. Casado, Electrochemical Destruction of Aniline
and 4‐Chloroaniline for Wastewater Treatment Using a Carbon‐PTFE O 2 ‐
Fed Cathode. Journal of The Electrochemical Society, 1995. 142(6): p. 1733-1741.

127. Brillas, E., M.Á. Baños, M. Skoumal, P.L. Cabot, J.A. Garrido, and R.M. Rodríguez,
Degradation of the herbicide 2,4-DP by anodic oxidation, electro-Fenton and
photoelectro-Fenton using platinum and boron-doped diamond anodes. Chemosphere,
2007. 68(2): p. 199-209.

128. Liu, Y., S. Chen, X. Quan, H. Yu, H. Zhao, and Y. Zhang, Efficient Mineralization of
Perfluorooctanoate by Electro-Fenton with H2O2 Electro-generated on Hierarchically
Porous Carbon. Environmental Science & Technology, 2015. 49(22): p. 13528-13533.

129. Brillas, E., J.C. Calpe, and J. Casado, Mineralization of 2,4-D by advanced
electrochemical oxidation processes. Water Research, 2000. 34(8): p. 2253-2262.

165
130. Boye, B., M.M. Dieng, and E. Brillas, Degradation of Herbicide 4-Chlorophenoxyacetic
Acid by Advanced Electrochemical Oxidation Methods. Environmental Science &
Technology, 2002. 36(13): p. 3030-3035.

131. Song, C. and J. Zhang, Electrocatalytic Oxygen Reduction Reaction, in PEM Fuel Cell
Electrocatalysts and Catalyst Layers: Fundamentals and Applications, J. Zhang, Editor.
2008, Springer London: London. p. 89-134.

132. Cheng, Y., J. Zhang, and S.P. Jiang, Are metal-free pristine carbon nanotubes
electrocatalytically active? Chemical Communications, 2015. 51(72): p. 13764-13767.

133. Daneshvar, N., S. Aber, V. Vatanpour, and M.H. Rasoulifard, Electro-Fenton treatment
of dye solution containing Orange II: Influence of operational parameters. Journal of
Electroanalytical Chemistry, 2008. 615(2): p. 165-174.

134. Ge, X., A. Sumboja, D. Wuu, T. An, B. Li, F.W.T. Goh, T.S.A. Hor, Y. Zong, and Z.
Liu, Oxygen Reduction in Alkaline Media: From Mechanisms to Recent Advances of
Catalysts. ACS Catalysis, 2015. 5(8): p. 4643-4667.

135. Byers, J.C., A.G. Güell, and P.R. Unwin, Nanoscale Electrocatalysis: Visualizing Oxygen
Reduction at Pristine, Kinked, and Oxidized Sites on Individual Carbon Nanotubes.
Journal of the American Chemical Society, 2014. 136(32): p. 11252-11255.

136. Randviir, E.P. and C.E. Banks, The Oxygen Reduction Reaction at Graphene Modified
Electrodes. Electroanalysis, 2014. 26(1): p. 76-83.

137. Liu, Y., X. Quan, X. Fan, H. Wang, and S. Chen, High-Yield Electrosynthesis of
Hydrogen Peroxide from Oxygen Reduction by Hierarchically Porous Carbon.
Angewandte Chemie International Edition, 2015. 54(23): p. 6837-6841.

138. Miao, J., H. Zhu, Y. Tang, Y. Chen, and P. Wan, Graphite felt electrochemically
modified in H2SO4 solution used as a cathode to produce H2O2 for pre-oxidation of
drinking water. Chemical Engineering Journal, 2014. 250: p. 312-318.

139. Lu, Z., G. Chen, S. Siahrostami, Z. Chen, K. Liu, J. Xie, L. Liao, T. Wu, D. Lin, Y. Liu,
T.F. Jaramillo, J.K. Nørskov, and Y. Cui, High-efficiency oxygen reduction to hydrogen
peroxide catalysed by oxidized carbon materials. Nature Catalysis, 2018. 1(2): p. 156-
162.

140. Jürmann, G. and K. Tammeveski, Electroreduction of oxygen on multi-walled carbon


nanotubes modified highly oriented pyrolytic graphite electrodes in alkaline solution.
Journal of Electroanalytical Chemistry, 2006. 597(2): p. 119-126.

141. Sarapuu, A., K. Vaik, D.J. Schiffrin, and K. Tammeveski, Electrochemical reduction of
oxygen on anthraquinone-modified glassy carbon electrodes in alkaline solution. Journal
of Electroanalytical Chemistry, 2003. 541: p. 23-29.

166
142. Sun, Y., I. Sinev, W. Ju, A. Bergmann, S. Dresp, S. Kühl, C. Spöri, H. Schmies, H.
Wang, D. Bernsmeier, B. Paul, R. Schmack, R. Kraehnert, B. Roldan Cuenya, and P.
Strasser, Efficient Electrochemical Hydrogen Peroxide Production from Molecular
Oxygen on Nitrogen-Doped Mesoporous Carbon Catalysts. ACS Catalysis, 2018: p.
2844-2856.

143. Park, J., Y. Nabae, T. Hayakawa, and M.-a. Kakimoto, Highly Selective Two-Electron
Oxygen Reduction Catalyzed by Mesoporous Nitrogen-Doped Carbon. ACS Catalysis,
2014. 4(10): p. 3749-3754.

144. Petrucci, E., A. Da Pozzo, and L. Di Palma, On the ability to electrogenerate hydrogen
peroxide and to regenerate ferrous ions of three selected carbon-based cathodes for
electro-Fenton processes. Chemical Engineering Journal, 2016. 283: p. 750-758.

145. Sirés, I., J.A. Garrido, R.M. Rodríguez, E. Brillas, N. Oturan, and M.A. Oturan, Catalytic
behavior of the Fe3+/Fe2+ system in the electro-Fenton degradation of the antimicrobial
chlorophene. Applied Catalysis B: Environmental, 2007. 72(3): p. 382-394.

146. Yatagai, T., Y. Ohkawa, D. Kubo, and Y. Kawase, Hydroxyl radical generation in
electro-Fenton process with a gas-diffusion electrode: Linkages with electro-chemical
generation of hydrogen peroxide and iron redox cycle. Journal of Environmental Science
and Health, Part A, 2017. 52(1): p. 74-83.

147. Petrucci, E., D. Montanaro, and S. Le Donne, Effect of carbon material on the
performance of a gas diffusion electrode in electro-Fenton process. Journal of
Environmental Engineering Management, 2009. 19(5): p. 299-305.

148. Zhao, H., Y. Wang, Y. Wang, T. Cao, and G. Zhao, Electro-Fenton oxidation of
pesticides with a novel Fe3O4@Fe2O3/activated carbon aerogel cathode: High activity,
wide pH range and catalytic mechanism. Applied Catalysis B: Environmental, 2012. 125:
p. 120-127.

149. Ganiyu, S.O., T.X. Huong Le, M. Bechelany, G. Esposito, E.D. van Hullebusch, M.A.
Oturan, and M. Cretin, A hierarchical CoFe-layered double hydroxide modified carbon-
felt cathode for heterogeneous electro-Fenton process. Journal of Materials Chemistry A,
2017. 5(7): p. 3655-3666.

150. Pinheiro, V.S., E.C. Paz, L.R. Aveiro, L.S. Parreira, F.M. Souza, P.H.C. Camargo, and
M.C. Santos, Ceria high aspect ratio nanostructures supported on carbon for hydrogen
peroxide electrogeneration. Electrochimica Acta, 2018. 259: p. 865-872.

151. Brillas, E., B. Boye, I. Sirés, J.A. Garrido, R.M.a. Rodrı́guez, C. Arias, P.-L.s. Cabot, and
C. Comninellis, Electrochemical destruction of chlorophenoxy herbicides by anodic
oxidation and electro-Fenton using a boron-doped diamond electrode. Electrochimica
Acta, 2004. 49(25): p. 4487-4496.

167
152. Oturan, N., E. Brillas, and M.A. Oturan, Unprecedented total mineralization of atrazine
and cyanuric acid by anodic oxidation and electro-Fenton with a boron-doped diamond
anode. Environmental Chemistry Letters, 2012. 10(2): p. 165-170.

153. Sopaj, F., N. Oturan, J. Pinson, F. Podvorica, and M.A. Oturan, Effect of the anode
materials on the efficiency of the electro-Fenton process for the mineralization of the
antibiotic sulfamethazine. Applied Catalysis B: Environmental, 2016. 199: p. 331-341.

154. Lin, H., N. Oturan, J. Wu, H. Zhang, and M.A. Oturan, Cold incineration of sucralose in
aqueous solution by electro-Fenton process. Separation and Purification Technology,
2017. 173: p. 218-225.

155. GarcÍA-OtÓN, M., F. Montilla, M.A. Lillo-RÓDenas, E. MorallÓN, and J. VÁZquez,


Electrochemical Regeneration of Activated Carbon Saturated with Toluene. Journal of
Applied Electrochemistry, 2005. 35(3): p. 319-325.

156. Karimi-Jashni, A. and R.M. Narbaitz, Electrochemical reactivation of granular activated


carbon: pH dependence. Journal of Environmental Engineering and Science, 2005. 4(3):
p. 187-194.

157. Zhang, H., L. Ye, and H. Zhong, Regeneration of phenol-saturated activated carbon in
an electrochemical reactor. Journal of Chemical Technology & Biotechnology, 2002.
77(11): p. 1246-1250.

158. Bañuelos, J., O. García-Rodríguez, F. Rodríguez-Valadez, J. Manríquez, E. Bustos, A.


Rodríguez, and L. Godínez, Cathodic polarization effect on the electro-Fenton
regeneration of activated carbon. Journal of Applied Electrochemistry, 2015. 45(5): p.
523-531.

159. Roth, H., Y. Gendel, P. Buzatu, O. David, and M. Wessling, Tubular carbon nanotube-
based gas diffusion electrode removes persistent organic pollutants by a cyclic
adsorption – Electro-Fenton process. Journal of Hazardous Materials, 2016. 307: p. 1-6.

160. Trellu, C., Y. Péchaud, N. Oturan, E. Mousset, D. Huguenot, E.D. van Hullebusch, G.
Esposito, and M.A. Oturan, Comparative study on the removal of humic acids from
drinking water by anodic oxidation and electro-Fenton processes: Mineralization
efficiency and modelling. Applied Catalysis B: Environmental, 2016. 194: p. 32-41.

161. Kendrick, K.L. and W.R. Gilkerson, The state of aggregation of methyl orange in water.
Journal of Solution Chemistry, 1987. 16(4): p. 257-267.

162. Oliveira, L.C.A., R.V.R.A. Rios, J.D. Fabris, V. Garg, K. Sapag, and R.M. Lago,
Activated carbon/iron oxide magnetic composites for the adsorption of contaminants in
water. Carbon, 2002. 40(12): p. 2177-2183.

168
163. Mohan, D., A. Sarswat, V.K. Singh, M. Alexandre-Franco, and C.U. Pittman Jr,
Development of magnetic activated carbon from almond shells for trinitrophenol removal
from water. Chemical Engineering Journal, 2011. 172(2–3): p. 1111-1125.

164. Zhou, M., Q. Yu, and L. Lei, The preparation and characterization of a graphite–PTFE
cathode system for the decolorization of C.I. Acid Red 2. Dyes and Pigments, 2008.
77(1): p. 129-136.

165. Wang, Y., Y. Liu, X.-z. Li, F. Zeng, and H. Liu, A highly-ordered porous carbon
material based cathode for energy-efficient electro-Fenton process. Separation and
Purification Technology, 2013. 106: p. 32-37.

166. Brunauer, S., P.H. Emmett, and E. Teller, Adsorption of Gases in Multimolecular Layers.
Journal of the American Chemical Society, 1938. 60(2): p. 309-319.

167. Thommes, M., K. Kaneko, V. Neimark Alexander, P. Olivier James, F. Rodriguez-


Reinoso, J. Rouquerol, and S.W. Sing Kenneth, Physisorption of gases, with special
reference to the evaluation of surface area and pore size distribution (IUPAC Technical
Report), in Pure and Applied Chemistry. 2015. p. 1051.

168. Jagiello, J. and J.P. Olivier, 2D-NLDFT adsorption models for carbon slit-shaped pores
with surface energetical heterogeneity and geometrical corrugation. Carbon, 2013. 55: p.
70-80.

169. Lee, M., X-ray Diffraction for Materials Research: From Fundamentals to Applications.
2016: CRC Press.

170. Langford, J.I. and A.J.C. Wilson, Scherrer after sixty years: A survey and some new
results in the determination of crystallite size. Journal of Applied Crystallography, 1978.
11(2): p. 102-113.

171. Mayoral, M.C., M.T. Izquierdo, J.M. Andrés, and B. Rubio, Different approaches to
proximate analysis by thermogravimetry analysis. Thermochimica Acta, 2001. 370(1): p.
91-97.

172. Suart, B., Infrared spectroscopy: Fundamental and applications. Analytical Techniques
in the Sciences. 2004: John Wiley & Sons, Ltd.

173. Christy, A.A., Y. Ozaki, and V.G. Gregoriou, Modern Fourier transform infrared
spectroscopy. 2001: Elsevier.

174. Eisenberg, G., Colorimetric Determination of Hydrogen Peroxide. Industrial &


Engineering Chemistry Analytical Edition, 1943. 15(5): p. 327-328.

175. Narbaitz, R.M. and J. Cen, Alternative methods for determining the percentage
regeneration of activated carbon. Water Research, 1997. 31(10): p. 2532-2542.

169
176. Hadi, P., J. Guo, J. Barford, and G. McKay, Multilayer Dye Adsorption in Activated
Carbons—Facile Approach to Exploit Vacant Sites and Interlayer Charge Interaction.
Environmental Science & Technology, 2016. 50(10): p. 5041-5049.

177. Oleszczuk, P., B. Pan, and B. Xing, Adsorption and Desorption of Oxytetracycline and
Carbamazepine by Multiwalled Carbon Nanotubes. Environmental Science &
Technology, 2009. 43(24): p. 9167-9173.

178. Brown, N.W., E.P.L. Roberts, A.A. Garforth, and R.A.W. Dryfe, Electrochemical
regeneration of a carbon-based adsorbent loaded with crystal violet dye. Electrochimica
Acta, 2004. 49(20): p. 3269-3281.

179. Sajab, M.S., C.H. Chia, C.H. Chan, S. Zakaria, H. Kaco, S.W. Chook, S.X. Chin, and
A.A.M. Noor, Bifunctional graphene oxide-cellulose nanofibril aerogel loaded with
Fe(iii) for the removal of cationic dye via simultaneous adsorption and Fenton oxidation.
RSC Advances, 2016. 6(24): p. 19819-19825.

180. Darmograi, G., B. Prelot, A. Geneste, L.-C. De Menorval, and J. Zajac, Removal of three
anionic orange-type dyes and Cr(VI) oxyanion from aqueous solutions onto strongly
basic anion-exchange resin. The effect of single-component and competitive adsorption.
Colloids and Surfaces A: Physicochemical and Engineering Aspects, 2016. 508: p. 240-
250.

181. Mohammadi, N., H. Khani, V.K. Gupta, E. Amereh, and S. Agarwal, Adsorption process
of methyl orange dye onto mesoporous carbon material–kinetic and thermodynamic
studies. Journal of Colloid and Interface Science, 2011. 362(2): p. 457-462.

182. Gregor, C., M. Hermanek, D. Jancik, J. Pechousek, J. Filip, J. Hrbac, and R. Zboril, The
Effect of Surface Area and Crystal Structure on the Catalytic Efficiency of Iron(III) Oxide
Nanoparticles in Hydrogen Peroxide Decomposition. European Journal of Inorganic
Chemistry, 2010. 2010(16): p. 2343-2351.

183. Alatalo, S.-M., E. Makila, E. Repo, M. Heinonen, J. Salonen, E. Kukk, M. Sillanpaa, and
M.-M. Titirici, Meso- and microporous soft templated hydrothermal carbons for dye
removal from water. Green Chemistry, 2016. 18(4): p. 1137-1146.

184. Alkan, M., Ö. Demirbaş, and M. Doğan, Adsorption kinetics and thermodynamics of an
anionic dye onto sepiolite. Microporous and Mesoporous Materials, 2007. 101(3): p. 388-
396.

185. Cheung, W.H., Y.S. Szeto, and G. McKay, Intraparticle diffusion processes during acid
dye adsorption onto chitosan. Bioresource Technology, 2007. 98(15): p. 2897-2904.

186. Zhang, S., M. Zeng, J. Li, J. Li, J. Xu, and X. Wang, Porous magnetic carbon sheets from
biomass as an adsorbent for the fast removal of organic pollutants from aqueous
solution. Journal of Materials Chemistry A, 2014. 2(12): p. 4391-4397.

170
187. Lücking, F., H. Köser, M. Jank, and A. Ritter, Iron powder, graphite and activated
carbon as catalysts for the oxidation of 4-chlorophenol with hydrogen peroxide in
aqueous solution. Water Research, 1998. 32(9): p. 2607-2614.

188. Rey, A., J.A. Zazo, J.A. Casas, A. Bahamonde, and J.J. Rodriguez, Influence of the
structural and surface characteristics of activated carbon on the catalytic decomposition
of hydrogen peroxide. Applied Catalysis A: General, 2011. 402(1–2): p. 146-155.

189. An, D., P. Westerhoff, M. Zheng, M. Wu, Y. Yang, and C.-A. Chiu, UV-activated
persulfate oxidation and regeneration of NOM-Saturated granular activated carbon.
Water Research, 2015. 73: p. 304-310.

190. Israelachvili, J.N., Intermolecular and surface forces: revised third edition. 2011:
Academic press.

191. Velichkova, F., C. Julcour-Lebigue, B. Koumanova, and H. Delmas, Heterogeneous


Fenton oxidation of paracetamol using iron oxide (nano)particles. Journal of
Environmental Chemical Engineering, 2013. 1(4): p. 1214-1222.

192. Attri, P., Y.H. Kim, D.H. Park, J.H. Park, Y.J. Hong, H.S. Uhm, K.-N. Kim, A. Fridman,
and E.H. Choi, Generation mechanism of hydroxyl radical species and its lifetime
prediction during the plasma-initiated ultraviolet (UV) photolysis. Scientific Reports,
2015. 5: p. 9332.

193. Jiang, P.-Y., Y. Katsumura, R. Nagaishi, M. Domae, K. Ishikawa, K. Ishigure, and Y.


Yoshida, Pulse radiolysis study of concentrated sulfuric acid solutions. Formation
mechanism, yield and reactivity of sulfate radicals. Journal of the Chemical Society,
Faraday Transactions, 1992. 88(12): p. 1653-1658.

194. Lee, J., W. Choi, and J. Yoon, Photocatalytic Degradation of N-Nitrosodimethylamine:


Mechanism, Product Distribution, and TiO2 Surface Modification. Environmental
Science & Technology, 2005. 39(17): p. 6800-6807.

195. Padmaja, S. and S.A. Madison, Hydroxyl radical‐induced oxidation of azo dyes: a pulse
radiolysis study. Journal of physical organic chemistry, 1999. 12(3): p. 221-226.

196. Xiao, Y. and J.M. Hill, Impact of Pore Size on Fenton Oxidation of Methyl Orange
Adsorbed on Magnetic Carbon Materials: Trade-Off between Capacity and
Regenerability. Environmental Science & Technology, 2017. 51(8): p. 4567-4575.

197. Hu, Z., M. Beuret, H. Khan, and P.A. Ariya, Development of a Recyclable Remediation
System for Gaseous BTEX: Combination of Iron Oxides Nanoparticles Adsorbents and
Electrochemistry. ACS Sustainable Chemistry & Engineering, 2014. 2(12): p. 2739-2747.

198. Naghizadeh, A., Regeneration of Carbon Nanotubes Exhausted with Humic Acid Using
Electro-Fenton Technology. Arabian Journal for Science and Engineering, 2016. 41(1): p.
155-161.

171
199. Oturan, M.A., E. Guivarch, N. Oturan, and I. Sirés, Oxidation pathways of malachite
green by Fe3+-catalyzed electro-Fenton process. Applied Catalysis B: Environmental,
2008. 82(3–4): p. 244-254.

200. Zhou, M., Q. Tan, Q. Wang, Y. Jiao, N. Oturan, and M.A. Oturan, Degradation of
organics in reverse osmosis concentrate by electro-Fenton process. Journal of Hazardous
Materials, 2012. 215–216: p. 287-293.

201. Wang, Y., X. Li, L. Zhen, H. Zhang, Y. Zhang, and C. Wang, Electro-Fenton treatment
of concentrates generated in nanofiltration of biologically pretreated landfill leachate.
Journal of Hazardous Materials, 2012. 229–230: p. 115-121.

202. Oturan, M.A., J. Peiroten, P. Chartrin, and A.J. Acher, Complete Destruction of p-
Nitrophenol in Aqueous Medium by Electro-Fenton Method. Environmental Science &
Technology, 2000. 34(16): p. 3474-3479.

203. Özcan, A., Y. Şahin, A. Savaş Koparal, and M.A. Oturan, Carbon sponge as a new
cathode material for the electro-Fenton process: Comparison with carbon felt cathode
and application to degradation of synthetic dye basic blue 3 in aqueous medium. Journal
of Electroanalytical Chemistry, 2008. 616(1–2): p. 71-78.

204. Berenguer, R., J.P. Marco-Lozar, C. Quijada, D. Cazorla-Amorós, and E. Morallón,


Effect of electrochemical treatments on the surface chemistry of activated carbon.
Carbon, 2009. 47(4): p. 1018-1027.

205. Berenguer, R., J.P. Marco-Lozar, C. Quijada, D. Cazorla-Amorós, and E. Morallón,


Electrochemical regeneration and porosity recovery of phenol-saturated granular
activated carbon in an alkaline medium. Carbon, 2010. 48(10): p. 2734-2745.

206. Wang, L. and N. Balasubramanian, Electrochemical regeneration of granular activated


carbon saturated with organic compounds. Chemical Engineering Journal, 2009. 155(3):
p. 763-768.

207. Ai, Z., T. Mei, J. Liu, J. Li, F. Jia, L. Zhang, and J. Qiu, Fe@Fe2O3 Core−Shell
Nanowires as an Iron Reagent. 3. Their Combination with CNTs as an Effective Oxygen-
Fed Gas Diffusion Electrode in a Neutral Electro-Fenton System. The Journal of Physical
Chemistry C, 2007. 111(40): p. 14799-14803.

208. GARCÍA-OTÓN, M., F. MONTILLA, M.A. LILLO-RÓDENAS, E. MORALLÓN, and


J.L. VÁZQUEZ, Electrochemical Regeneration of Activated Carbon Saturated with
Toluene. Journal of Applied Electrochemistry, 2005. 35(3): p. 319-325.

209. Zhou, M.H. and L.C. Lei, Electrochemical regeneration of activated carbon loaded with
p-nitrophenol in a fluidized electrochemical reactor. Electrochimica Acta, 2006. 51(21):
p. 4489-4496.

172
210. Liu, R.-H., G.-P. Sheng, M. Sun, G.-L. Zang, W.-W. Li, Z.-H. Tong, F. Dong, M. Hon-
Wah Lam, and H.-Q. Yu, Enhanced reductive degradation of methyl orange in a
microbial fuel cell through cathode modification with redox mediators. Applied
Microbiology and Biotechnology, 2011. 89(1): p. 201-208.

211. Mu, Y., K. Rabaey, R.A. Rozendal, Z. Yuan, and J. Keller, Decolorization of Azo Dyes in
Bioelectrochemical Systems. Environmental Science & Technology, 2009. 43(13): p.
5137-5143.

212. Ma, H., B. Wang, and X. Luo, Studies on degradation of Methyl Orange wastewater by
combined electrochemical process. Journal of Hazardous Materials, 2007. 149(2): p. 492-
498.

213. Xiao, X., W.-W. Zhu, Y.-B. Lei, Q.-Y. Liu, Q. Li, and W.-W. Li, Zwitterionic buffer-
induced visible light excitation of TiO2 for efficient pollutant photodegradation. RSC
Advances, 2016. 6(42): p. 35449-35454.

214. Comparelli, R., E. Fanizza, M.L. Curri, P.D. Cozzoli, G. Mascolo, R. Passino, and A.
Agostiano, Photocatalytic degradation of azo dyes by organic-capped anatase TiO2
nanocrystals immobilized onto substrates. Applied Catalysis B: Environmental, 2005.
55(2): p. 81-91.

215. Gemeay, A.H., Kinetics and mechanism of the reduction of some azo-dyes by inorganic
oxysulfur compounds. Dyes and Pigments, 2002. 54(3): p. 201-212.

216. Liu, Y., L. Jin, and J.-f. Liu, A New Route for Synthesis of 2-Substituted-3-amino-5-
phenyl-7-N,N-dimethylamino Phenazinium Chloride Salts. Journal of Heterocyclic
Chemistry, 2017. 54(3): p. 1931-1936.

217. Salgado, L.T., R. Tomazetto, L.P. Cinelli, M. Farina, and G.M. Amado Filho, The
influence of brown algae alginates on phenolic compounds capability of ultraviolet
radiation absorption in vitro. Brazilian Journal of Oceanography, 2007. 55: p. 145-154.

218. Bazhin, N., The Born Formula Describes Enthalpy of Ions Solvation. ISRN
Thermodynamics, 2012. 2012: p. 3.

219. Han, Y., X. Quan, S. Chen, S. Wang, and Y. Zhang, Electrochemical enhancement of
adsorption capacity of activated carbon fibers and their surface physicochemical
characterizations. Electrochimica Acta, 2007. 52(9): p. 3075-3081.

220. Li, X., S. Chen, X. Quan, and Y. Zhang, Enhanced Adsorption of PFOA and PFOS on
Multiwalled Carbon Nanotubes under Electrochemical Assistance. Environmental
Science & Technology, 2011. 45(19): p. 8498-8505.

221. Han, Y., X. Quan, S. Chen, H. Zhao, C. Cui, and Y. Zhao, Electrochemically enhanced
adsorption of phenol on activated carbon fibers in basic aqueous solution. Journal of
Colloid and Interface Science, 2006. 299(2): p. 766-771.

173
222. Han, Y., X. Quan, S. Chen, H. Zhao, C. Cui, and Y. Zhao, Electrochemically enhanced
adsorption of aniline on activated carbon fibers. Separation and Purification Technology,
2006. 50(3): p. 365-372.

223. Kötz, R. and M. Carlen, Principles and applications of electrochemical capacitors.


Electrochimica Acta, 2000. 45(15): p. 2483-2498.

224. Wu, C., A. De Visscher, and I.D. Gates, Reactions of hydroxyl radicals with benzoic acid
and benzoate. RSC Advances, 2017. 7(57): p. 35776-35785.

225. Álvarez, P.M., F.J. Beltrán, V. Gómez-Serrano, J. Jaramillo, and E.M. Rodrıǵ uez,
Comparison between thermal and ozone regenerations of spent activated carbon
exhausted with phenol. Water Research, 2004. 38(8): p. 2155-2165.

226. Cannon, F.S., V.L. Snoeyink, R.G. Lee, and G. Dagois, Reaction mechanism of calcium-
catalyzed thermal regeneration of spent granular activated carbon. Carbon, 1994. 32(7):
p. 1285-1301.

227. Panizza, M. and G. Cerisola, Electro-Fenton degradation of synthetic dyes. Water


Research, 2009. 43(2): p. 339-344.

228. Le, T.X.H., R. Esmilaire, M. Drobek, M. Bechelany, C. Vallicari, S. Cerneaux, A. Julbe,


and M. Cretin, Nitrogen-Doped Graphitized Carbon Electrodes for Biorefractory
Pollutant Removal. The Journal of Physical Chemistry C, 2017. 121(28): p. 15188-
15197.

229. Wang, A., J. Qu, J. Ru, H. Liu, and J. Ge, Mineralization of an azo dye Acid Red 14 by
electro-Fenton's reagent using an activated carbon fiber cathode. Dyes and Pigments,
2005. 65(3): p. 227-233.

230. Dong, H., H. Yu, X. Wang, Q. Zhou, and J. Feng, A novel structure of scalable air-
cathode without Nafion and Pt by rolling activated carbon and PTFE as catalyst layer in
microbial fuel cells. Water Research, 2012. 46(17): p. 5777-5787.

231. Dong, H., H. Yu, and X. Wang, Catalysis Kinetics and Porous Analysis of Rolling
Activated Carbon-PTFE Air-Cathode in Microbial Fuel Cells. Environmental Science &
Technology, 2012. 46(23): p. 13009-13015.

232. Bañuelos, J.A., A. El-Ghenymy, F.J. Rodríguez, J. Manríquez, E. Bustos, A. Rodríguez,


E. Brillas, and L.A. Godínez, Study of an Air Diffusion Activated Carbon Packed
Electrode for an Electro-Fenton Wastewater Treatment. Electrochimica Acta, 2014. 140:
p. 412-418.

233. Zarei, M., D. Salari, A. Niaei, and A. Khataee, Peroxi-coagulation degradation of C.I.
Basic Yellow 2 based on carbon-PTFE and carbon nanotube-PTFE electrodes as
cathode. Electrochimica Acta, 2009. 54(26): p. 6651-6660.

174
234. Alghunaim, A., S. Kirdponpattara, and B.-m.Z. Newby, Techniques for determining
contact angle and wettability of powders. Powder Technology, 2016. 287: p. 201-215.

235. Horsthemke, A. and J.J. Schröder, The wettability of industrial surfaces: Contact angle
measurements and thermodynamic analysis. Chemical Engineering and Processing:
Process Intensification, 1985. 19(5): p. 277-285.

236. Law, K.-Y., Definitions for Hydrophilicity, Hydrophobicity, and Superhydrophobicity:


Getting the Basics Right. The Journal of Physical Chemistry Letters, 2014. 5(4): p. 686-
688.

237. García, A.B., A. Martínez-Alonso, C.A. Leon y Leon, and J.M.D. Tascón, Modification
of the surface properties of an activated carbon by oxygen plasma treatment. Fuel, 1998.
77(6): p. 613-624.

238. Gruger, A., A. Régis, T. Schmatko, and P. Colomban, Nanostructure of Nafion®


membranes at different states of hydration: An IR and Raman study. Vibrational
Spectroscopy, 2001. 26(2): p. 215-225.

239. Mawhinney, D.B., V. Naumenko, A. Kuznetsova, J.T. Yates, J. Liu, and R.E. Smalley,
Infrared Spectral Evidence for the Etching of Carbon Nanotubes: Ozone Oxidation at
298 K. Journal of the American Chemical Society, 2000. 122(10): p. 2383-2384.

240. Cao, P., K. Xu, J.O. Varghese, and J.R. Heath, The Microscopic Structure of Adsorbed
Water on Hydrophobic Surfaces under Ambient Conditions. Nano Letters, 2011. 11(12):
p. 5581-5586.

241. Hawaiah Imam, M., D. Wan Mohd Ashri Wan, and A. Mohamed Kheireddine, Effect of
varying the amount of binder on the electrochemical characteristics of palm shell
activated carbon. IOP Conference Series: Materials Science and Engineering, 2017.
210(1): p. 012011.

242. Park, K.-K., J.-B. Lee, P.-Y. Park, S.-W. Yoon, J.-S. Moon, H.-M. Eum, and C.-W. Lee,
Development of a carbon sheet electrode for electrosorption desalination. Desalination,
2007. 206(1): p. 86-91.

243. Show, Y. and H. Itabashi, Electrically conductive material made from CNT and PTFE.
Diamond and Related Materials, 2008. 17(4): p. 602-605.

244. Moreira, F.C., S. Garcia-Segura, R.A.R. Boaventura, E. Brillas, and V.J.P. Vilar,
Degradation of the antibiotic trimethoprim by electrochemical advanced oxidation
processes using a carbon-PTFE air-diffusion cathode and a boron-doped diamond or
platinum anode. Applied Catalysis B: Environmental, 2014. 160-161: p. 492-505.

245. Almeida, L.C., S. Garcia-Segura, C. Arias, N. Bocchi, and E. Brillas, Electrochemical


mineralization of the azo dye Acid Red 29 (Chromotrope 2R) by photoelectro-Fenton
process. Chemosphere, 2012. 89(6): p. 751-758.

175
246. Zheng, Q., D.P. Durkin, J.E. Elenewski, Y. Sun, N.A. Banek, L. Hua, H. Chen, M.J.
Wagner, W. Zhang, and D. Shuai, Visible-Light-Responsive Graphitic Carbon Nitride:
Rational Design and Photocatalytic Applications for Water Treatment. Environmental
Science & Technology, 2016. 50(23): p. 12938-12948.

247. Largitte, L. and R. Pasquier, A review of the kinetics adsorption models and their
application to the adsorption of lead by an activated carbon. Chemical Engineering
Research and Design, 2016. 109: p. 495-504.

248. Namasivayam, C. and D. Kavitha, Removal of Congo Red from water by adsorption onto
activated carbon prepared from coir pith, an agricultural solid waste. Dyes and
Pigments, 2002. 54(1): p. 47-58.

249. Hameed, B.H., A.T.M. Din, and A.L. Ahmad, Adsorption of methylene blue onto
bamboo-based activated carbon: Kinetics and equilibrium studies. Journal of Hazardous
Materials, 2007. 141(3): p. 819-825.

250. Huang, H., X. Xiao, B. Yan, and L. Yang, Ammonium removal from aqueous solutions by
using natural Chinese (Chende) zeolite as adsorbent. Journal of Hazardous Materials,
2010. 175(1–3): p. 247-252.

251. Wahab, M.A., S. Jellali, and N. Jedidi, Ammonium biosorption onto sawdust: FTIR
analysis, kinetics and adsorption isotherms modeling. Bioresource Technology, 2010.
101(14): p. 5070-5075.

252. Fan, X., D.J. Parker, and M.D. Smith, Adsorption kinetics of fluoride on low cost
materials. Water Research, 2003. 37(20): p. 4929-4937.

253. Ma, W., F.-Q. Ya, M. Han, and R. Wang, Characteristics of equilibrium, kinetics studies
for adsorption of fluoride on magnetic-chitosan particle. Journal of Hazardous Materials,
2007. 143(1–2): p. 296-302.

254. Gujar, R.B., S.A. Ansari, and P.K. Mohapatra, Highly Efficient Composite Polysulfone
Beads Containing a Calix[4]arene–Monocrown-6 Ligand and a Room Temperature Ionic
Liquid for Radiocesium Separations: Remediation of Environmental Samples. Industrial
& Engineering Chemistry Research, 2016. 55(48): p. 12460-12466.

255. Zhang, M., Q. Gao, C. Yang, L. Pang, H. Wang, H. Li, R. Li, L. Xu, Z. Xing, J. Hu, and
G. Wu, Preparation of Amidoxime-Based Nylon-66 Fibers for Removing Uranium from
Low-Concentration Aqueous Solutions and Simulated Nuclear Industry Effluents.
Industrial & Engineering Chemistry Research, 2016. 55(40): p. 10523-10532.

256. Ho, Y.-S., Review of second-order models for adsorption systems. Journal of Hazardous
Materials, 2006. 136(3): p. 681-689.

257. Ho, Y.S. and G. McKay, The kinetics of sorption of divalent metal ions onto sphagnum
moss peat. Water Research, 2000. 34(3): p. 735-742.

176
258. Ho, Y.S. and G. McKay, Pseudo-second order model for sorption processes. Process
Biochemistry, 1999. 34(5): p. 451-465.

259. Blanchard, G., M. Maunaye, and G. Martin, Removal of heavy metals from waters by
means of natural zeolites. Water Research, 1984. 18(12): p. 1501-1507.

260. Azizian, S., Kinetic models of sorption: a theoretical analysis. Journal of Colloid and
Interface Science, 2004. 276(1): p. 47-52.

261. Bhargava, D.S. and S.B. Sheldarkar, Use of TNSAC in phosphate adsorption studies and
relationships. Literature, experimental methodology, justification and effects of process
variables. Water Research, 1993. 27(2): p. 303-312.

262. Mitsuishi, M. and A. Datyner, Diffusion of methyl orange and its homologs in water and
in micellar solution of dodecyltrimethylammonium bromide. Sen'i Gakkaishi, 1980.
36(4): p. 175-178.

263. El-Khaiary, M.I., G.F. Malash, and Y.-S. Ho, On the use of linearized pseudo-second-
order kinetic equations for modeling adsorption systems. Desalination, 2010. 257(1–3):
p. 93-101.

264. Chaudhuri, H., S. Dash, and A. Sarkar, Fabrication of New Synthetic Routes for
Functionalized Si-MCM-41 Materials as Effective Adsorbents for Water Remediation.
Industrial & Engineering Chemistry Research, 2016. 55(38): p. 10084-10094.

265. Yousef, R.I., B. El-Eswed, and A.a.H. Al-Muhtaseb, Adsorption characteristics of


natural zeolites as solid adsorbents for phenol removal from aqueous solutions: Kinetics,
mechanism, and thermodynamics studies. Chemical Engineering Journal, 2011. 171(3): p.
1143-1149.

266. Hui, K.S., C.Y.H. Chao, and S.C. Kot, Removal of mixed heavy metal ions in wastewater
by zeolite 4A and residual products from recycled coal fly ash. Journal of Hazardous
Materials, 2005. 127(1–3): p. 89-101.

267. Chayes, F., On ratio correlation in petrography. The Journal of Geology, 1949: p. 239-
254.

268. Lin, J. and L. Wang, Comparison between linear and non-linear forms of pseudo-first-
order and pseudo-second-order adsorption kinetic models for the removal of methylene
blue by activated carbon. Frontiers of Environmental Science & Engineering in China,
2009. 3(3): p. 320-324.

269. Khare, S.K., K.K. Panday, R.M. Srivastava, and V.N. Singh, Removal of victoria blue
from aqueous solution by fly ash. Journal of Chemical Technology & Biotechnology,
1987. 38(2): p. 99-104.

177
270. Yahya, M.S., N. Oturan, K. El Kacemi, M. El Karbane, C.T. Aravindakumar, and M.A.
Oturan, Oxidative degradation study on antimicrobial agent ciprofloxacin by electro-
fenton process: Kinetics and oxidation products. Chemosphere, 2014. 117: p. 447-454.

271. Guivarch, E., S. Trevin, C. Lahitte, and M.A. Oturan, Degradation of azo dyes in water
by Electro-Fenton process. Environmental Chemistry Letters, 2003. 1(1): p. 38-44.

272. Pimentel, M., N. Oturan, M. Dezotti, and M.A. Oturan, Phenol degradation by advanced
electrochemical oxidation process electro-Fenton using a carbon felt cathode. Applied
Catalysis B: Environmental, 2008. 83(1–2): p. 140-149.

273. Panizza, M. and M.A. Oturan, Degradation of Alizarin Red by electro-Fenton process
using a graphite-felt cathode. Electrochimica Acta, 2011. 56(20): p. 7084-7087.

274. Hammami, S., N. Oturan, N. Bellakhal, M. Dachraoui, and M.A. Oturan, Oxidative
degradation of direct orange 61 by electro-Fenton process using a carbon felt electrode:
Application of the experimental design methodology. Journal of Electroanalytical
Chemistry, 2007. 610(1): p. 75-84.

275. Özcan, A. and M. Gençten, Investigation of acid red 88 oxidation in water by means of
electro-Fenton method for water purification. Chemosphere, 2016. 146: p. 245-252.

276. Barhoumi, N., N. Oturan, H. Olvera-Vargas, E. Brillas, A. Gadri, S. Ammar, and M.A.
Oturan, Pyrite as a sustainable catalyst in electro-Fenton process for improving
oxidation of sulfamethazine. Kinetics, mechanism and toxicity assessment. Water
Research, 2016. 94: p. 52-61.

277. Haque, E., J.W. Jun, and S.H. Jhung, Adsorptive removal of methyl orange and methylene
blue from aqueous solution with a metal-organic framework material, iron terephthalate
(MOF-235). Journal of Hazardous Materials, 2011. 185(1): p. 507-511.

278. Ni, Z.-M., S.-J. Xia, L.-G. Wang, F.-F. Xing, and G.-X. Pan, Treatment of methyl orange
by calcined layered double hydroxides in aqueous solution: Adsorption property and
kinetic studies. Journal of Colloid and Interface Science, 2007. 316(2): p. 284-291.

279. Zhu, H.Y., R. Jiang, L. Xiao, and G.M. Zeng, Preparation, characterization, adsorption
kinetics and thermodynamics of novel magnetic chitosan enwrapping nanosized γ-Fe2O3
and multi-walled carbon nanotubes with enhanced adsorption properties for methyl
orange. Bioresource Technology, 2010. 101(14): p. 5063-5069.

280. Liu, Y., J. Wei, Y. Tian, and S. Yan, The structure-property relationship of manganese
oxides: highly efficient removal of methyl orange from aqueous solution. Journal of
Materials Chemistry A, 2015. 3(37): p. 19000-19010.

281. Haldorai, Y. and J.-J. Shim, An efficient removal of methyl orange dye from aqueous
solution by adsorption onto chitosan/MgO composite: A novel reusable adsorbent.
Applied Surface Science, 2014. 292: p. 447-453.

178
282. Liu, J., S. Ma, and L. Zang, Preparation and characterization of ammonium-
functionalized silica nanoparticle as a new adsorbent to remove methyl orange from
aqueous solution. Applied Surface Science, 2013. 265: p. 393-398.

283. Eren, E., Adsorption Performance and Mechanism in Binding of Azo Dye by Raw
Bentonite. CLEAN – Soil, Air, Water, 2010. 38(8): p. 758-763.

284. Ravikovitch, P.I., A. Vishnyakov, R. Russo, and A.V. Neimark, Unified Approach to
Pore Size Characterization of Microporous Carbonaceous Materials from N2, Ar, and
CO2 Adsorption Isotherms. Langmuir, 2000. 16(5): p. 2311-2320.

179
APPENDIX A: PRELIMINARY RESULTS-REGENERATION OF IRON AMENDED

CARBON BY HETEROGENEOUS FENTON OXIDATION

A.1. Introduction

Fenton oxidation processes, both homogeneous and heterogeneous Fenton, have been

investigated to regenerate carbon material. The reported regeneration efficiency varied

significantly among different studies. The differences in the carbon materials properties,

preparation method, adsorbates, and regeneration conditions may account for this phenomenon.

The main purpose of this appendix is to repeat experiments of other researchers.

A.2. Materials and methods

A.2.1. Materials

PAC (ColorSorb G5) was obtained from Jacobi CARBON (Columbus, USA). Ferrous sulfate,

heptahydrate (FeSO4•7H2O, ≥99.0%) was purchased from Anachemia (Canada). MO (ACS

reagent grade) was bought from Ricca chemical company (TX, USA). H2O2 solution (30 wt. %

in water, ACS reagent) was purchased from Sigma-Aldrich. Sulfuric acid (H2SO4, 95.0-98.0%)

was got from EM science.

A.2.2. Carbon modification

About 2 g of fresh PAC without any pretreatment was added into 100 mL of deionized water

in an Erlenmeyer flask. Different amount of FeSO4•7H2O was transferred into the flask. The pH

of the solution was adjusted to 3 by using diluted H2SO4 solution. Then, the flask was sealed by

Parafilm, and the solution was kept constant stirring for 24 h. After that, the mixture was filtered

180
by using a filter paper, and the carbon on the filter paper was washed using 1 L deionized water.

The obtained carbon was dried in an oven at 105 °C for overnight. According to the difference in

the Fe loading on carbon surface, 0.6%Fe/C, 2%Fe/C, and 10%Fe/C was prepared. Fresh carbon

was also treated under same conditions in the absence of FeSO4•7H2O, and the obtained carbon

was denoted as treated carbon.

A.2.3. Adsorption

The adsorption kinetics of the fresh carbon and the adsorption isotherms of different carbon

materials were evaluated by using MO as model compound. For adsorption kinetics, 0.0100 g of

fresh carbon was added into 10 mL of ~200 mg/L MO solution, and the MO concentration in

water phase was measured after different adsorption time by using UV-Visible

spectrophotometer (EVOLUTION 220, Thermo scientific, USA). During adsorption isotherm

test, different amount of carbon material was added into 10 mL of ~200 mg/L MO solution.

After two days adsorption, the MO concentration in water phase was measured. All the

adsorption experiments were carried out in batch model in an incubating orbital shaker (VWR,

Canada), the temperature was set at 25 °C, and the shaking speed was 250 rpm.

A.2.4. Regeneration

Before regeneration, the carbon materials were saturated with ~240 mg/g of MO by mixing

0.4 g of carbon with 50 mL of ~2000 mg/L MO solution. ~0.0500 g of MO saturated carbon

material was transferred into a bottle, and then 10 mL of deionized water was added into the

bottle. The bottle was put into the incubating orbital shaker and kept shaking for 30 min to reach

the designed temperature. Then the pH of the solution was adjusted to 3 by adding diluted H2SO4

181
solution, and different amount of H2O2 solution was added into the bottle. The bottle was then

sealed by using Parafilm and kept shaking in the shaker for different time. After that, the mixture

was filtered by using a filter paper, and the obtained solid was dried in the oven at 105 °C for

overnight. The regenerated carbon was used for adsorption again with the carbon dosage of 0.8

g/L.

A.3. Results and discussion

A.3.1. Adsorption isotherms of modified carbon materials

300
Adsorption capacity (mg MO/g Carbon)

250

200 Fresh carbon


Treated carbon
0.6%Fe/C
2%Fe/C
10%Fe/C

150

0
0 20 40 60 80 100 120 140
Equilibrium concentration (mg/L)

Figure A-1. Adsorption isotherms of different carbon material. Adsorption conditions: different
amount of carbon material; 10 mL of ~200 mg/L MO solution; 250 rpm; 25 °C; 2 days
adsorption.

The adsorption isotherms of modified carbon materials were studied as shown in Figure A-1.

There were no significant differences among the isotherms of different carbon materials. The

182
maximum MO capacity of all the carbon materials was ~250 mg/g. The lack of difference on the

isotherms indicated the modification process did not change the carbon surface a lot.

A.3.2. Effect of regeneration conditions

300
Before regeneration -H2O2;+filtration
-H2O2;-filtration +H2O2;+filtration
250 +H2O2;-filtration
Adsorption capacity (mg/g)

200

150

100

50

0
Fresh carbon 0.6%Fe/C
Carbon materials

Figure A-2 Effect of post-regeneration treatment on the adsorption capacity of regenerated


carbon. +/- H2O2: presence/absence of H2O2 during regeneration; +/- filtration: with/without
filtration after regeneration; 25 °C; initial pH 3; 4 h.

In the first set of experiments, the carbon materials were not filtrated out after regeneration.

The whole suspensions were put into an oven set at 105 °C to let the water vaporize, the carbon

materials were further dried overnight in the same oven. As shown in the Figure A-2, the

adsorption capacity of regenerated carbon was higher than the one with filtration. On the other

hand, without filtration, there were no differences between adsorption capacities of regenerated

fresh carbon and 0.6%Fe/C. However, notable differences on the adsorption capacity of these

two carbon materials after regeneration were observed when a filtration was carried out before
183
drying. These results indicated that carbon materials can be regenerated under higher

temperatures especially in the presence of H2O2.

Meanwhile, the effect of temperature on the adsorption capacity of regenerated carbon was

studied as shown in the Figure A-3.

250
Adsorption capacity (mg/g)

200 Initial capacity

150

100

50

0
20 30 40 50 60 70

Temperature (oC)

Figure A-3 Effect of regeneration temperature on the adsorption capacity of regenerated carbon.
Reaction conditions: 0.05 g 0.6%Fe/C loaded with ~240 mg/g MO; 10 mL deionized water; 1
mL 30 wt. % H2O2; initial pH 3; 4 h.

The increase in regeneration temperature tended to increase the adsorption capacity of the

regenerated carbon materials. However, the increase in adsorption capacity was small even

though the temperature was increased from 25 °C to 65 °C.

184
The impact of H2O2 concentration on Fenton oxidation regeneration was also tested as shown

in Figure A-4. The increase in H2O2 concentration can significantly increase the adsorption

capacity of regenerated carbons especially for the iron loaded carbon material.

300
Before regeneration 1.5% H2O2
0% H2O2 3% H2O2
250
Adsorption capacity (mg/g)

0.3% H2O2

200

150

100

50

0
Fresh carbon Treated carbon 0.6%Fe/C
Carbon material

Figure A-4. Effect of initial H2O2 concentration on the adsorption capacity of regenerated
carbon. Reaction conditions: 0.05 g of carbon loaded with ~240 mg/g MO; 10 mL deionized
water; initial pH 3; 4 h.

The influence of solution pH on the regeneration process was also tested with the results

showing in Figure A-5. The pH of the regeneration solution had little effect on the adsorption

capacity of the regenerated carbon material. Slight decrease in adsorption capacity was observed

when the regeneration was carried out at initial pH of 4 compared with those obtained at initial

pH of 2 and 3.

185
250

Adsorption capacity (mg/g)


200 initial capacity

150

100

50

0
2 3 4
Initial pH

Figure A-5. Effect of initial pH on the adsorption capacity of regenerated carbon. Reaction
conditions: 0.05 g 0.6%Fe/C loaded with ~240 mg/g MO; 10 mL deionized water; 1 mL 30 wt.
% H2O2; 65 °C, 1 h.

As indicated in some studies, the H2O2 solution for regeneration was added into the reactor

for several times, because this can avoid the H2O2 spikes in concentration and •OH scavenging to

make the Fenton oxidation more efficiency [120]. The impact of H2O2 addition procedure on the

regeneration process was investigated as shown in Figure A-6. The results indicated that the

H2O2 addition procedure had no effect on the adsorption capacity of regenerated carbon

materials at the tested temperature up to 65 °C. The similar regeneration efficiency of different

H2O2 addition procedure indicated that the H2O2 in solution was not completely consumed after

regeneration.

186
250

Adsorption capacity (mg/g)


200 initial capacity
Add 1 mL H2O2 at beginning
Add 0.25 mL H2O2 every 1 h
150

100

50

0
25°C 50°C 65°C
Temperature

Figure A-6. Effect of H2O2 addition procedure on the adsorption capacity of regenerated carbon.
Reaction conditions: 0.05 g 0.6%Fe/C loaded with ~240 mg/g MO; 10 mL deionized water; 4 h.

A.3.3. Effect of carbon modification

The effect of iron loading on the Fenton oxidation regeneration of carbon materials were also

tested. The results are shown in the following Figure A-7. The highest regeneration efficiency

was obtained by the carbon materials loaded with 0.6% of Fe, which was consistent with that

reported in the literature. Further increase in iron loading did not increase the regeneration

efficiency, and the regeneration efficiency even decreased when the iron loading increased to

10%. Compared with fresh carbon, the treated carbon without addition of any iron also had

improved regeneration efficiency after Fenton oxidation.

187
250
Before regeneration regenerated

Adsorption capacity (mg/g)


200

150

100

50

0
on on e/C e /C e /C
s h carb ed carb 0.6%F 2 %F 1 0 %F
Fre t
Trea

Carbon material

Figure A-7 Effect of different Fe loading on the adsorption capacity of regenerated carbon.
Reaction conditions: 0.05 g of carbon loaded with ~240 mg/g MO; 10 mL deionized water;
initial pH 3; 65 °C; 1 h.

A.4. Conclusions

According to these preliminary results based on Fenton oxidation regeneration process, the

following conclusions can be obtained. (1) The adsorption capacity of regenerated carbon

increased with the increasing of regeneration temperature and initial H2O2. (2) The adsorption

capacity of regenerated carbon increased by adding iron on the carbon surface. (3) The lack of

difference on adsorption capacity of regenerated carbon under different regeneration pH and the

effect of post-regeneration treatment indicated continuous regeneration during the carbon drying

process.

188
A.5. Problems with G5 powder activated carbon

During the preliminary experiments on G5 PAC regeneration, some problems were identified

as follows. First, the difficulty on carbon recovery after adsorption and regeneration. The particle

size of the carbon material is so small that it required at least overnight to settle. The only way to

separate the carbon from water is filtration, and most of the carbon cannot be recovered once

they stick to the filter paper. In addition, the regeneration process may continue when the carbon

was drying in the oven at high temperature. However, these problems can be solved by preparing

some magnetic carbon which can be separated by using a magnetic field.

189
APPENDIX B: ERRONEOUS APPLICATION OF PSEUDO SECOND ORDER

ADSORPTION KINETICS MODEL: IGNORED ASSUMPTIONS AND SPURIOUS

CORRELATIONS

In this appendix, we revisit the pseudo second order model for adsorption kinetics, its

assumptions, and its application to simulated, random and published data. In particular, a widely

used linear form of the pseudo second order model - plotting 𝑡/𝑞𝑡 against 𝑡 - is shown to result in

spurious correlations for typical adsorption experimental data. Depending on the range of data

used, data from pseudo first and third order models can also appear to be well fit by the pseudo

second order model. Inspection of the residual errors, however, indicates that the errors are not

randomly distributed, as they should be. Based on this study, it is recommended to always verify

the assumptions of a model, fit the data with the non-linear form of the model equation, and

inspect the residual plot to determine the goodness of fit.

The content in this appendix has been published in the journal Industrial & Engineering

Chemistry Research (Xiao, Y., Azaiez, J., & Hill, J. M. (2018). Erroneous Application of Pseudo-

Second-Order Adsorption Kinetics Model: Ignored Assumptions and Spurious Correlations.

Industrial & Engineering Chemistry Research, 57(7), 2705-2709.).

B.1. Introduction

Adsorption is an effective and efficient method to remove hazardous substances from water.

To put the performance of an adsorbent in context, adsorption isotherms and adsorption kinetics

are collected. A model is fit to this data for quantification and process design. Many adsorption

kinetics models have been developed based on reaction control (e.g., pseudo first order [49] and

190
pseudo second order [50] models), or diffusion control [48, 51, 247] (e.g., the WM model). One

of the most commonly applied models is the pseudo second order model, which was developed

by Ho and Mckay in 1998 [50]. Despite being developed on the basis of a reaction-controlled

adsorption, the pseudo second order model has been widely applied to various adsorbents

including GAC [78, 248, 249] for which diffusion control would be expected. In addition, it has

also been widely used for different adsorbates, such as single-charged ammonium [250, 251] and

fluoride [252, 253], for which a pseudo second order mechanism is highly unlikely. Researchers

often ignore the assumptions of the model and base the applicability of the model solely on the

obtained determination coefficients (R2). Application of a linear form of the pseudo second order

model, in which 𝑡/𝑞𝑡 is plotted versus t [254-256], often results in very high correlation

coefficients (𝑅 2 >0.99) regardless of the actual model fit.

The fact that so many adsorption processes with widely different adsorbents and adsorbates

were well fit by this one model prompted us to investigate further. In this study, we revisited the

assumptions of the pseudo second order adsorption kinetics model, analyzed the determination

coefficients obtained with various sets of data, and analyzed the applicability of the model to

several sets of literature data.

B.2. Results and discussion

B.2.1. Pseudo second order model development

If the adsorption rate is pseudo second order with respect to the adsorption site, the kinetics

equation can be expressed as [50, 257]:

191
𝑑𝑞𝑡
= 𝑘(𝑞𝑒 − 𝑞𝑡 )2 (B-1)
𝑑𝑡

where 𝑞𝑡 (mg/g) is the amount adsorbed at time 𝑡 (min), 𝑞𝑒 (mg/g) is the equilibrium adsorption

capacity, and 𝑘 (g/mg min) is the adsorption rate constant. The adsorption reaction

corresponding to Eqn. (B-1) may be as follows [257, 258]:

𝑘1
𝐴 + 2𝑆 ⇄ 𝐴𝑆2 (B-2)
𝑘−1

where 𝐴 is the adsorbate, 𝑆 is an adsorption site, and 𝐴𝑆2 is the adsorbate adsorbed on two sites.

With the assumption that the adsorption process is not limited by diffusion, the reaction rate for

Eqn. (B-2) is expressed as:

𝑑𝐶𝐴𝑆2
= 𝑘1 𝐶𝐴 𝐶𝑆2 − 𝑘−1 𝐶𝐴𝑆2 (B-3)
𝑑𝑡

where Ci corresponds to the concentration of species i at adsorption time of t, k1 is the rate

constant for the forward reaction, and k-1 is the rate constant for the reverse reaction. In order to

simplify eqn. (B-3) to eqn. (B-1), two other assumptions must be made: first, the adsorption is

irreversible so that the rate of desorption, 𝑘−1 𝐶𝐴𝑆2 , can be neglected, and second, the

concentration of adsorbate in the liquid remains essentially constant during the adsorption

process (i.e., CA is constant) [259]. The second assumption is rarely valid in studies reported in

the literature. With the addition of the following relationships in eqns. (B-4) and (B-5), eqn. (B-

1) is obtained from eqn. (B-3).

192
𝐶𝐴𝑆2 ∝ 𝑞𝑡 (B-4)

𝐶𝑆 ∝ (𝑞𝑒 − 𝑞𝑡 ) (B-5)

According to Azizian, the pseudo second order model expression, eqn. (B-1), could be

obtained if the adsorption reaction follows a first order mechanism but the adsorbate

concentration in the liquid phase changes significantly during adsorption [260]. Thus, an R2

value close to 1 would be misleading in this situation.

To further illustrate the importance of verifying that the model assumptions are valid, several

sets of literature data were analyzed. Bhargava and Sheldarkar used an unrinsed tamarind nut

shell AC to adsorb phosphate from water with the results plotted in Figure B-1(a) [261]. A non-

linear form of the pseudo second order adsorption model fit the data resulted in an R2 value of

0.9459. The authors used a linear form of this model and obtained an R2 value of 0.998 but stated

that zinc ions had leached from the AC resulting in removal of phosphate mainly by precipitation

of zinc phosphate, and not a second order adsorption process [261]. In a second example, León et

al. investigated the adsorption of MO on GAC from water [78]. As plotted in Figure B-1(b), the

value of R2 was 0.9926 for the pseudo second order model fit. Given the large particle size (2-5

mm) and microporosity of the adsorbent [78], and the low diffusion coefficient of MO (~6.0 ×

10-6 cm2/s) [262], the adsorption process was likely diffusion limited, thus violating one of the

assumptions of the pseudo second order model.

193
(a) (b)

Figure B-1. Adsorption kinetics for (a) phosphate adsorption on unrinsed tamarind nut shell AC
[261], and (b) MO adsorption on GAC [78]. Solid lines are non-linear fits to the pseudo second
order model.

B.2.2. Linearization of the model and subsequent model fitting

Using the initial condition of 𝑞𝑡 = 0 at 𝑡 = 0, the integral form of Eqn. (B-1) is as follows:

𝑞𝑒2 𝑘𝑡
𝑞𝑡 = (B-6)
1 + 𝑞𝑒 𝑘𝑡

In order to obtain the adsorption kinetics parameters, a linear form of Eqn. (B-6) has often

been used. There are many linear forms [263], but eqn. (B-7) has been most widely used [248-

256, 264-266], possibly because this form was proposed when Ho and Mckay developed the

pseudo second order adsorption kinetics model [50].

194
𝑡 1 1
= 2+ 𝑡 (B-7)
𝑞𝑡 𝑘𝑞𝑒 𝑞𝑒

In a typical adsorption experiment, the adsorbed amount increases quickly initially as the

most vacant sites are available. As adsorption progresses, the rate of adsorption decreases as the

equilibrium adsorption capacity is approached. If data is collected at longer times, the variation

of adsorption time (𝑡) is much bigger than the variation of adsorption capacity (𝑞𝑡 ) with the

result that the fit to eqn. (B-7) improves. According to Chayes, R2 for the fit of eqn. (B-7) to data

can be estimated using eqns. (B-8)-(B-12) [267].

2
(𝐶𝑡 − 𝑟𝑞𝑡, 𝑡 𝐶𝑞𝑡 )
𝑅 2𝑡 = 2 (B-8)
𝑞𝑡
,𝑡 𝐶𝑡 + 𝐶𝑞2𝑡 − 2𝑟𝑞𝑡, 𝑡 𝐶𝑡 𝐶𝑞𝑡

𝑛
1
𝑥̅ = ∑ 𝑥𝑖 (B-9)
𝑛
𝑖=1

1 𝑛
𝑠𝑥 = √ ∑ (𝑥𝑖 − 𝑥̅ )2 (B-10)
𝑛 − 1 𝑖=1

𝐶𝑥 = 𝑠𝑥 /𝑥̅ (B-11)

1
∑𝑛𝑖=1(𝑥𝑖 − 𝑥̅ )(𝑦𝑖 − 𝑦̅)
𝑟𝑥𝑦 = 𝑛 − 1 (B-12)
𝑠𝑥 𝑠𝑦

In these equations, 𝑅 2𝑡 , 𝑡 is the determination coefficient calculated for eqn. (B-7), 𝐶𝑡 and 𝐶𝑞𝑡
𝑞𝑡

are the variation coefficients of 𝑡 and 𝑞𝑡 respectively, which are calculated by using eqn. (B-11),

and 𝑟𝑞𝑡, 𝑡 is the correlation coefficient between 𝑡 and 𝑞𝑡 which is calculated according to eqn. (B-

195
12). To illustrate the relationships between these parameters, several sets of simulated data were

generated by changing the relative values of the variation coefficients, 𝐶𝑡 and 𝐶𝑞𝑡 . As shown in

Figure B-2(a), the value of 𝑅 2𝑡 , 𝑡 depends on both the correlation between 𝑡 and 𝑞𝑡 (i.e., 𝑟𝑡,𝑞𝑡 ) and
𝑞𝑡

the relative values of the variation coefficients. For experiments over longer times (i.e., larger Ct

values relative to Cqt), the value of R2 is close to 1 regardless of the correlation between 𝑡 and 𝑞𝑡 .

Figure B-2(a) illustrates how the fit of eqn. (6) to even random data (𝑟𝑡,𝑞𝑡 = 0), can result in high

(>0.8) values of R2 when the change in time is large compared to the change in adsorption

capacity. The curve shapes in Figure B-2(a) vary depending on the relationship between Ct and

Cqt. In an actual experiment, this relationship will change over time. That is, initially qt increases

rapidly (Ct < Cqt) while at longer times, qt is relatively constant (Ct >> Cqt). Figure B-2(b)

illustrates the fit obtained using eqn. (B-7) (i.e., t/qt versus t, red points) with data randomly

distributed between 0-1000 min for t and 100-150 mg/g for 𝑞𝑡 . The value of R2 was 0.9463

despite there being no correlation between the data. In addition, if the relationship between qt and

t is linear (𝑟𝑞𝑡, 𝑡 = 1), the R2 value will be 1.0 for eqn. (B-7).

196
(a) (b)

Figure B-2. Relationships between (a) the coefficient of determination, R2 and 𝑟𝑡,𝑞𝑡 for different
relative ranges of time (𝑡) and adsorption capacity (𝑞𝑡 ), and (b) the fit of eqn. (B-7) to a random
set of data.

To further illustrate the problems with fitting data using eqn. (B-7), two sets of simulated data

were generated using pseudo first order and third order adsorption kinetics models and these data

are shown in Figure B-3(a). The curves are distinct initially but overlap at longer times. By using

the linear form of eqn. (B-7) to fit the simulated data, R2 values greater than 0.99 were obtained

(Figure B-3(b)) for both sets of data. Relying only on the coefficient of determination, the model

fits would incorrectly lead to the conclusion that the adsorption processes all followed pseudo

second order mechanisms.

(a) (b)

Figure B-3. Simulated (a) adsorption kinetics data for pseudo first, second, and third order data
with rate constants of 0.01 min-1, 0.001 L/(mg min), and 0.0001 L2/(mg min), respectively, and
(b) fit to a linear form of the pseudo second order model (eqn. (B-7)), where the data was

197
generated from first (○), second (●), and third (∆) order models in (b). The maximum adsorption
capacity was set to be the same for the three models.

B.2.3. Recommendations for fitting adsorption data

As has been suggested [263, 268], a non-linear fitting method is superior to a linear fitting

method for the pseudo second order model. Using a non-linear regression (SigmaPlot software)

with the pseudo first and third order model data in Figure B-3(a), however, resulted in R2 values

of 0.973 and 0.952, respectively. The values of R2 are not always reported but values greater than

0.95 are generally accepted as a “good” fit. Thus, an inappropriate conclusion could still be

drawn and additional assessments of the fit are required. A residual plot is one type of

assessment. The standardized residuals are calculated by the following equation:

𝑞𝑡 − 𝑞𝑡′
𝑆𝑡𝑎𝑛𝑑𝑎𝑟𝑑𝑖𝑧𝑒𝑑 𝑅𝑒𝑠𝑖𝑑𝑢𝑎𝑙 = (B-13)
𝑠

where 𝑞𝑡′ (mg/g) is the adsorbed amount at time 𝑡 predicted by the model, and s is the standard

deviation of the residuals (mg/g).

If a model, such as the pseudo second order model, truly represents the behavior of the data,

the residuals should be randomly distributed and close to zero. The residuals calculated for the

non-linear fitting of the data in Figure B-3(a) using eqn. (6) are shown in Figure B-4. In this plot,

the poor fits of the pseudo second order model to the data generated from the pseudo first and

third order equations are clear – the residuals are not randomly distributed around zero.

198
Figure B-4. Residual plots for data generated from pseudo first (○), second (●), and third (∆)
order adsorption kinetics models fit non-linearly to the pseudo second order model (eqn. (B-6)).

As further examples, R2 values of 1.000 were obtained for the fits of simulated pseudo first

order data with rate constants of either 0.005 min-1 or 0.0005 min-1, while an R2 value of 0.981

was obtained for simulated data from a WM model [51]. Again, spurious conclusions may be

drawn based only on the R2 values. The residual plots (Figure B-5), however, clearly show that

the standardized residuals are not randomly distributed, and the pseudo second order model is

unsuitable for all three sets of data.

199
Figure B-5. Residual plot for data generated from pseudo first order kinetics model with different
rate constants (▼ k=0.005 min-1; □ k=0.0005 min-1) and WM diffusion model (■) fit non-
linearly to the pseudo second order model (eqn. (B-6)).

Finally, data from the literature was examined. Khare et al. investigated the adsorption of

Victoria blue (VB) dye ([C33H40N3]OH) by fly ash at different initial dye concentrations and

their results are plotted in Figure B-6(a). Using a non-linear fitting method, the R2 values for the

pseudo second order model fit are 0.934, 0.995, and 0995 for initial VB concentrations of

1.0×10-4 M, 4.5×10-4 M, and 8.5×10-4 M, respectively. The residuals for these plots are given in

Figure B-6(b). The residuals are not randomly distributed around zero and so the pseudo second

order model is not a good fit to the data. The authors concluded that the adsorption of the dye

was diffusion limited, consistent with the pseudo second order model not being an appropriate

model for this data [269].

200
(a) (b)

Figure B-6. Adsorption kinetics of Victoria blue dye from aqueous solutions by fly ash at
different initial concentrations (● 1.0×10-4 M, ○ 4.5×10-4 M, and ▲ 8.5×10-4 M): (a) non-linear
fitting to pseudo second order model (eqn. (B-6)), and (b) residual plots. Data from Khare et al.
[269].

B.3. Conclusions

In this appendix, the pseudo second order adsorption kinetics model is developed with the

assumptions explicitly stated. These assumptions - (i) reaction controlled adsorption, (ii)

essentially constant liquid concentration, and (iii) no desorption – must be satisfied before the

model is applied. If the assumptions are not satisfied, the value of the determination coefficient

(R2) will be meaningless, and a high value may be misinterpreted. Similarly, a spuriously high

value of R2 will be obtained if a linear form of the pseudo second order model is used (i.e.,

plotting 𝑡/𝑞𝑡 vs 𝑡). Using a non-linear form of the equation for the fitting should be followed by

a check of the residual plot, and the model only said to be consistent with the data if the residuals

201
are randomly distributed and close to zero. These issues have been illustrated with simulated and

experimental data.

202
APPENDIX C: ELECTRO-FENTON WATER TREATMENT RESULTS IN THE LITERATURE

Some typical research and results on the EF oxidation of organic pollutants in water were listed as follows.

Pollutants C0 Cathode Anode Electrol O2 Fe pH Voltage/curr Efficienc Referen


yte ent y ce
Malachite green 0.5 Graphite-felt Pt 0.05 Compress 0.2 3.0 60/200 mA >95 % [199]
mM cylindri mM ed air 1 mM (540
cal Na2SO4 L/min Fe2+/Fe min)
3+
mesh 100 %
(22min)
*
0.244
min-1 **
Perfluorooctanoat 50 Hierarchically Pt 0.05 M oxygen 0.3-1.5 2- -0.2 ~ -0.6 90.7- [128]
e mg/L porous carbon Na2SO4 mM 6 V (vs SCE) 70.4% (-
coated carbon FeSO4 0.4 V, pH
paper 2-6, 4h)
1.15-0.69
h-1 (-
0.4V) **

Atrazine 0.1 Carbon felt Pt 0.1 M Compress 0.1 3.0 1A 81% (Pt, [152]
mM cylindri Na2SO4 ed air 1 mM 8 h)
cal L/min Fe2+ 97%
mesh (BDD, 8
BDD h)
Cyanuric acid 0.2 Carbon felt Pt 0.1 M Compress 0.1 3.0 1A 4.1% (pt, [152]
mM cylindri Na2SO4 ed air 1 mM 10 h)
cal L/min Fe2+
mesh
203
BDD 90%
(BDD, 10
h)
RO concentrate 120- Graphite felt Pt wire Cl-, 0.6 L/min 0.2 3.0 -0.72 V (vs 62% (3 [200]
150 Ca2+, O2 mM SCE) h)
mg/L and Fe3+
COD Mg2+
(22300
μS/cm)
Leachate 3896 Carbon-PTFE Pt plate Na+, K+, 0.4 L/min 1-40 2.0 10-40 82% (30 [201]
concentrate mg/L electrode Ca2+, Cl- O2 mM - mA/cm2 mA/cm2,
COD (23.4 FeSO4 5.0 20 cm2,
mS/cm) 10 mM
Fe2+, pH
3.0, 6 h)
p-Nitrophenol 4 mM Carbon felt Pt grid - Oxygen 0.5 2.0 -0.5 V (vs 95% (800 [202]
bubbling mM SCE) C)
Fe3+
Herbicides 100 O2-diffusion BDD 0.05 M 20 0.5-2 3.0 100-450 mA 75-97% [151]
mg/L carbon-PTFE Na2SO4 mL/min mM (1 mM
(TOC) O2 Fe2+ Fe2+, 35
°C, 3 h)
2,4- 230 Carbon-PTFE Pt sheet 0.05 M 20 1 mM 3.0 100-450 mA 87% (4 h, [129]
dichlorophenoxya ppm Na2SO4 mL/min FeSO4 450 mA)
cetic acid (2,4-D) O2
4- 40- Carbon-PTFE Pt sheet 0.05 M 20 0.5-2 2- 100-450 mA 80% (100 [130]
Chlorophenoxyac 387 Na2SO4 mL/min mM 6 mA, 6 h,
etic Acid ppm O2 Fe2+ pH 3.0,
35 °C)
Ciprofloxacin ~33 Carbon felt Pt mesh 0.05 M - 0.1 3.0 60-500 mA 91% (400 [270]
hydrochloride mg/L Na2SO4 mM mA, 6 h)
Fe2+

204
Alizarin red (dye) 120 GDE (carbon- Pt grid 0.05 M 20 mL/s 0-2 2- 100-200 mA 90% (4 [227]
mg/L PTFE) Na2SO4 air mM 6 h)
Fe2+
Azo dyes 0.032 Carbon felt Pt sheet - Oxygen 0.5 2 -0.5 V (vs ~80% [271]
mM mM SCE) (5000 C)
Fe3+
Phenol 1 mM Carbon felt Pt grid 0.05 M Compress Co2+, 3 300 mA ~100% [272]
Na2SO4 ed air Cu2+, (7 h, 0.1
Fe2+, mM
Mn2+ Fe2+)
Orange II 0.4-5 Graphite felt Pt sheet NaCl, Oxygen 0.1 3 -0.5 V and - 75% (3 [133]
× 10-5 Na2SO4, mM 1.0 V (vs h)
M NaClO4. Fe3+ SCE)
Orange II 70 GDE (carbon- Pt plate 0.05 M 1.5 L/min Fe2+ 3.0 0.1 A (-1.5 50% (5h) [146]
mg/L PTFE) Na2SO4 air V vs
Ag/AgCl
Basic blue 3 0.2 Carbon Pt gauze 0.1 M 100 0.1 3 0.1 A 91.6% [203]
mM sponge (CS); NaNO3 mL/min mM (CS);
Carbon felt O2 Fe3+ 50.8%
(CF) (CF) (8
h)
Dimethyl 50 CMK-3-PTFE Pt foil 0.1 M 300 0.5 3.0 -0.5 V (vs 75% (1 [165]
phthalate mg/L coated on Na2SO4 mL/min mM SCE) h)
carbon paper O2 Fe2+
Methyl red 100 Graphite- Pt wire 0.1 M 400 0.2 M 3.0 -0.55 V (vs 80% (20 [164]
mg/L PTFE Na2SO4 mL/min Fe2+ SCE) min) *
electrode O2
Alizarin Red S 100- Graphite felt - 0.05 M 1 L/min 0.1-0.3 3.0 100-300 mA 95% (300 [273]
300 Na2SO4 O2 mM mA, 0.2
mg/L Fe2+ mM Fe2+,
3.5 h)

205
2-(2,4- 217 Carbon-PTFE Pt or 0.05 M 20 0.5-2.0 3.0 100-450 mA 63 % [127]
dichlorophenoxy)- mg/L electrode BDD Na2SO4 mL/min mM (450 mA,
propionic acid (2, anode O2 Fe2+ 4h, Pt)
4-DP) 97% (450
mA, 4h,
BDD)
Orange 61 0.53m Carbon felt Pt sheet 0.05 M Air 0.1 3.0 250 mA 98% (6h) [274]
M Na2SO4 mM
Fe2+
Sulfamethazine 0.2 Graphite felt Pt; 0.05 M - 0.2 3.0 2.08 to 4.16 [153]
mM (GF) Ti/RuO2 Na2SO4 mM 41.66 mA mA/cm2,
(28.8 -IrO2 Fe2+ 6 h Pt
mg/L) (DSA); (69.6);
BDD; DSA
GF (62.2);
BDD(91.
9);
GF(70.2)
Humic acids 16.2 Carbon Pt; BDD 0.05 M Air 0.1 3.0 300 mA ~50% [160]
mg/L sponge; Na2SO4 mM (Pt, 9h)
(TOC) Stainless steel Fe2+ ~60%
(BDD,
9h)
Sucralose 0.2 Carbon felt Pt mesh; 0.05 M 0.5 L/min 0.05- 3.0 100-500 mA ~100% at [154]
mM BDD Na2SO4 air 0.5 0.2 mM
mM Fe2+ and
Fe2+ 200 mA
Acid red 88 0.25 Carbon felt Pt gauze 0.075 M air 0.1 3.0 500 mA 87% (2h) [275]
mM Na2SO4 mM
Fe2+
Sulfamethazine 0.20 Carbon felt Pt mesh; 0.05 M 1 L/min 0.5-4 6.1 100-1000 95% (8h) [276]
mM BDD Na2SO4 air g/L mA BDD, 2.0

206
pyrite g/L
(FeS2) pyrite,
300 mA.
Acid Orange II 0.1 CoFe-layered Pt gauze 0.05 M 1 L/min - 2- 40 mA 97% (pH [149]
mM double Na2SO4 air 7.1 3.0, 8h)
hydroxide/car
bon felt

207
APPENDIX D: METHYL ORANGE ADSORPTION MECHANISM PROPOSED IN THE LITERATURE

The adsorption mechanism of MO by carbon materials and oxides in some literature was listed below.

Adsorption
Surface area Solution pHpzc/ zeta
Carbon capacity Mechanism Reference
(m2/g) pH potential
(mg/g)
π-π interaction;
AC nanotubes 534.6 ~7.0 2.0 149 hydrogen bond; [81]
pore filling
ZnO/CuO/γ- Al2O3 ~32 mV 341 electrostatic interaction [87]
acid modified carbon coated
232 6.0 6.6 88 electrostatic interaction [88]
monolith (ACCM)
Physiadsorption
(electrostatic
GAC 520 5.7 4.7 ~35 [78]
and Van der Waals
interactions)
hydrogen bonding;
Copper sulfide nanoparticles electrostatic
1286 5.0 - 122 [85]
loaded on AC soft–soft and dipole–
ion interaction
porous polyimide 635.5 3 - 609.8 Electrostatic interaction [83]
iron terephthalate (MOF-235) - ~4 ~20 mV 477 Electrostatic interaction [277]
calcined layered double Not 317.9 (200
- - chemisorption [278]
hydroxides adjusted mg/L)
m-CS (chitosan )/γ- 32.78 (20
- 4.29 - Electrostatic interaction [279]
Fe2O3/MWCNTs mg/L)
Electrostatic
30.4 (20
MWCNTs 160 4 - interaction; [77]
mg/L)
Interlayer filling

208
331 (1000
mesoporous carbon CMK-3 940 3 3.3 Electrostatic interaction [181]
mg/L)
9 278
AC/Fe3O4 nanoparticle Electrostatic interaction
1110 5.0 - 303 [27]
composite for Fe3O4
291 (150
ZnO–Al2O3 nanocomposite 279 4.5 6.1 Electrostatic interaction [84]
mg/L)
290.55 (100
Electrostatic
mg/L)
A-MnO2 183.57 3.0 - interaction; [280]
1135.57
Physical adsorption
(500 mg/L)
MnO2–graphene– ~105 (100 electrostatic interaction;
173.45 3 - [86]
CNT hybrid material mg/L) physical adsorption
chitosan/MgO composite - 8 50.7 mV 60 (100 mg/L) electrostatic interaction [281]
ammonium-functionalized silica
- 7.0 105.4 electrostatic interaction [282]
nanoparticle
electrostatic forces;
Bentonite - 4.0 6.8 34.34 dipole–dipole [283]
interactions

209
APPENDIX E: COMPARISON OF NITROGEN ADSORPTION ISOTHERM AND

CARBON DIOXIDE ADSORPTION ISOTHERM DATA

Carbon dioxide adsorption isotherms of Fe2O3/G5 and Fe2O3/NORIT were also determined at

0 °C and shown in Figure E-1. The pore structure properties of these two carbon materials were

derived according to the carbon dioxide adsorption isotherms and the nitrogen adsorption

isotherms using the 2D-NLDFT model as shown in Figure E-2 and Table E-1.

Figure E-1. Carbon dioxide adsorption isotherm of Fe2O3 loaded carbon materials.

Table E-1. Comparison of micropore volume of Fe2O3 loaded carbon materials derived from
carbon dioxide adsorption isotherm (0 °C) and nitrogen adsorption-desorption isotherm (-196
°C) based on 2D-NLDFT model.

Micropore volume (cm3/g)


Carbon materials
N2 CO2
Fe2O3/G5 0.26 0.28
Fe2O3/NORIT 0.42 0.41

210
(a) (b)
Figure E-2. Pore size distributions of Fe2O3/G5 (a) and Fe2O3/NORIT (b) derived from carbon

dioxide adsorption isotherm (0 °C) and nitrogen adsorption isotherm (-196 °C) based on 2D-

NLDFT model.

Because carbon dioxide adsorption isotherm is mainly applied for the micropore measurement

[284], the derived micropore volume and micropore size distribution were compared with those

obtained from the nitrogen adsorption isotherm. For both Fe2O3/G5 and Fe2O3/NORIT, the

micropore volume obtained from the carbon dioxide adsorption isotherm was similar to those

obtained from the nitrogen adsorption isotherm. The pore size distribution derived from the

carbon dioxide adsorption isotherm was different compared with those obtained from the

nitrogen adsorption isotherm with the peak shifting to a lower pore size. The differences between

Fe2O3/G5 and Fe2O3/NORIT, however, were distinguishable according to the nitrogen adsorption

isotherm. Therefore, nitrogen adsorption isotherm itself was sufficient for this research.

211
APPENDIX F: FENTON OXIDATION OF METHYL ORANGE

Figure F-1. Adsorption (■) and Fenton oxidation (●) kinetics of MO removal on Fe2O3/NORIT.
~0.025 g of carbon, 500 mL of MO solution at ~30 mg/L at pH 3.0, room temperature (~20 °C).
0.3 wt.% of H2O2 was applied during Fenton oxidation process.

Figure F-3. Adsorption (■) and Fenton oxidation (●) kinetics of MO removal on Fe2O3/G5.
~0.025 g of carbon, 500 mL of MO solution at ~30 mg/L at pH 3.0, room temperature (~20 °C).
0.3 wt.% of H2O2 was applied during Fenton oxidation process.

212
Figure F-4. Adsorption (■) and Fenton oxidation (●) kinetics of MO removal on Fe2O3/SMC.
~0.025 g of carbon, 500 mL of MO solution at ~30 mg/L at pH 3.0, room temperature (~20 °C).
0.3 wt.% of H2O2 was applied during Fenton oxidation process.

213
APPENDIX G: COPYRIGHT PERMISSIONS

214
215
216

You might also like