207
University of Calgary PRISM: University of Calgary's Digital Repository Graduate Studies The Vault: Electronic Theses and Dissertations 2014-04-28 Light Emitting Diode Based Photochemical Treatment of Contaminants in Aqueous Phase Yu, Linlong Yu, L. (2014). Light Emitting Diode Based Photochemical Treatment of Contaminants in Aqueous Phase (Unpublished doctoral thesis). University of Calgary, Calgary, AB. doi:10.11575/PRISM/26762 http://hdl.handle.net/11023/1447 doctoral thesis University of Calgary graduate students retain copyright ownership and moral rights for their thesis. You may use this material in any way that is permitted by the Copyright Act or through licensing that has been assigned to the document. For uses that are not allowable under copyright legislation or licensing, you are required to seek permission. Downloaded from PRISM: https://prism.ucalgary.ca

1.3 Light emitting diode

  • Upload
    others

  • View
    12

  • Download
    0

Embed Size (px)

Citation preview

Page 1: 1.3 Light emitting diode

University of Calgary

PRISM: University of Calgary's Digital Repository

Graduate Studies The Vault: Electronic Theses and Dissertations

2014-04-28

Light Emitting Diode Based Photochemical Treatment

of Contaminants in Aqueous Phase

Yu, Linlong

Yu, L. (2014). Light Emitting Diode Based Photochemical Treatment of Contaminants in Aqueous

Phase (Unpublished doctoral thesis). University of Calgary, Calgary, AB.

doi:10.11575/PRISM/26762

http://hdl.handle.net/11023/1447

doctoral thesis

University of Calgary graduate students retain copyright ownership and moral rights for their

thesis. You may use this material in any way that is permitted by the Copyright Act or through

licensing that has been assigned to the document. For uses that are not allowable under

copyright legislation or licensing, you are required to seek permission.

Downloaded from PRISM: https://prism.ucalgary.ca

Page 2: 1.3 Light emitting diode

UNIVERSITY OF CALGARY

Light Emitting Diode Based Photochemical Treatment of Contaminants in Aqueous Phase

by

Linlong Yu

A THESIS

SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

IN PARTIAL FULFILMENT OF THE REQUIREMENTS FOR THE

DEGREE OF DOCTOR OF PHILOSOPHY

DEPARTMENT OF CIVIL ENGINEERING

CALGARY, ALBERTA

April, 2014

© Linlong Yu 2014

Page 3: 1.3 Light emitting diode

ii

Abstract

In this research, photochemical treatment of pesticides and polychlorinated biphenyls

(PCBs) in aqueous medium were investigated. The studies on photochemical treatment

of these two groups of compounds, along with radiation field modelling, further, led to

the design of an efficient light emitting diode (LED) based flow-through photocatalytic

reactor.

Sensitized photodechlorination of PCBs in surfactant solutions was studied. Three types

of surfactants at different concentrations were investigated. The neutral and cationic

surfactants were found to be more effective than the anionic one. In each case the

surfactant concentration was found to play a significant role in the rate of dechlorination.

LED based photocatalytic degradation of pesticides and chlorophenols, namely 2,4-

dichlorophenoxyacetic acid (2,4-D), 2-methyl-4-chlorophenoxyacetic acid (MCPA) , 4-

chlorophenol (4-CP) and 2,4-dichlorophenol (2,4-DCP) was studied. Further, the impact

of photocatalyst loading and light intensity on the degradation rate was evaluated. The

degradation of 2,4-D under LED irradiation was compared to that with mercury discharge

lamp irradiation. The results show these compounds can be efficiently degraded using

LED based TiO2 photocatalysis. They are completely mineralized upon prolonged

irradiation. Our results indicate that LEDs are a better light source than the mercury

lamps.

Page 4: 1.3 Light emitting diode

iii

To design an efficient LED based photocatalytic reactor, a radiation field model was

developed in this research. The model was tested with experimental data and good

agreement between two was noted. The model can be used to optimize the photoreactor

and chose the optimal gap between adjacent LEDs, the irradiated distance and the light

output of LEDs for a homogenous radiation field.

Finally, an LED based photocatalytic reactor was designed and fabricated. The reactor

uses anodized TiO2 nanostructure as a photocatalyst. The performance of reactor was

evaluated and optimized by studying the degradation of 2,4-D. The effect of different

operational parameters on the reactor performance were investigated, including light

intensity, distance between the LED module and photocatalytic plate (DL-P), the flow rate

through the reactor, presence of external electron scavengers and photocatalyst

configuration. A power law relationship was observed between the light intensity (2.2

mW cm-2~17.3 mW cm-2) and the first order degradation rate constant for 2,4-D. A

suitable flow rate and DL-P was determined for the reactor. Enhanced performance of the

reactor was observed where electron scavengers were introduced.

Page 5: 1.3 Light emitting diode

iv

Acknowledgements

I would like to express my sincerest appreciation and gratitude to my supervisor Dr.

Gopal Achari and my co-supervisor Dr. Cooper H. Langford for their continuous

encouragement, intellectual advice, precious guidance and enthusiastic supports

throughout my doctoral program.

It is my fortune to have friendly colleagues, Dr. Jyoti Ghosh, Dr. Maryam Izadifard,

Jiansong Kong, Chien-Kai Kenneth Wang, Upasana Chamoli and Mitra mehrabani. I

greatly appreciate their helps. My gratitude is also extended to Mr. Daniel Larson for his

assistance with instruments and laboratory facilities during my research. Thanks to Mr.

Edward C. Cairns, Mr. Andrew Read, Mr. Mark Toonen and Robert Thomson for their

help on fabricating LED reactors.

I gratefully acknowledge the financial support provided by Samuel Hanen Foundation,

RES'EAU WaterNet Strategric Research Network, Natural Science and Engineering

Council of Canada and Department of Civil Engineering.

Finally, I would like to show my gratitude to my sister, my uncles and my aunties for

their supports in the past five years.

Page 6: 1.3 Light emitting diode

v

Dedication

This thesis is dedicated to my beloved parents.

Page 7: 1.3 Light emitting diode

vi

Table of Contents

Abstract ............................................................................................................................... ii Acknowledgements ............................................................................................................ iv Dedication ............................................................................................................................v Table of Contents ............................................................................................................... vi List of Tables .......................................................................................................................x List of Figures and Illustrations ......................................................................................... xi List of Symbols, Abbreviations and Nomenclature ...........................................................xv

CHAPTER ONE: INTRODUCTION ..................................................................................1 1.1 Background ................................................................................................................1 1.2 Photochemical Treatment Processes ..........................................................................3 1.3 Light emitting diode (LED) in photocatalytic reactors ..............................................4 1.4 Research Objectives and Scopes ................................................................................5 1.5 Thesis Overview ........................................................................................................6

CHAPTER TWO: LITERATURE REVIEW ......................................................................9 2.1 Principle of Photochemistry.......................................................................................9

2.1.1 Light and photon ................................................................................................9 2.1.2 The electronic excited states ............................................................................11 2.1.3 Quantum yield .................................................................................................11 2.1.4 Direct photolysis ..............................................................................................12 2.1.5 Photosensitized degradation ............................................................................13 2.1.6 Photocatalysis ..................................................................................................13 2.1.7 Advanced Oxidation Processes (AOPs) ..........................................................14

2.2 TiO2 Photocatalysis .................................................................................................14 2.2.1 TiO2 as a photocatalyst....................................................................................14 2.2.2 Mechanism of TiO2 photocatalysis .................................................................17 2.2.3 The kinetics of photocatalytic degradation ......................................................20 2.2.4 Factors affecting the photocatalytic degradation kinetics ...............................21

2.2.4.1 TiO2 loading ..........................................................................................21 2.2.4.2 Light intensity ........................................................................................22 2.2.4.3 pH ...........................................................................................................23 2.2.4.4 Electron acceptor ...................................................................................24 2.2.4.5 Hole/hydroxyl radicals scavenger ..........................................................26

2.3 Contaminants (Pesticides and PCBs) .......................................................................27 2.3.1 Pesticides .........................................................................................................27

2.3.1.1 2,4-dichlorophenoxyacetic acid (2,4-D) ...............................................29 2.3.1.2 2-methyl-4-chlorophenoxyacetic acid (MCPA) ....................................30 2.3.1.3 Chlorophenols ........................................................................................32

2.3.2 PCBs ................................................................................................................33 2.4 Photochemical treatment of pesticides and PCBs ....................................................36

2.4.1 Direct photolytic degradation of pesticides and PCBs ....................................36 2.4.2 Photosensitized degradation of pesticides and PCBs ......................................37 2.4.3 Photocatalytic degradation of pesticides and PCBs ........................................37

Page 8: 1.3 Light emitting diode

vii

2.5 Design of a photocatalytic reactor ...........................................................................38 2.5.1 State of Photocatalyst in the Reactor ...............................................................38

2.5.1.1 Slurry photocatalytic reactor vs immobilized photocatalytic reactor ....38 2.5.1.2 TiO2 immobilization through electrochemical anodization ..................40

2.5.2 Light source .....................................................................................................41 2.5.2.1 Sunlight ..................................................................................................42 2.5.2.2 Mercury lamps .......................................................................................44 2.5.2.3 Light emitting diode ...............................................................................45

2.5.3 Artificially illuminated photocatalytic reactors ...............................................47 2.5.4 Solar photocatalytic reactors: ..........................................................................53

2.6 Radiation-field modelling ........................................................................................56 2.6.1 The Radiation transport equation (RTE) .........................................................57 2.6.2 Numerical methods to solve the RTE ..............................................................60 2.6.3 Radiation source models ..................................................................................61

CHAPTER THREE: ELECTRON TRANSFER SENSITIZED PHOTODECHLORINATION OF SURFACTANT SOLUBILIZED PCB138 .......63

3.1 Introduction ..............................................................................................................63 3.2 Materials and methods .............................................................................................64

3.2.1 Materials ..........................................................................................................64 3.2.2 Methods ...........................................................................................................65

3.2.2.1 PCB 138 solubilization with surfactants ................................................65 3.2.2.2 Photochemical reaction ..........................................................................65 3.2.2.3 Sampling, extraction and GC analysis ...................................................66

3.3 Results and Discussion ............................................................................................67 3.3.1 Selectivity of surfactants .................................................................................67 3.3.2 Dechlorination of PCBs in TEA and NaBH4 systems ....................................72

3.3.2.1 MB and TEA ..........................................................................................72 3.3.2.2 MB and NaBH4 .....................................................................................73 3.3.2.3 Photodegradation of Aroclor 1254 with NaBH4 and TEA ....................74

3.3.3 The dechlorination pathways of PCB 138 using CTAB and TWEEN 80 .......75 3.4 Conclusions ..............................................................................................................77

CHAPTER FOUR: LED-BASED PHOTOCATALYTIC TREATMENT OF PESTICIDES AND CHLOROPHENOLS ...............................................................79

4.1 Introduction ..............................................................................................................79 4.2 Methods and Materials .............................................................................................81

4.2.1 Photoreactor .....................................................................................................81 4.2.2 Chemicals ........................................................................................................83 4.2.3 Photocatalytic degradation ..............................................................................83 4.2.4 Actinometric Experiment ................................................................................84 4.2.5 Analysis of sample ..........................................................................................85

4.2.5.1 HPLC Analysis ......................................................................................85 4.2.5.2 TOC Analysis: .......................................................................................85

4.3 Results and Discussions ...........................................................................................86 4.3.1 Photocatalytic degradation of pesticides and chlorophenols ...........................86 4.3.2 Photocatalytic degradation of pesticides mixtures ..........................................89

Page 9: 1.3 Light emitting diode

viii

4.3.3 Effect of Photocatalyst Loading ......................................................................93 4.3.4 Effect of Light Intensity ..................................................................................95 4.3.5 Comparison between LED and Mercury Lamp Irradiation .............................97

4.4 Conclusions ..............................................................................................................99

CHAPTER FIVE: DESIGN A HOMOGENEOUS RADIATION FIELD IN A UV-LED BASED PHOTOCATALYTIC REACTOR ..................................................100

5.1 Introduction ............................................................................................................100 5.2 Advantage of homogeneous radiation field in a photocatalytic reactor ................101 5.3 Development of radiation field model ...................................................................103

5.3.1 UV-LED array and photocatalyst plate .........................................................103 5.3.2 Radiation field model without shielding glass plate ......................................104 5.3.3 Radiation field model with a shielding glass plate ........................................109

5.4 Calibration and validation of the radiation field model .........................................110 5.4.1 Light intensity measurement .........................................................................110 5.4.2 Model light intensities vs measured light intensities .....................................112

5.5 Design of a homogenous radiation filed ................................................................114 5.5.1 The effect of ID on the homogeneity of radiation field for a fixed gap ........114 5.5.2 Optimal combination of ID and gap ..............................................................116 5.5.3 Selection of the output of the UV-LED .........................................................117

5.6 Conclusions ............................................................................................................118

CHAPTER SIX: A NOVEL LIGHT EMITTING DIODE BASED PHOTOCATALYTIC REACTOR FOR WATER TREATMENT ........................119

6.1 Introduction ............................................................................................................119 6.2 Experimental details ..............................................................................................121

6.2.1 Chemicals ......................................................................................................121 6.2.2 Design and fabrication of an LED based photocatalytic reactor ...................121

6.2.2.1 Preparation of anodized TiO2 photocatalytic plate. .............................121 6.2.2.2 UV-LEDs module ................................................................................122 6.2.2.3 Photocatalytic system ..........................................................................123

6.2.3 Radiation field and light intensity estimation ................................................124 6.2.4 Experimental set-up and sample analysis ......................................................126

6.3 Result and discussion .............................................................................................127 6.3.1 Degradation of phenoxy pesticides and chlorophenols in a flow-through

LED based photocatalytic reactor ..................................................................127 6.3.2 Degradation of 2,4-D with different combination of (UV, TiO2

photocatalyst plate, H2O2 and O2) in the UV-LED photoreactor . ...............128 6.3.3 Effect of DL-P .................................................................................................130 6.3.4 Effect of flow rates on the photocatalytic degradation of 2,4-D. ..................131 6.3.5 Effect of UV light intensity ...........................................................................133 6.3.6 Comparison of three different photocatalyst configurations .........................135

6.4 Conclusions ............................................................................................................136

CHAPTER SEVEN: CONCLUSION AND RECOMMENDATION FOR FUTURE RESEARCH ............................................................................................................137

7.1 Conclusions ............................................................................................................137

Page 10: 1.3 Light emitting diode

ix

7.1.1 Photosensitized dechlorination of PCBs solubized in surfactant solution ....137 7.1.2 LED based photocatalytic treatment of pesticides and chlorophenols ..........138 7.1.3 Design a homogenous radiation field model for photocatalytic reactor ........138 7.1.4 A novel light emitting diode photocatalytic reactor for water treatment ......139

7.2 Recommendations for Future Research .................................................................140 7.2.1 Incorporating PCBs extraction using surfactants and PCBs

photodechlorination using sensitized visible light .........................................140 7.2.2 UVC-LED ......................................................................................................140 7.2.3 The decay of photocatalytic activity and its life time ....................................140 7.2.4 Hollow microsphere coated with TiO2 (HGMT)...........................................140 7.2.5 Scale-up of the reactor ...................................................................................141

REFERENCES ................................................................................................................142

APPENDIX A: INVESTIGATION OF ULTRTRASONIC EXTRACTION OF POLYCHORINATED BIPHENYLS FROM SOIL ...............................................177

A.1. Experimental ........................................................................................................177 A.1.1. Chemicals ....................................................................................................177 A.1.2. Pre-Processing of contaminated soil ............................................................177 A.1.3. Ultrasonic extraction of PCBs .....................................................................177 A.1.4. Soxhlet extraction of the remaining PCBs in soil : ....................................178 A.1.5. Calculation of ultrasonic extraction efficiency ............................................178 A.2. Results and Discussions ..................................................................................179

A.3. Reference .............................................................................................................180

APPENDIX B: INVESTIGATION OF PHOTODEGRADATION OF BIPHENYL IN ULTRAVIOLET WATER PURIFICATION SYSTEMS.................................181

B.1. Experimental ........................................................................................................181 B.1.1. Chemicals .....................................................................................................181 B.1.2. Photoreactor .................................................................................................181 B.1.3. Photodegradation of biphenyl in IPA ..........................................................181

B.2. Results and discussions ........................................................................................182

APPENDIX C: UV VIS ABSORPTION SPECTRUM OF DIFFERENT PESTICIDES ..........................................................................................................184

APPENDIX D: THE CALCULATION OF PERCENTAGE OF AVAILABLE PHOTONIC ENERGY FOR PHOTOCATALYTIC REACTION ........................186

Page 11: 1.3 Light emitting diode

x

List of Tables

Table 2-1: Definitions of the quantum yield (Oppenlander, 2003). ................................. 12

Table 2-2: Generation of hydroxyl radicals for different AOPs. ...................................... 15

Table 2-3: Sales/use of the top 20 pesticide active ingredient in Canada (Brimble et al., 2005). .................................................................................................................. 28

Table 2-4: Slurry versus Immobilized Photocatalytic Systems (Lasa et al., 2005). ........ 39

Table 3-1: First order rate coefficients (K) for PCB138 dechlorination in TEA-MB system with different concentration of surfactants. .................................................. 71

Table 3-2: First order rate coefficients (K) for PCB138 dechlorination in NaBH4-MB system. ...................................................................................................................... 74

Table 4-1: Light intensity of different photoreactors. ....................................................... 84

Table 4-2: First order rate coefficients (K) for photocatalytic pegradation of different pesticides. .................................................................................................................. 89

Table 4-3: Mixtures of Pesticides. .................................................................................... 89

Table 4-4: Percentage removal of pesticides at 0.028 kJ energy dosage. ......................... 91

Table 4-5: Percentage of pesticides adsorbed on the surface of TiO2 after 30 minutes of stirring in the dark. ................................................................................................ 93

Table 5-1: Parameters used for radiation field model calculation. ................................. 113

Table 5-2: Comparison of modeled light intensity and measured light intensity. .......... 113

Table 6-1: Average light intensity received by the photocatalytic plate. ....................... 124

Table 6-2: First order kinetic rate constants for different photocatalyst configurations . 135

Page 12: 1.3 Light emitting diode

xi

List of Figures and Illustrations

Figure 2-1: Classification of electromagnetic radiation in the wavelength range below 1200 nm. [Reproduced from (Oppenlander, 2003) with the permission]. ................ 10

Figure 2-2: Phenomenological subdivision of ultraviolet radiation into four sub-bands and their characteristic effects. [Reproduced from (Oppenlander, 2003) with the permission]. ............................................................................................................... 10

Figure 2-3: Photochemical activation of TiO2. ................................................................. 19

Figure 2-4: Molecular structure of 2,4-dichlorophenoxyacetic acid. ............................... 29

Figure 2-5: Molecular structure of 2-methyl-4-chlorophenoxyacetic acid. ...................... 31

Figure 2-6: Molecular structure of chlorophenols. ........................................................... 32

Figure 2-7: Molecular structure of polychlorinated biphenyls. ........................................ 34

Figure 2-8: Solar spectral irradiance distribution on the surface of earth. [Reproduced from (Hulstrom et al., 1985) with permission] ......................................................... 42

Figure 2-9: Fractional cumulative integrated irradiance vs. wavelength. [Reproduced from (Hulstrom et al., 1985) with permission] ......................................................... 43

Figure 2-10: An inner working on an LED. [Adapted from (Wikipedia, 2011)] ............. 46

Figure 2-11: Various photochemical reactor configurations. [Reproduced from (Pareek et al., 2008) with permission] ...................................................................... 48

Figure 2-12: Scheme of a multiple tube reactor. [Reproduced from (Ray and Beenackers, 1998) with permission] ......................................................................... 50

Figure 2-13: Scheme of an optical fibre photocatalytic reactor. [Reproduced from (Nguyen and Wu, 2008) with the permission] .......................................................... 51

Figure 2-14: Scheme of a rotating disk reactor. [Reproduced from (Hamill et al., 2001) with the permission] ....................................................................................... 51

Figure 2-15: Top view of a distributive photocatalytic reactor. [Reproduced from (Ray and Beenackers, 1998) with permission] ......................................................... 52

Figure 2-16: Experimental setup of taylor vortex photocatalytic reactor:(1) motor, (2) speed controller, (3) gear coupling, (4) UV lamp (5) sample collection point (6) lamp holder (7) outer cylinder and (8) catalyst-coated inner cylinder. [Reproduced from (Dutta and Ray, 2004) with the permission]............................... 52

Page 13: 1.3 Light emitting diode

xii

Figure 2-17: Scheme of a fluidized photocatalytic reactor. [Reproduced from (Vaisman et al., 2005) with permission] ................................................................... 53

Figure 2-18: Solar photocatalytic reactor: (a) parabolic trough reactor (PTR) (b) compound parabolic collector (CPC). [Reproduced from (Braham and Harris, 2009) with permission] ............................................................................................. 54

Figure 2-19. Typical reactor layout for an (a) inclined plate collector and (b) double skin sheet photoreactor. [Reproduced from (Braham and Harris, 2009) with permission] ................................................................................................................ 55

Figure 2-20: Typical reactor layout for (a) horizontal rotating disk reactor and (b) water bell reactor. [Reproduced from (Braham and Harris, 2009) with permission] ................................................................................................................ 56

Figure 2-21: Schematic for photon transport. .................................................................. 57

Figure 3-1: Reductive dechlorination of PCB 138 using LMB with TEA as the reducing agent; [PCB 138] = 6.6 mg L-1, [SDS] = [CTAB] = [TWEEN 80] = 3.2 g L-1, [MB] = 750 mg L-1, [TEA] = 68 g L-1, Io = 5.2×1016 photon s-1 . ................... 68

Figure 3-2: Reductive dechlorination of PCB 138 using LMB with NaBH4 as the reducing agent; [PCB 138] = 20 mg L-1, [SDS] = [CTAB] = [TWEEN 80] = 3.2 g L-1; [MB] = 600 mg L-1; [NaBH4] = 20 g L-1, Io = 3.0×1016 photon s-1

. ................ 69

Figure 3-4: Dechlorination of Aroclor1254 solubilized with TWEEN 80 in the presence of MB and TEA or NaBH4: [Aroclor1254] = 10 mg L-1, [MB] = 600 mg L-1, [TWEEN80] = 1.6 g L-1, [TEA] = 108 g L-1, [NaBH4] = 20 g L-1, Io = 3.0 ×1016 photon s-1. ........................................................................................................ 75

Figure 3-5: The product distribution (after six minutes irradiation) for dechlorination of PCB 138 solubilized by TWEEN 80 (1.6 g L-1) or CTAB (1.6 g L-1) in the presence of MB (600 mg L-1) and TEA (68 g L-1), Io = 3.0 ×1016 photon/s, P: peak area of each congener from GC, Po: the peak area of initial PCB 138. ........... 76

Figure 3-6: The product distribution (after six minutes irradiation) for dechlorination of PCB 138 solubilized by TWEEN 80 (1.6 g L-1) or CTAB (1.6 g L-1) in the presence of MB (600 mg L-1) and NaBH4 (10 g L-1); Io = 3.0 ×1016 photon s-1, P: peak area of each congener from GC, Po: the peak area of initial PCB 138. ........... 77

Figure 4-1: LED photoreactor and insert. ......................................................................... 82

Figure 4-2: Photocatalytic degradation of different pesticides with UV-LED photoreactor (Io = 8.55×1016 photon s-1, CTiO2=2.0 g L-1, Co=20 mg L-1): (a) loss of parent pesticides; and (b) loss of total organic carbon. ........................................ 87

Figure 4-3: Photocatalytic degradation of pesticides mixture with UV-LED photoreactor based on the loss of pesticides detected by HPLC (Io =8.55×1016

Page 14: 1.3 Light emitting diode

xiii

photon s-1, CTiO2=2 g L-1, Co=20 mg L-1): (a) mixture containing 4-CP and 2,4-DCP; (b) mixture containing 4-CP and 2,4-D; (c) mixture containing 2,4-DCP and 2,4-D. .................................................................................................................. 90

Figure 4-4: Photocatalytic degradation of 2,4-D with different TiO2 loadings and LED irradiation (Io =8.55×1016 photon s-1, Co=20 mg L-1). ..................................... 94

Figure 4-5: Photocatalytic degradation of 2,4-D with UV-LED photoreactor under different light conditions (Co=20 mg L-1, CTiO2=2 g L-1). ......................................... 96

Figure 4-6: Photocatalytic degradation of 2,4-D in the two photoreactors: Co=20 mg L-1, CTiO2=2 g L-1. (a): LED reactor, Io =8.55×1016 photon s-1; (b): Rayonet reactor, Io =8.25×1016 photon s-1. ............................................................................. 98

Figure 5-1: UV-LED array and photocatalyst plate. ....................................................... 103

Figure 5-2: Directivity of radiation (NICHIA, 2013). ................................................... 105

Figure 5-3: Cartesian and polar coordinates in radiation system. ................................... 107

Figure 5-4: Scheme of UV-LED radiation. .................................................................... 109

Figure 5-5: Geometry of sensor. ..................................................................................... 111

Figure 5-6: The radiation field with different ID: (a) ID=0.01 m, gap=0.025 m; (b) ID= 0.04 m, gap=0.025 m. ...................................................................................... 115

Figure 5-7: The effect of irradiated distance (ID) on Maximum Error........................... 116

Figure 5-8: Optimal combination of ID and gap. ........................................................... 116

Figure 5-9: Selection of light output of UV-LED. .......................................................... 117

Figure 6- 1: SEM image of anodized TiO2 nanostructrure. ............................................ 122

Figure 6-2: Scheme of an LED based photocatalytic reactor. ........................................ 123

Figure 6-3: Radiation field on a photocatalyst plate under different conditions; (a) DL-

P = 0.014 m, 4 by 4 LEDs panel; (b) DL-P = 0.034 m, 4 by 4 LEDs panel; (c) DL-P = 0.054 m, 4 by 4 LEDs panel. ............................................................................... 125

Figure 6-4: Photodegradation of MCPA, 2,4-D, 2,4-DCP and 4-CP in a UV-LED photoreactor: flow rate = 2.03 L min-1; DL-P = 0.54 cm; Ia=17.3 mW cm-2. .......... 128

Figure 6-5: Photodegradation of 2,4-D in a flow-through UV-LED photoreactor: flow rate = 2.03 L min-1; DL-P = 0.54 cm ; Ia=17.3 mW cm-2. ........................................ 129

Figure 6- 6: The effect of DL-P on 2,4-D degradation: flow rate =2.03 L min-1, Ia=17.3 mW cm-2. ................................................................................................................. 131

Page 15: 1.3 Light emitting diode

xiv

Figure 6-7: The effect of flow rate on degradation of 2,4-D: DL-P = 5.4 cm, Iaverage=17.3 mW cm-2. ............................................................................................ 132

Figure 6-8: The effect of light intensity on degradation of 2,4-D: DL-P = 5.4 cm, Flow rate=2.03 L min-1. ................................................................................................... 134

Figure 7-1: Scheme of a scale-up LED based photocatalytic reactor. ............................ 141

Figure A-A-1: Ultrasonic extraction efficiency of PCBs from 10 g soil under different experimental conditions. ......................................................................................... 180

Figure A-B-1: Ultraviolet water purification system. ..................................................... 181

Figure A-B-2: Degradation of biphenyl under different flow rate. ................................ 183

Figure A-B-3: The pseudo first order kinetics of biphenyl degradation under different flow rate. ................................................................................................................. 183

Figure A-C-1: UV-Vis absorption spectra of 40 mg/L of 2,4-D in water. ..................... 184

Figure A-C-2: UV-Vis absorption spectra of 40 mg/L of 2,4-DCP in water. ................ 184

Figure A-C-3: UV-Vis absorption spectra of 40 mg/L of 4-CP in water. ...................... 185

Figure A-C-4: UV-Vis absorption spectra of 40 mg/L of MCPA in water. ................... 185

Figure A-D-1: The emission spectrum and TiO2 band edge. ......................................... 187

Page 16: 1.3 Light emitting diode

xv

List of Symbols, Abbreviations and Nomenclature

Symbol Definition

A Cross surface area, m2

Ap Area of photocatalyst plate, m2

c Light speed in vacuum, m s-1

C Concentration of substrate, mole L-1

Co Initial concentration of parent compound, mg L-1

d Distance, m

Ds-p The distance between shielding glass and photocatalyst, cm

Dl-p The distance between LEDs and photocatalyst, cm

do Specific distance, m

E Energy of photon, J

g Distance between point (x, y) and point (xo, yo), m

Gv Incident light intensity, photon s-1 m-2

h Plank constant, 6.62*10-34 J s

I Light intensity, mw cm-2

Ia Average light intensity, mw cm-2

Imax Maximum of light intensity, mw cm-2

Imeasured Light intensity measured by UV meter, mw cm-2

Imin Minimum of light intensity, mw cm-2

Imodel Light intensity calculated by model, mw cm-2

Io Light intensity measured by actinometry, photon s-1

It Light output of an LED lamp, mw

Iλ Specific light intensity, photon s-1 m-2

k Apparent reaction rate constant, mol L-1 s-1

K First order rate coefficient, s-1

Ka The average of first order rate coefficient, s-1

Kad Adsorption coefficient, L mol-1

lg The thickness of glass plate, m

Me Max error

Page 17: 1.3 Light emitting diode

xvi

p(Ω-Ω') Phase function for scattering in RTE

Q Number of photons

qa Rate of photon absorption, photon s-1

qe Rate of photon emission, photon s-1

qin Rate of photon in-scattered, photon s-1

qout Rate of photon out-scattered, photon s-1

r Radius, m

R Radial distance, m

Re Radiation directivity function

Re' Modified radiation directivity function

rp Kinetic reaction rate, mole L-1 s-1

rs Radius of the sensor, m

s Surface area, m2

S Direction vector, m

t Time, s

T Transmittance

v Frequency, s-1

V Elementary control volume, m3

Wa Volumetric rate of photon absorption, photon s-1 m-3

We Volumetric rate of photon emission, photon s-1 m-3

Win Volumetric rate of photon in-scattered, photon s-1 m-3

Wout Volumetric rate of photon out-scattered, photon s-1 m-3

x x-coordinates

xo x-coordinates of LED position

y y-coordinates

yo y-coordinates of LED position

z z-coordinates

α Volumetric absorption coefficient, m-1

β Extinction coefficient, m-1

γ Constant

η Constant

Page 18: 1.3 Light emitting diode

xvii

θ View angle, radian

λ Wavelength

λmax Wavelength of maximum emission, nm

σ Volumetric scattering coefficient, m-1

τ Asymmetry factor

ψ Scattering angle, radian

Ω Solid angle, steradian

Abbreviations Definition

2,4-D 2,4-Dichlorophenoxyacetic Acid

2,4-DCP 2,4-Dichlorophenol

4-CP 4-Chlorophenol

AOPs Advanced Oxidation Processes

ARPs Advanced Reduction Processes

CCA Chromated Copper Arsenate

CMC Critical Micelle Concentration

CTAB Cetyltrimethylammonium Bromide

DDT Dichlorodiphenyltrichloroethane

ECD Electron Captured Detector

EQE External Quantum Efficiency

GC Gas Chromatography

HGMT Hollow Glass Microspheres Coated with Anatase TiO2

HPLC High Performance Liquid Chromatography

IARC International Agency for Research Cancer

ID Irradiated Distance

LED Light Emitting Diode

LMB Leuco-methylene Blue

LVREA Local Volumetric Rate of Energy Absorption

MB Methylene Blue

MC Monte Carlo

MCPA 2-methyl-4-chlorophenoxyacetic acid

Page 19: 1.3 Light emitting diode

xviii

PCB138 2,2',3,4,4',5'-Hexachlorobiphenyl

PCBs Polychlorinated biphenyls

PEPO Photon Energy per Order

PFP Pentafluorophenyl

PVC Polyvinyl Chloride

RTE Radiation Transport Equation

SDS Sodium Dodecyl Sulfate

SEM Scan Electron Microscopy

TEA Triethylamine

TOC Total Organic Carbon

TWEEN80 Polyoxyethylene (80) Sorbitanmonooleate

USEPA United States Environmental Protection Agency

UV Ultraviolet

UVA Ultraviolet, subtype A

UVB Ultraviolet, subtype B

UVC Ultraviolet, subtype C

UV-Vis Ultraviolet-visible

VUV Vacuum Ultraviolet

Page 20: 1.3 Light emitting diode

1

Chapter One: INTRODUCTION

1.1 Background

In the past several decades, increased population, industrialization and agricultural

activities have led to an increase in the level of water contamination of receiving water

bodies. Pesticides, a major category of pollutants causing water contamination, pose a

potential threat to human health and the environment. Using pesticides is almost a

necessary way to maintain and improve the food production for an ever increasing world

population. However, extensive use of pesticides has resulted in water pollution in

different ways such as runoffs, run-ins and leaching (Polyrakis, 2009). The primary focus

of this thesis is to study the photochemical treatment of pesticides in water and to design

an efficient light emitting diode (LED) based photocatalytic reactor. Polychlorinated

biphenyls (PCBs) form a secondary interest in this thesis.

Pesticides exposure can cause different acute and chronic effects on human health

(Younes and Galal-Gorchev, 2000). A large number of pesticides, such as mancozeb,

dithiocarbamate and organophosphorus compound, manifest their toxicity through

functional and biochemical action in the central and peripheral nervous system (Kimura

et al., 2005). Several chronic diseases have been linked to the long-term exposure to

pesticides. Examples include porphyria following exposure to hexachlorobenzene,

delayed neuropathy from exposure to organophosphates and chloracne due to long-term

exposure to chlorophenoxy acid derivatives and chrolophenols (Younes and Galal-

Gorchev, 2000). Besides, cancers of the soft tissue, lung, gonads, liver, brain, the urinary

Page 21: 1.3 Light emitting diode

2

tract and the digestive system have been associated with long-term exposure to some

pesticides, although the association is not firm (Younes and Galal-Gorchev, 2000).

Polychlorinated biphenyls (PCBs) are toxic contaminants, which are less soluble in

water, but can bind to sediments of aquatic systems or adsorb on suspended particulates

(Sullivan et al., 1983, Manchester-Neesvig et al., 1996, Bergen et al., 1998). Once they

are released to the environment, they are difficult to remediate. The occurrence of water

contamination by PCBs is due to desorption from sediments or leaching from landfills

and contaminated soil. PCBs have been demonstrated to cause a variety of adverse health

effects. Data on animal experiments have provided conclusive evidence that PCBs are

carcinogenic to animals and can cause a number of non carcinogenic health effects,

including effects on the immune system, nervous system, endocrine system, reproductive

system and others (USEPA, 2013a). The studies also support that PCBs can cause

potential carcinogenic and non-carcinogenic effects to human beings (USEPA, 2013a).

Long term exposure to PCBs can cause damages to heart, kidney, liver and central

nervous systems (Erickson, 1997).

To alleviate water pollution with these two categories of pollutants, a variety of

techniques has been developed: bio-treatment (Hussain et al., 2009, Portier et al., 1990,

Zhang et al., 2004, Natarajan et al., 1996), membrane separation (Bhattacharya, 2006,

Boussahel et al., 2000), activated carbon adsorption (Foo and Hameed, 2010, Sotelo et

al., 2002), coagulation followed by settling (Dempsey and O'Melia, 1984) and others.

Page 22: 1.3 Light emitting diode

3

Although these methods, to some extent, can reduce the water contamination, several

drawbacks limit them from wider applications. Bio-technologies may require specific

pollutant resistant microbes as well as appropriate environmental conditions such as pH,

nutrients and temperature. Physical separation can remove contaminants from water and

transfer them to other phases. Nevertheless, the disposal of the concentrate or sludge can

be a serious problem. Besides, regeneration of adsorbents and fouling of membranes limit

the application of these techniques. To detoxify and degrade these compounds,

technologies based on photochemical processes are considered as good choices

(Devipriya and Yesodharan, 2005, Ollis et al., 1991, Parsons, 2004, Izadifard et al.,

2010a, Achari et al., 2003, Chu et al., 2005). These methods are quite fast and mostly

lead to complete degradation of the contaminants.

1.2 Photochemical Treatment Processes

Most photochemical treatment processes are based on advanced oxidation processes

(AOPs), using the generated hydroxyl radicals, positive holes, oxygen species and other

strong oxidants to degrade the organic compounds. They have been widely used in

removing organic contaminants in water and wastewater such as disinfection by-

products, pesticides, endocrine disruptors and so on (Parsons, 2004). Besides, during

some photochemical processes, highly reactive reducing radicals, such as free electrons,

may be formed. These strong reducing agents can be used to degrade the oxidized

contaminants such as nitrate, perchlorate, dichlorophenols and perfluorooctanoic acid

(Vellanki et al., 2013). Such photochemical treatment techniques are called the advanced

reduction processes (ARPs).

Page 23: 1.3 Light emitting diode

4

In this research, the major interest is focused on the removal of aqueous contaminants

such as pesticides. The most commonly used pesticides such as 2,4-D, MCPA and

chlorophenols were selected as the studied compounds. TiO2 photocatalysis based on

AOPs was chosen for degrading these compounds, as it does not consume large amount

of chemicals and is able to use the longer wavelength domain of ultraviolet light, which is

a part of UV region in the solar spectrum received on the surface of earth. Application of

TiO2 photocatalysis can lead to usage of longer (less energy) light sources. To treat PCBs

in aqueous medium, photosensitization based on ARPs are used.

1.3 Light emitting diode (LED) in photocatalytic reactors

To apply TiO2 photocatalysis in water and wastewater treatment, an efficient

photocatalytic reactor need to be designed and fabricated. The rapid development of LED

technology has made it a promising light source in photochemical applications. This

mercury-free light source is able to provide monochromatic light, has a longer lifetime,

and efficient electricity to light conversion (Würtele et al., 2011). Furthermore, the small

size of LEDs does not limit the geometry of the reactor. All these advantages have made

LEDs favourable in photocatalytic reactor designs. The application of LEDs has been

reported in photochemical treatment of air and water by several researchers (Huang et al.,

2009, Shie et al., 2008, Chen et al., 2005, Wang and Ku, 2006, Ghosh et al., 2008, Ghosh

et al., 2009). In this research, ultraviolet-light emitting diodes (UV- LEDs) are selected

for reactor design and fabrication.

Page 24: 1.3 Light emitting diode

5

1.4 Research Objectives and Scopes

The goal of this research is twofold: (1) develop efficient photochemical technologies to

treat contaminants (e.g. pesticides) in aqueous medium and design an efficient LED

based photocatalytic reactor; (2) study photochemical treatment of PCBs in aqueous

medium. To achieve this goal, four objectives are defined:

Dechlorinate PCBs in aqueous surfactant solutions using photosensitized visible

light irradiation.

o Investigate the photodechlorination of PCBs using Leuco-methylene blue

as a photosensitizer and cool white lamps as a light source

o Determine the effect of the type and the concentration of surfactants on the

photodechlorination of PCBs.

o Optimize the PCBs dechlorination conditions.

Investigate the photocatalytic degradation of certain pesticides in a batch UV-

LED photoreactor.

o Design a batch UVA-LED based photoreactor.

o Investigate the photocatalytic degradation of pesticides mixtures.

o Study the effect of photocatalyst loading and light intensity on the

photocatalytic degradation rate.

o Compare the photocatalytic degradation of pesticides in the LED

photoreactor with the mercury lamps.

Develop a radiation field model for a UV-LED photocatalytic reactor and design a

homogenous radiation field.

o Determine the most efficient radiation field for a photocatalytic reactor.

Page 25: 1.3 Light emitting diode

6

o Develop and validate a mathematical radiation field model for LED

arrays.

o Develop a method for designing a homogenous radiation field generated

by LED arrays.

Design, fabricate and test an efficient UV-LED based photocatalytic reactor, as

well as optimize the reactor performance and provide useful information for the

scale-up of the reactor.

o Design a novel photocatalytic reactor using UV-LED and TiO2 nanotubes.

o Evaluate the performance of the photocatalytic reactor by studying the

degradation of pesticides.

o Optimize the photocatalytic reactor performance through the study of the

effect of different operational parameters on the photocatalytic

degradation rate of pesticides.

1.5 Thesis Overview

This thesis contains seven chapters as outlined here.

Chapter one provides the general background information, research scope and objectives,

and an outline of the dissertation.

Chapter two provides a review of principles of photochemistry, photocatalysis,

photochemical treatment of PCBs and pesticides, designs of photocatalytic reactors and

radiation field modelling.

Page 26: 1.3 Light emitting diode

7

Chapter three describes a study of dechlorination of PCBs in surfactant solution with

visible light irradiation using a photosensitizer-Leuco methylene blue (LMB). In this

chapter, the generation of LMB through two different ways are studied. The impact of

surfactant type and surfactant concentration on PCBs photodechlorination efficiency is

investigated.

Chapter four describes photocatalytic degradation of phenoxy herbicides and

chlorophenols with a UV-LED light source in a TiO2 slurry system. During this research,

a batch UV-LED photoreactor is fabricated. The impact of light intensity and TiO2

loading on photocatalytic degradation is investigated.

Chapter five describes the development of a radiation field model for a LED based

photocatalytic reactor and the design of a homogenous radiation field.

Chapter six describes the design, fabrication and optimization of a flow-through LED

based photocatalytic reactor. Parameters such as flow rate, light intensity, and

photocatalyst configuration are studied.

Chapter seven provides a summary of research results as well as recommendations for

future research.

This thesis is written in a paper format where chapter 3, 4, 5 and 6 comprise separate

papers. Chapter 3 and 4 have been published as "Yu, Linlong; Izadifard Maryam; Achari,

Page 27: 1.3 Light emitting diode

8

Gopal; Langford, Cooper H., 2013. Electron transfer sensitized photodechlorination of

surfactant solubilized PCB 138. Chemosphere, 90, 2347-2351." and "Yu, Linlong;

Achari, Gopal; Langford, Cooper H., 2013. LED-Based Photocatalytic Treatment of

Pesticides and Chlorophenols. Journal of Environmental Engineering, 139, 1146-1151.",

respectively. Chapter 5 has been submitted to Journal of Environmental Engineering and

Science.

Page 28: 1.3 Light emitting diode

9

Chapter Two: LITERATURE REVIEW

2.1 Principle of Photochemistry

2.1.1 Light and photon

Photochemistry is the science of light-induced chemical reactions. The modern theory of

quantum mechanics considers light beam as consisted a number of photons which possess

the property of both waves and particles (Turro, 1991). Each photon has energy related to

its wavelength (Plank's Equation, Equation [2-1]). The shorter the wavelength the higher

the energy it carries.

E = hν =hcλ

[2-1]

Where E is the radiant energy of the photon (J), h is Plank constant (6.62*10-34 J·s), ν is

the frequency of photon (s-1), λ is the wavelength of photon (m) and c is the velocity of

photon travelling in vacuum (m s-1).

The wavelength range generally utilized in photochemistry lies between 170 nm and 1000

nm (Figure 2-1) (Oppenlander, 2003), which is divided into five sub region: the vacuum-

UV or VUV (below 200 nm), UV-C (200-280 nm), UV-B (280-315 nm), UV-A (315-380

nm), VIS (380-850 nm) and infrared (800-1000nm) The subdivisions of the UV spectral

domain are related to physical, chemical, biological or biochemical effects showed in

Figure 2-2.

Page 29: 1.3 Light emitting diode

10

Figure 2-1: Classification of electromagnetic radiation in the wavelength range below 1200 nm. [Reproduced from (Oppenlander, 2003) with the permission].

Figure 2-2: Phenomenological subdivision of ultraviolet radiation into four sub-bands and their characteristic effects. [Reproduced from (Oppenlander, 2003) with the permission].

Photoionization

M→M++e

0 200 400 600 800 1000

Wavelength (nm)

IR

Visible light UVA

UVB

UVC

VUV X-ray

γ-ray

Vibrational excitation

M→Mvib

Photoexcitation

M→M*

400 100 150 200 250 300 350

Wavelength (nm)

UVA (315-380 nm)

UVB (280-315 nm)

UVC(200-280 nm)

VUV (100-200nm)

Absorbed by Organic Chromophores

Absorbed by all substances including H2O,O2,CO2

Absorbed by all Proteins, DNA, RNA,O2

Sunburn Skin Cancer

Sun Tanning

Page 30: 1.3 Light emitting diode

11

2.1.2 The electronic excited states

In photochemistry, only the absorbed photon can cause a photochemical reaction, and

each photon is absorbed by a single molecule to initiate the reaction (Turro, 1991).

Absorption in the wavelength region of photochemical interest promotes the absorber

from its ground state to its excited state. Absorption at longer wavelengths (infra-red)

usually leads to the excitation of vibrations or rotations of a molecule in its ground state;

generally, only electronically excited states are involved in photochemical processes

(Wayne, 1988). The fates of excited species 'A*' are shown as below (Wayne, 1988) :

The excited species 'A*' can lose its energy by emitting a photon, which gives rise

to the phenomenon of luminescence.

The excess energy of 'A*' can also be relieved by an atom or molecule 'M' in the

form of physical quenching. Normally, in this process, the excess energy of 'A*' is

converted to translational or vibrational excitation of 'M*' at lower energy.

The excited species A* can transfer energy to other molecules to generate other

excited species, which can then participate in any of the processes including

relaxation to the ground state (radiationless decay);

The excited species A* may undergo dissociation, direct reaction, ionization or

spontaneous isomerization.

2.1.3 Quantum yield

The absorption of photons can cause other processes rather than the desired reaction. To

determine the efficiency of the photochemical reaction, the concept of quantum yield was

Page 31: 1.3 Light emitting diode

12

developed. Four commonly used definitions of quantum yield are shown in Table 2-1.

The quantum yield is a unitless constant, usually ranging from zero to one; the value of

quantum yield larger than one indicate a photo-induced chain reaction involving radicals

species or photo-generated catalysis (Oppenlander, 2003).

Table 2-1: Definitions of the quantum yield (Oppenlander, 2003).

Mathematical

expression

Definition

𝜙𝜆 =dn(event)/dt

Φpabs

Universally valid: Number n of events per unit time divided by the

number of photons absorbed during this period

𝜙𝜆 =dn(M)/dt

Φpabs

Number n of reactant molecules M consumed per unit time divided

by the number of photons absorbed during this period

𝜙𝜆 =dn(P′)/dt

Φpabs

Number n of photoproduct molecules P' formed per unit time

divided by the number of photons absorbed during this period

𝜙λ1−λ2

=dn(P′)/dt

Φpabs(λ1−λ2) ≠ 𝜙𝜆

Ratio of the number m of photoproduct molecules formed per unit

time to the total number of photons absorbed in the spectral region

λ1- λ2 during this period

note: Φpabs and Φp

abs(λ1−λ2)are the absorption rates of photons.

2.1.4 Direct photolysis

Direct photolysis involves the transformation of a chemical resulting from the direct

absorption of a photon. Absorption of photons with high energy can promote the

contaminants (e.g.2-chloro-N-methylacetanilide ) to their excited singlet states from

Page 32: 1.3 Light emitting diode

13

electronic ground state. This excited state can then undergo, among other processes; (i)

homolysis (ii) heterolysis or (iii) photoionization (Burrows et al., 2002). Most organic

compound show absorption bands at relatively short UV wavelengths capable of

producing direct photolytic degradation of these compounds.

2.1.5 Photosensitized degradation

Photosensitization is the process of initiating a reaction through the use of a

photosensitizer capable of absorbing light and transferring the energy or exchanging

electrons with the reactants (Burrows et al., 2002). The major advantage of

photosensitized photodegradation is its possibility to use light of wavelengths longer than

those corresponding to the absorption characteristics of the pollutants.

2.1.6 Photocatalysis

Photocatalysis is a chemical reaction induced by absorption light by a photocatalyst

(Ohtani, 2008a). With solid photocatalyst, the reaction is activated by absorption of a

photon with sufficient energy, i.e. equal or higher than the band-gap energy of the

photocatalysts (Fox and Dulay, 1993, Herrmann, 2005, Hoffmann et al., 1995). The

band-gap energy is the energy difference between the bottom of conduction band (lowest

unoccupied molecular orbital) and the top of the valance band (highest occupied

molecular orbital) related to the electronic structures of semiconducting materials.

Various semiconductors such as TiO2, CdO, ZnO, WO3, CdS, CdSe, GaP, GaAs, ZnS,

SnO2, Fe2O3, SrTiO3, BaTiO3 etc, have been used as photocatalysts. Generally, the best

photocatalytic performances are obtained with titanium dioxide as catalyst (Herrmann,

Page 33: 1.3 Light emitting diode

14

2005). The details of TiO2 photocatalysis fundamental and mechanism will be described

in the section 2.2.

2.1.7 Advanced Oxidation Processes (AOPs)

AOPs are processes designed to degrade recalcitrant organic compounds using chemical

oxidants. Most organic contaminants can be completely mineralized or partially

mineralized to innocuous compounds using appropriate AOPs. Currently, the major light

induced AOPs include UV&H2O2, Ozone&UV, vacuum UV, TiO2 photocatalysis, and

others. Besides, there are several non-light induced AOPs such as H2O2, Fenton’s reagent

and ozonation. Although different AOPs make use of different reaction systems (Table 2-

2), the chemistry of these reaction systems are similar: generation of highly reactive

oxidative species, such as hydroxyl radicals (OH•), positive holes and singlet oxygen

(Andreozzi et al., 1999). The oxidation potential of hydroxyl radicals are greater than that

of most conventional oxidants such as chlorine, oxygen, ozone, etc. (Parsons, 2004).

2.2 TiO2 Photocatalysis

2.2.1 TiO2 as a photocatalyst

TiO2 is considered to be the most successful photocatalyst as it has several advantages

such as: (1) photo active (2) low toxicity (3) biologically and chemically stable (4) able to

utilize near UV light and (5) economic (Bhatkhande et al., 2002, Linsebigler et al., 1995,

Hoffmann et al., 1995). Titanium dioxide naturally exists in three crystal forms: anatase,

rutile and brookite. Brookite is extremely difficult to synthesize, while anatase and rutile

Page 34: 1.3 Light emitting diode

15

Table 2-2: Generation of hydroxyl radicals for different AOPs.

Type of AOPs Spectral domain Reactions

Vacuum UV VUV 𝐻2𝑂ℎ𝑣��𝐻𝑂 ∙ +𝐻 ∙

UV/H2O2 UVC 𝐻2𝑂2ℎ𝑣��𝐻𝑂 ∙ +𝐻𝑂 ∙

TiO2/UV UVA-UVC

𝑇𝑖𝑂2ℎ𝑣�� 𝑒− + ℎ+

ℎ+ + 𝐻2𝑂 → 𝐻𝑂 ∙ +𝐻+

ℎ+ + 𝑂𝐻− → 𝐻𝑂 ∙

O3/UV UVC

𝑂3ℎ𝑣��𝑂(𝐷) + 𝑂2

𝑂(𝐷) + 𝐻2𝑂 → 𝐻2𝑂2

𝐻2𝑂2ℎ𝑣��𝐻𝑂 ∙ +𝐻𝑂 ∙

Ozonation No UV

𝐻𝑂− + 𝑂3 → 𝑂2 + 𝐻𝑂2−

𝐻𝑂2− + 𝑂3 ⇄ 𝐻𝑂2 ∙ +𝑂3 ∙−

𝐻𝑂2 ∙⇄ 𝐻+ + 𝑂2 ∙−

𝑂2 ∙−+ 𝑂3 → 𝑂2 + 𝑂3 ∙−

𝑂3 ∙−+ 𝐻+ → 𝐻𝑂3 ∙

𝐻𝑂3 ∙→ 𝐻𝑂 ∙ +𝑂2

𝐻𝑂 ∙ +𝑂3 → 𝐻𝑂2 ∙ +𝑂2

Fenton process No UV

𝐹𝑒2+ + 𝐻2𝑂2 ⇄ 𝐹𝑒𝑂2+ + 𝐻2𝑂

𝐹𝑒2+ + 𝐻2𝑂2 → 𝐹𝑒3+ + 𝑂𝐻− + 𝑂𝐻 ∙

𝐹𝑒3+ + 𝐻2𝑂2 ⇄ 𝐹𝑒𝑂𝑂𝐻2+ + 𝐻+

𝐹𝑒𝑂𝑂𝐻2+ ⟶ 𝐻𝑂2 ∙ +𝐹𝑒2+

Page 35: 1.3 Light emitting diode

16

can be produced easily in the laboratory (Bickley et al., 1991). Among these three crystal

forms, anatase and rutile are the two most commonly used types and have been employed

most in the photocatalytic study. The band-gap energy are, respectively, 3.0 eV, 3.2 eV,

for rutile phase and anatase phase, and its amorphous form is reported to have the band-

gap energy varying from 3.2 to 3.5 eV (Roy et al., 2011). Even though the most active

form of titanium dioxide is believed to be anatase, a mixed phase of anatase and rutile

appears to achieve better photocatalytic efficiency (Bickley et al., 1991). The co-presence

of anatase and rutile phase introduce mesoporosity and a wider pore size distribution,

which may be responsible for the high level of photocatalytic activity (Thiruvenkatachari

et al., 2008). Hurum et al. (2003) proposed three possible reasons for the greater

photocatalytic activity of TiO2 mixed phase: (1) the band-gap of rutile is smaller than that

of anatase and extends the useful wavelength range of photoactivity; (2) the transfer of

photoexcited electrons between rutile/anatase phase enhance the charge separation and

slows down electron-hole recombination; (3) the small size of the rutile crystallites

enhance the photocatalyst activity.

Degussa (Evonik) P25, Aeroxide TiO2 P25, via the chloride technology method is

currently the de-facto commercial reference TiO2 photocatalyst (Alonso-Tellez et al.,

2012). It is widely used in potocatalytic reaction systems because of its high

photocatalytic activity, and has been reported in more than one thousand papers since

1900 (Ohtani et al., 2010). P25 has a large surface area (50 m2 g-1) (Zertal et al., 2004)

and small crystal size (20 nm). Theoretically, a photocatalyst with larger surface area and

smaller particle size can provide more active sites for illumination and adsorption of the

Page 36: 1.3 Light emitting diode

17

reactants, leading to a higher expected photocatalytic activity. The composition of P25 is

reported to be 70% anatase and 30% rutile or 80% anatase and 20% of rutile, however,

the exact crystalline composition seems to be unknown, presumably due to a lack of

determination techniques for crystalline contents in nano-sized particle samples (Ohtani

et al., 2010, Ohtani, 2008b). Except Degussa P25, other commercial TiO2, such as

products from Millennium and Hombikat also show their high photocatalytic activities

(Zertal et al., 2004, Alonso-Tellez et al., 2012) .

2.2.2 Mechanism of TiO2 photocatalysis

The fundamentals and mechanism of TiO2 photocatalysis have been intensively reported

in many literatures (Fujishima et al., 2000, Gaya and Abdullah, 2008, Fox and Dulay,

1993, Herrmann, 1999). The overall process of TiO2 photocatalysis can be broken into

five independent steps (Herrmann, 2005, Mozia, 2010) :

Transfer of the reactants in the bulk solution to the TiO2 surface;

Adsorption of the reactants on the surface of TiO2;

Reaction in the adsorbed phase;

Desorption of the products;

Removal of by-products from the interface region.

The third step includes all the photochemical processes (Herrmann, 2005) and is

summarized in Equations [2-2]~[2-14] (Mozia, 2010) and Figure 2-3. The initial step of

photon-induced reaction is the excitation of TiO2 by absorbing photon with formation of

electron-hole pair. Once TiO2 absorbs photons with sufficient energy, i.e. equal or larger

than its band-gap energy, electrons are promoted from the valence band to the conduction

Page 37: 1.3 Light emitting diode

18

band, while positive holes (h+) are left in the valence band (Equation [2-2]). The electron

and the hole can migrate to the catalyst surface and participate in the redox reactions

(Equations [2-4] ~ [2-14]) with different species adsorbed on the catalyst surface. A

recombination of the electron and hole will occur if no suitable electron and hole

scavengers are present (Equation [2-3]). If oxygen is present in the water (e.g. water

open to air), it will capture the electron in the conduction band to form the superoxide

radical ion while the remaining hole can react with surface-bond H2O molecule or

hydroxide ion to produce hydroxyl radicals. Hydroxyl radicals can also be generated

following the pathways through reactions shown in Equations [2-7] ~ [2-11]. The

hydroxyl radicals generated on the surface of illuminated TiO2 are supposed to be the

primary oxidizing species in the photocatalytic oxidation processes, which are highly

reactive and can degrade most organic compound and eventually convert them to CO2,

H2O and other inorganic compounds.

𝑇𝑖𝑂2ℎ𝑣�� 𝑇𝑖𝑂2 (𝑒− + ℎ+) [2-2]

𝑒− + ℎ+ → heat [2-3]

ℎ+ + 𝐻2𝑂 → 𝑂𝐻 ∙ +𝐻+ [2-4]

ℎ+ + 𝑂𝐻− → 𝑂𝐻 ∙ [2-5]

𝑒− + 𝑂2 → 𝑂2− ∙ [2-6]

𝑂2−. +𝐻+ → 𝐻𝑂2 ∙ [2-7]

𝐻𝑂2 ∙ +𝐻𝑂2 ∙→ 𝐻2𝑂2 + 𝑂2 [2-8]

𝑒− + 𝐻2𝑂2 → 𝑂𝐻 ∙ +𝑂𝐻− [2-9]

Page 38: 1.3 Light emitting diode

19

𝑂2− ∙ +𝐻2𝑂2 → 𝑂𝐻 ∙ +𝑂𝐻− + 𝑂2 [2-10]

𝐻2𝑂2ℎ𝑣�� 2𝑂𝐻 ∙ [2-11]

𝑂𝑟𝑔𝑎𝑛𝑖𝑐 𝑐𝑜𝑚𝑝𝑜𝑢𝑛𝑑 + 𝑂𝐻 ∙→ 𝑑𝑒𝑔𝑟𝑎𝑑𝑎𝑡𝑖𝑜𝑛 𝑝𝑟𝑜𝑑𝑢𝑐𝑡𝑠 [2-12]

𝑂𝑟𝑔𝑎𝑛𝑖𝑐 𝑐𝑜𝑚𝑝𝑜𝑢𝑛𝑑 + ℎ+ → 𝑜𝑥𝑖𝑑𝑎𝑡𝑖𝑜𝑛 𝑝𝑟𝑜𝑑𝑢𝑐𝑡𝑠 [2-13]

𝑂𝑟𝑔𝑎𝑛𝑖𝑐 𝑐𝑜𝑚𝑝𝑜𝑢𝑛𝑑 + 𝑒− → 𝑟𝑒𝑑𝑢𝑐𝑡𝑖𝑜 𝑝𝑟𝑜𝑑𝑢𝑐𝑡𝑠 [2-14]

Figure 2-3: Photochemical activation of TiO2.

Page 39: 1.3 Light emitting diode

20

2.2.3 The kinetics of photocatalytic degradation

The Langmuir-Hinshelwood (LH) model proves to be the best for simulation of the

kinetic rate of initial photocatalytic degradation (Kumar et al., 2008, Matthews, 1988,

Mills and Morris, 1993). In the LH model, the rate of photocatalytic reaction is controlled

by the reaction of the adsorbed molecules. Firstly, the substrate adsorbs on the surface of

the photocatalyst and then undergoes photocatalytic degradation. This model is based on

several assumptions (Fox and Dulay, 1993):

Adsorption by the substrate is identical for each site and is independent of surface

coverage;

At equilibrium, the number of surface adsorption sites is fixed;

Each surface site is only combined with one substrate molecule;

The adjacent adsorbed molecules do not react with each other;

The rate of adsorption is greater than other chemical reactions;

No irreversible blocking of active sites by binding to product occurs.

Ollis (2005) has shown that the model can fit the data well even though the adsorption

process is not at equilibrium as is shown by the dependence of the adsorption coefficient

dependence on light intensity.

The Langmuir-Hinsheldwood Kinetic Expression for the photocatalytic degradation were

shown in Equation [2-15] (Fox and Dulay, 1993).

rp =𝑑𝑐𝑑𝑡

=𝐾𝑎𝑑𝑘𝐶

1 + 𝐾𝑎𝑑𝐶 [2-15]

Page 40: 1.3 Light emitting diode

21

Where rp is the reaction rate (mol L-1 s-1), C is the concentration of substrate (mole L-1),

Kad is adsorption coefficient for substrate (L mol-1), k is the apparent kinetics rate

constant occurring at the active site on the photocatalyst surface (mol L-1 s-1)

When the initial concentration contaminants is very high, 𝐾𝑎𝑑𝐶 ≫ 1, consequently,

Equation [2-15] can be simplified as zero order reaction kinetics:

𝑑𝑐𝑑𝑡

= k [2-16]

At highly diluted concentration, 𝐾𝑎𝑑𝐶 ≪ 1, the photocatalytic degradation become first

order reaction (Equation [2-17]) (Mozia, 2010, Herrmann, 2005).

𝑑𝑐𝑑𝑡

= 𝐾𝑎𝑑𝑘𝐶 [2-17]

2.2.4 Factors affecting the photocatalytic degradation kinetics

2.2.4.1 TiO2 loading

The effect of TiO2 loading on photocatalytic degradation of different contaminants in

aqueous solution has been widely investigated (Chen and Liu, 2007, Singh et al., 2007,

Kaneco et al., 2009, Liu et al., 2009, Wu et al., 2010, Muneer et al., 2005, Qamar et al.,

2006, Pizarro et al., 2005, Garcia and Takashima, 2003). Generally, at a low

photocatalyst loading range, the photocatalytic reaction rates were observed to be

proportional to the catalyst loading. As the photocatalyst loading reaches an optimal

Page 41: 1.3 Light emitting diode

22

value, the photocatalytic reaction rate becomes independent of photocatalyst amount and

becomes constant. A further increase of photocatalyst loading beyond the optimum can

even inhibit photocatalytic reaction. This phenomena is associated with the effect of

photocatalyst loading on active surface area of photocatalyst and the lack of light

penetration into solution (Ahmed et al., 2011, Mozia, 2010, Chen and Liu, 2007,

Adesina, 2004). An increase of TiO2 loading can enlarge the active surface area available

for reactant adsorption and photon absorption, hence a higher photocatalytic degradation

rate was expected. On the other hand, high loading of photocatalyst can cause light

scattering and screening effects, which impede the penetration of light into the solution

far from radiation source (Chen and Liu, 2007, Lea and Adesina, 1998, Singh et al., 2007,

Rahman and Muneer, 2005). Moreover, the agglomeration of photocatalyst at high solid

loading can result in a loss of active surface area (Chen and Liu, 2007, Lea and Adesina,

1998). The trade-off between these two opposite effects leads to an optimal photocatalyst

loading for the photocatalytic reaction.

2.2.4.2 Light intensity

Light intensity is another key parameter in the TiO2 photocatalysis. A power law

relationship (Equation [2-18] ) between the photocatalytic reaction rate (k) and light

intensity (I) was observed in a number of experimental studies (Wang et al., 2012, Choi

et al., 2000, Kim and Hong, 2002, Obee and Brown, 1995, Ohko et al., 1997). The

exponent (α) varies from one to zero as the light intensity increases.

Page 42: 1.3 Light emitting diode

23

𝐾 ∝ 𝐼𝛼 [2-18]

The effect of light intensity on the kinetics of the photocatalytic degradation process due

to the competition between electron-hole generation and electron-hole recombination

were summarized by Ollis et al.(1991) as follows:

At low light intensity range, electron-hole formation dominates and the apparent

photocatatytic reaction rate is proportional to the light intensity;

At intermediated light intensities, electron-hole pair generation competes with the

recombination of electron-hole and the photocatalytic reaction rate is linear to the

square root of light intensity;

At high light intensities, the photocatalyst loading become a limiting factor,

consequently, the increased light intensity does not improve the photocatalyic

reaction rate.

2.2.4.3 pH

The effect of pH on photocatalytic process is complicated (Fox and Dulay, 1993,

Konstantinou and Albanis, 2004, Mozia, 2010, Akpan and Hameed, 2009). Firstly, the

surface charge of TiO2 and ionization state of some contaminants is strongly influenced

by pH, which thus impacts the adsorption behaviour of contaminants. The surface charge

of TiO2 at different pH is determined by the following reactions.

𝑇𝑖𝑂𝐻 + 𝐻+ ⇔ 𝑇𝑖𝑂𝐻2+ [2-19]

Page 43: 1.3 Light emitting diode

24

𝑇𝑖𝑂𝐻 + 𝑂𝐻− ⇔ 𝑇𝑖𝑂− + 𝐻2𝑂 [2-20]

The isoelectric point of commercial available TiO2 (Degussa P25) is observed at pH=6.8

(Pelton et al., 2006, Mozia, 2010). When pH is lower than 6.8, the surface of P25 is

positively charged and the adsorption of negatively charged contaminants is favoured,

while at pH>6.8, the negatively charged P25 more likely attract the positively charged

contaminants. Secondly, at low pH condition, the TiO2 particles tend to agglomerate, as

a result, the available surface area for reactants adsorption and photon illumination is

reduced, which finally limit the photocatalytic reaction. Thirdly, the reaction between

hydroxide ions and positive holes can generate hydroxyl radicals. At low pH, the positive

holes are expected to be the major oxidation species whereas at high pH levels, the

predominant oxidation species are considered to be hydroxyl radicals. In alkaline

solution, the generation of hydroxyl radicals are easier through oxidizing more hydroxide

ions available on TiO2 surface, thus the efficiency of the process is logically enhanced

(Gonçalves et al., 1999, Shourong et al., 1997). Optimal pH for photocatalytic studies at

both low pH or at high pH have been observed. Higher photocatalytic degradation

efficiency for chlorophenols (Augugliaro et al., 1988), glyphosate (Muneer and Boxall,

2008) were observed at higher pH. However, some other contaminants like 2,4-D (Trillas

et al., 1995) and anionic dyes (Sakthivel et al., 2003) favor an acidic condition.

2.2.4.4 Electron acceptor

In the application of TiO2 photocatalysis, electron-hole recombination is the major step

of energy waste. Without suitable electron acceptors, the recombination step is

Page 44: 1.3 Light emitting diode

25

predominant and limits the photocatalytic reaction. The presence of electron acceptors in

solution can accelerate the photocatalytic degradation rate by (1) preventing the electron-

hole recombination (2) increasing the concentration of reactive oxygen species and

oxidation rate of intermediate compound. (Muruganandham and Swaminathan, 2006,

Singh et al., 2007, Ahmed et al., 2011). Usually, the dissolved oxygen in solution is used

as electron acceptor and promotes the photocatalytic reaction. Besides oxygen, other

oxidants such as H2O2, K2S2O8 KBrO3 can also act as electron acceptors. The effects of

these electron acceptors on photocatalytic degradation of pesticides has been investigated

by several researchers (Chen and Liu, 2007, Bahnemann et al., 2007, Singh et al., 2003,

Singh and Muneer, 2004, Rahman et al., 2006, Wei et al., 2009). All results indicate a

higher photocatalytic degradation rate when additional electron acceptors were

introduced.

Chen and Liu (2007) reported that adding a small amount of H2O2 (up to 0.1 mM) can

improve the efficiency of photocatalytic degradation of glyphosate. However, at high

concentration of H2O2 (larger than 0.1mM) the photocatalytic degradation of glyphosate

is retarded. Similar effects were also found in photocatalytic degradation of other

contaminants such as azo dyes (So et al., 2002), dicamba (Chu and Wong, 2004)

monochlorbenzene (Tseng et al., 2012) and triclosan (Yu et al., 2006). Hydrogen

peroxide is considered to be a better electron acceptor than oxygen (Equation [2-21]).

Moreover, under UV irradiation, hydrogen peroxide can also undergo direct photolysis

and generate hydroxyl radicals ( Equation [2-22]) (Ahmed et al., 2011).

Page 45: 1.3 Light emitting diode

26

𝐻2𝑂2 + 𝑒𝐶𝐵− → 𝑂𝐻 ∙ +𝑂𝐻− [2-21]

𝐻2𝑂2 + ℎ𝑣 → 2 𝑂𝐻 ∙ [2-22]

However, excess hydrogen peroxide can scavenge the generated hydroxyl radicals

(Equations [2-23]~[2-24]) and retard the photocatalytic degradation. Besides, the high

concentration of hydrogen peroxide can absorb and attenuate the UV light available for

TiO2 excitation (Muruganandham and Swaminathan, 2006, Chu and Wong, 2004).

𝐻2𝑂2 + 𝑂𝐻 ∙→ H2O + HO2 ∙ [2-23]

HO2 ∙ +𝑂𝐻 ∙→ H2O + O2 [2-24]

2.2.4.5 Hole/hydroxyl radicals scavenger

The photocatalytic degradation of pesticides occurs through reactions with the generated

holes or surface hydroxyl radicals. Some inorganic anions present in solution such as Cl-,

NO3- SO4

2- , CO32- and HCO3

- can act as hydroxyl radicals scavenger and inhibit the

photocatalytic oxidation (Konstantinou and Albanis, 2004, Wu et al., 2009, Chen et al.,

1997). Although the hydroxyl radical scavengers can react with hydroxyl radicals/holes

to form corresponding radicals, the reactivity of these radicals is lower than that of

hydroxyl radical or holes. Therefore, a decrease of photocatalytic degradation efficiency

in the presence of inorganic ions is usually observed. Wu et al. (2009) reported that the

presence of 0.05 mM Cl- and NO3- significantly inhibit photocatalytic degradation of

terbufos. However, the same phenomenon was not observed in some photocatalytic

studies. D'Oliveira et al. (1993) found that the presence of 0.1M inorganic anions (Cl-,

Page 46: 1.3 Light emitting diode

27

SO42-, and NO3

-) did not change the initial rate of photocatalytic degradation of 3-

chlorophenol. Mehrvar et al. (2001) studied the effect of hydroxyl radical scavengers on

photocatalytic degradation of 1,4-dioxane and tetrahydrofuran and found that bicarbonate

and carbonate ions slowed down the 1,4-dioxane degradation rate but did not

significantly affect the tetrahydrofuran degradation rate.

2.3 Contaminants (Pesticides and PCBs)

2.3.1 Pesticides

Pesticides are defined as any substance or mixture of substances intended to prevent,

destroy, repel, mitigate or attack any pest such as insects, weeds, microorganisms, fungi,

and others (USEPA, 2011). Pesticides are broadly classified into two groups: chemical

pesticides and bio-pesticides. Most conventional pesticides in large scale use are

chemically based. Pesticides are further classified as herbicides, fungicides, insecticides,

molluscicides, nematicides, plant growth regulators, pheromones, acaricides, repellents

and rodenticides (Tadeo, 2008). The active portion of a chemical pesticide is known as

the active ingredient (Kamrin, 2000).

Pesticides are very important in increasing food production and controlling weeds

(Polyrakis, 2009). Since 1950, pesticide usage has grown 50-fold to about 2.5 million

tons per year (Tadeo, 2008). The pesticides sold in Canada add up to more than 40

million kilograms, which represents approximately 3% of pesticide sale in the world. In

Canada, pesticides are regulated by the Pest Management Regulatory Agency (PMRA)

Page 47: 1.3 Light emitting diode

28

under the pest control products act. Currently, more than 7000 pesticide products are

registered for use in Canada (Brimble et al., 2005). Table 2-3 lists the top 20 pesticides

used in Canada.

Table 2-3: Sales/use of the top 20 pesticide active ingredient in Canada (Brimble et al., 2005).

RANK Active ingredient Type Amount (×106 kg)

1 Glyphosate H 4.608

2 Creosote AM 2.163

3 MCPA H 1.540

4 2,4-D H 1.490

5 CCA AM 0.824

6 Triallate H 0.707

7 Mancozeb F 0.655

8 Ethalfluralin H 0.598

9 Atrazine H 0.553

10 Brommoxynil H 0.544

11 Surfactant bend A 0.505

12 Mineral oil A/G/H/I 0.394

13 Petroleum Hydrocarbon Blend A 0.375

14 Trifluralin H 0.356

15 S-metolachlor H 0.293

16 Chlorothalonil F 0.265

17 Metolachlor H 0.261

18 Chlorpyrifos I 0.252

19 Mecoprop H 0.252

20 1,3-dichloroproprene I 0.248

A: adjuvant AM: anti-microbial F: fungicide G: growth regulator H: herbicide I:

insecticide

Page 48: 1.3 Light emitting diode

29

There is evidence that extensive usage of pesticides has an effect on water quality and is

associated with various environmental and human health problems. Persistent pesticides,

such as DDT and lindane, showing highly toxic effects and cause severe damage to

ecological system are forbidden or strictly controlled; consequently, the impact of these

pesticides become less and less. Currently, most of pesticides frequently used in Canada

are non-persistent and are considered less harmful to the environment and humans.

Nonetheless, from a long term perspective they may cause chronic effects which are

difficult to characterize with current technology. The pesticides studied in this thesis are

two commonly used phenoxy herbicides and chlorophenols.

2.3.1.1 2,4-dichlorophenoxyacetic acid (2,4-D)

2,4-D (Figure 2-4) is a systemic phenoxy herbicide, capable of controlling many types of

broadleaf weeds, e.g. dandelion (Humburg, 1989). It has been widely applied in forest

management, cultivated agriculture, pasture rangeland, and lawns and to control aquatic

vegetation. The name brands for 2,4-D related herbicides include Aqua-Kleen, Barrage,

Malerbane, Planotox, Lawn-keep, Salvo, Weedone , among others (Kamrin, 2000).

Figure 2-4: Molecular structure of 2,4-dichlorophenoxyacetic acid.

Page 49: 1.3 Light emitting diode

30

2,4-D is considered to be potentially harmful to both animals and humans and can cause

toxic effect on aquatic wildlife and aquatic ecosystem. The toxicity of 2,4-D in animals

has been studied extensively (USEPA, 2005): the acute oral LD50 varies from 638 mg/kg-

1250 mg/kg in rat; the acute dermal LD50 is higher than 2000 mg/kg in rabbits; the acute

inhalation LD50 is higher than 1.179mg/L in rats. Experiments on rats show that high

doses of 2,4-D may result in fetuses with abdominal cavity bleeding and increased

mortality (Kamrin, 2000). Long term exposure to 2,4-D may result in an increase of the

probability of malignant tumours (Kamrin, 2000). For humans, a high level of 2,4-D can

result in coughing, dizziness, burning and temporary loss of muscle coordination (Laws

and Hayes, 1991). Hardell (1981) suggested that 2,4-D is associated with Hodgkin’s

disease, non-Hodgkin’s lymphoma, and soft tissue sarcoma. Nevertheless, no evidence

from epidemiologic studies show that the exposure to 2.4-D can cause cancers (Garabrant

and Philbert, 2002). It is classified as Group D chemical (USEPA, 2005), one that is not

classifiable as to human carcinogenicity.

2.3.1.2 2-methyl-4-chlorophenoxyacetic acid (MCPA)

MCPA (Figure 2-5 ) is also a systemic post-emergence phenoxy herbicide used to control

a wide spectrum of broadleaf weeds (Kamrin, 2000). In Canada, it is registered for use on

agricultural sites, on fine turf (parks, golf courses, zoos, botanical gardens, athletic

playing fields and play ground) and lawns (residences public and commercial buildings)

and sod (grown in sod farms harvested for transplanting), in forestry (spruce seedlings for

reforestation) and at industrial sites (vegetation control) (Health Canada, 2010 ). There

Page 50: 1.3 Light emitting diode

31

are four MCPA related active ingredients: MCPA acid, MCPA dimethylamine salt,

MCPA sodium salt, MCPA and MCPA 2-ethylhexyl ester. Trade names for MCPA or

MCPA related product include Agroxone, Class MCPA, Agritox, Agroxone, Agronzone,

Class MCPA, Dakatota, Envoy, Gordon's Amine and among others (Kamrin, 2000).

Figure 2-5: Molecular structure of 2-methyl-4-chlorophenoxyacetic acid.

The toxicological experiments on rats and rabbits show that MCPA is a slightly toxic

compound (Kennepohl et al., 2010, Kamrin, 2000, USEPA, 2004). Symptoms in humans

due to very high acute exposure include twitching, drooling, low blood pressure, slurred

speech, jerking and spasms and unconsciousness (Kennepohl et al., 2010). Long-term

exposure to MCPA can result in reduced feeding rates and retarded growth rates in rates

(World Health Organization, 2004). MCPA has a moderate to low toxicity to birds, with

reported LC50 value of 377 mg/kg in bobwhite quail; and it is slightly toxic to freshwater

fish, with reported LC50 values ranging from 117 to 232 mg/L in rainbow trout (Kamrin,

2000, World Health Organization, 2004). All of the available evidence indicates that

MCPA does not cause cancer (Kennepohl et al., 2010).

Page 51: 1.3 Light emitting diode

32

2.3.1.3 Chlorophenols

Chlorinated phenols are a group of compounds consisting of phenol with substituted

chlorines (Figure 2-6). There are 19 chlorophenol congeners including three

monochlorophenols, six dichlorophenols, six trichlorophenols, three tetrachlorophenols

and one pentachlorophenol (Exon, 1984). Most of purified chlorinated phenols are

colorless crystalline solids; with an exception that 2-chlorophenol is a clear liquid.

(Health Canada, 2008). They have an unpleasant odor, which is medicinal, pungent,

phenolic, strong, or persistent (USEPA, 1980) .

Figure 2-6: Molecular structure of chlorophenols.

Chlorophenols can be formed by direct chlorination or the hydrolysis of the higher

chlorinated derivatives of benzene (USEPA, 1980). They can also be formed through

chlorination of water containing natural phenol or phenolic wastes. They have been

widely used in the production of dyes, pigments, phenolic resins, pesticides (USEPA,

1980). Certain chlorophenols are also used directly as pesticides such as fungicides, flea

repellents, wood preservatives, and so on. In Canada, chlorophenols are no longer in

production. However, they are continued to be imported. There are 110 chlorophenols

Page 52: 1.3 Light emitting diode

33

related pesticide products registered for used in Canada under the Pest Control Products

Act (Health Canada, 2008) .

The toxic effects of chlorophenols are related to the degree of chlorination. Generally

chlorophenols with higher degree of chlorination are more toxic. Acute exposure to lesser

chlorinated phenols in humans results in muscular twitching, tremors, spasms, ataxia,

weakness, convulsions and collapse (Health Canada, 2008). Acute poisoning by

pentachlorophenol can cause general weakness, anorexia, sweating, nausea, fatigue,

ataxia, headache, hyperpyrexia, vomiting, tachycardia, abdominal pain, terminal spasms

and death (Health Canada, 2008). Soft-tissue sarcomas, Hodgkin's disease and leukaemia

have been reported in epidemiological studies of occupational groups exposed to

chlorinated phenols and phenoxy acids. IARC (1987) identified chlorophenols as possible

humans carcinogens (Group 2B compound).

2.3.2 PCBs

PCBs are a class of nonpolar components which consist of 1 to 10 chlorine atoms on a

biphenyl ring (Figure 2-7). There are 209 different PCB configurations, commonly

referred as congeners, based on the number of chlorines and their positions on the

biphenyl ring (Erickson, 1997). PCBs were commercially produced as complex mixtures

containing multiple congeners, which were manufactured and sold under many different

names. In North America, Aroclor is the best known trade name for PCB mixtures. The

brand Aroclor is always followed with four digit suffix number: the first two of that

generally refers to the number of carbon atoms in the biphenyl ring and the last two of

Page 53: 1.3 Light emitting diode

34

that indicates the degree of chlorination (USEPA, 2012). The molecular structures of

PCBs lead to high chemical stability, low dielectric constants, high thermal conductivity

and low flammability. These properties led PCBs to be widely used in the manufacture of

hydraulic fluids, plasticizers, carbonless copy paper, fluorescent lamp ballasts, flame

retardants, ink, adhesives, and other consumer products (USEPA, 2000).

Figure 2-7: Molecular structure of polychlorinated biphenyls.

PCBs are not readily soluble in water, but soluble in organic solvents, oils and fats. PCBs

are highly stable under most environmental conditions, and can be bioaccumulated in

plants, fish and other living tissues (Erickson, 1997). A number of toxicological studies

have identified PCBs as toxic compounds to animals, humans and ecosystems (Erickson,

1997, USEPA, 2013a). Like any other toxic substance, the toxicological effect depends

on the exposure dosage, exposure duration and routes. The structure of PCBs also

determines its toxicological effect. Normally, those PCB structures that contain no ortho-

chlorine substituent or only a single ortho-chlorine substitute are more toxic. Studies in

Page 54: 1.3 Light emitting diode

35

animals provide conclusive evidence that PCBs can cause cancer in animals and a

number of non carcinogenic health effects, including effects on the immune system,

nervous system, endocrine system and reproductive system (USEPA, 2013a).

Furthermore, PCBs can cause potential carcinogenic and non-carcinogenic effects of

PCBs to human beings (USEPA, 2013a). Therefore, USEPA and IARC have classified

PCBs as probable human carcinogens (Group B2). The human non-carcinogenic health

effects associated with the exposure to PCBs include chloracne and rashes on the skin,

liver damage, dermal and ocular lesions, irregular menstrual cycles and lowered immune

responses, fatigues, headaches, coughs, and unusual skin sores, and among others

(Erickson, 1997).

Concern about the adverse effects of PCBs has caused the production of PCBs to be

banned in 1979 in US (USEPA, 2013b). The global production of PCBs has been banned

by the Stockholm Convention on Persistent Organic Pollutants in May 2004 (Fiedler,

2007). Today, PCBs previously introduced into the environment have become the major

source of PCB related problems. PCBs do not readily break down in the environment and

can remain for long periods of time cycling between water, soil and air (USEPA, 2013b).

They are released to water from contaminated soil, sediments and landfill. PCBs in the

water and soils can move into atmosphere through volatilization, and return back to the

soil and surface waters through wet and dry deposition (USEPA, 2000).

Page 55: 1.3 Light emitting diode

36

2.4 Photochemical treatment of pesticides and PCBs

2.4.1 Direct photolytic degradation of pesticides and PCBs

Most pesticides show UV-Vis absorption bands at relatively short UV wavelengths,

therefore, a short UV wavelength light source is required in direct photolysis of

pesticides. Direct photolysis using UVC (254nm) are reported to treat different pesticides

(Gal et al., 1992, Herweh and Hoyle, 1980, Zepp and Cline, 1977). As well, sunlight or

simulated sunlight for direct photodegradation of pesticides were also investigated

(Samanidou et al., 1988, Wilson and Mabury, 2000, Okamura et al., 1999, Miille and

Crosby, 1983, Ellis and Mabury, 2000, Ying and Williams, 1999). Since sunlight

reaching the earth's surface contain a very small fraction of short wavelength UV

radiation, the direct photolysis of pesticides under sunlight irradiation is not efficient.

Direct photolysis of PCBs using short wavelength UV (254nm) has been reported to take

place in different organic solvents (Yao et al., 1997, Miao et al., 1999, Hawari et al.,

1992, Dhol, 2005). Direct photolysis of PCBs in alkaline isopropanol media were

observed as the most effective. Hawari et al. (1992) reported a high quantum yield (~30)

for direct photolysis of Aroclor1254 in alkaline isopropanol. Direct PCB photolysis

involving the use of sunlight irradiation has not shown to be effective, since PCBs do not

absorb light with wavelength above 300 nm.

Page 56: 1.3 Light emitting diode

37

2.4.2 Photosensitized degradation of pesticides and PCBs

Photosensitized degradation of pesticides have been successfully achieved using the

sensitizers like anthraquinone, N,N,N',N'-tetrarnethylbenzidine and humate (Galadi and

Julliard, 1996, Stangroom et al., 1998, Galadi et al., 1995, Crank and Mursyidi, 1992).

Photosensitization of PCBs using phenothiazine and hydroquinoes has also been studied

(Hawari et al., 1992, Chu and Kwan, 2002). A high quantum yield (2.33) was observed in

the photosensitized dechlorination of PCBs with phenothiazine in alkaline isopropanol

(Hawari et al., 1992). Izadifard et al. (2008) successfully used leuco-methylene blue

(LMB) as a photosensitizer to treat PCBs in acetonitrile/water mixture under visible light

irradiation.

2.4.3 Photocatalytic degradation of pesticides and PCBs

TiO2 photocatalysis has been applied to the treatment of various pesticides including

amide herbicides, bipyridium herbicides, carbamate insecticides, chloroniotinoid

insectids, chlorophenol pesticides, halobenzonitrile pesticides, organochlroine

insecticides, organophosphorus pesticides, phenol-based pesticides, pyrimidine

pesticides, thiocarbamate herbicides, micellaneous, etc. (Kamble et al., 2004, Echavia et

al., 2009, Herrmann et al., 1998, Trillas et al., 1995, Serra et al., 1994, Chen et al., 1999,

Muneer and Boxall, 2008, Topalov et al., 2001, Ormad et al., 2010, Gelover et al., 2004,

Bamba et al., 2008, Kim et al., 2006, Zaleska et al., 2000, Muszkat et al., 1992, Burrows

et al., 2002). Photocatalytic degradation of PCBs with TiO2 was reported using light

ranging from 340 nm to 365 nm (Carey et al., 1976, Chiarenzelli et al., 1995, Wang and

Page 57: 1.3 Light emitting diode

38

Hong, 2000). Some researchers believe that hydroxyl radicals generated during TiO2

photocatalysis oxidize PCBs molecules, leading to PCBs photodegradation and the

eventual formation of CO2 (Wang and Hong, 2000). However, there are also reports of

hydroxyl radicals leading to oxygenation of the PCB ring which produces more toxic

compounds (Safe, 1994, Gierthy et al., 1997) .

2.5 Design of a photocatalytic reactor

A good photocatalytic reactor should be an appropriate combination of photocatalyst,

light source and geometry. The photocatalyst should be easily separated or immobilized;

the light source should be energy efficient; the geometry of reactor should make the

photocatalyst, target compound and photons come together efficiently; and it should be

scalable.

2.5.1 State of Photocatalyst in the Reactor

2.5.1.1 Slurry photocatalytic reactor vs immobilized photocatalytic reactor

A variety of photocatalytic reactors have been designed in the past two decades.

Generally, photocatalytic reactors can be classified into two major groups: slurry reactors

and fixed film reactors. In slurry reactors, the nanoparticle TiO2 is dispersed in the

solution. In the immobilized reactors, the photocatalysts are immobilized on an inert

substrate such as alumina pellets, molecular sieve, glass wall, glass fibre or ceramic

membranes (Parsons, 2004). Table 2-4 compares the two categories of reactors and

summarizes their advantages and disadvantages. The slurry photocatalytic reactors are

Page 58: 1.3 Light emitting diode

39

very efficient in terms of photons due to high surface area to volume ratio. However, the

use of slurries requires further separation steps involving either filtration, centrifugation

or coagulation, which increases the complexity of the overall processes and the

operational cost (Denny et al., 2009). The immobilized photoreactors are less efficient

Table 2-4: Slurry versus Immobilized Photocatalytic Systems (Lasa et al., 2005).

Slurry reactor Immobilized reactor

Advantages

• Uniform catalyst distribution

• High surface area /volume ratio

• Limited mass transfer

• Minimum catalyst fouling effects

• Well mixed particle suspension

• Low pressure drop

Advantages

• Continuous operation

• Improved removal of organic material

from water phase while using a

support with adsorption properties

• No need for catalyst separation

operation

Disadvantages

• Requires post-process seperation

step

• Important light scattering and

adsorption in the particle suspended

medium (Ollis et al., 1991))

Disadvantages

• Low light utilization efficiencies

• Restricted processing capacities

(Turchi and Ollis, 1988, Matthews

and McEvoy, 1992, Matthews, 1991)

• Possible catalyst deactivation and

catalyst wash out (Serrano and de

Lasa, 1997)

Page 59: 1.3 Light emitting diode

40

systems for pollutant degradation due to its smaller availability of illuminated surface

area per mass and substrate mass transport issues but it offers an advantage as the

secondary separation of catalyst from the treated water is not needed. One way to

improve the surface area to reaction volume ratio in an immobilized photocatalytic

reactor is to use supported semiconductor photocatalysts. They are a form of slurry

reactor with improved separation. In this type of photocatalytic reactor, TiO2 is coated on

small particles which can be easily separated (Geng and Cui, 2010, Imoberdorf et al.,

2008a, Vaisman et al., 2005, Pozzo et al., 2005, Kanki et al., 2005, Chiovetta et al., 2001,

Haarstrick et al., 1996).

2.5.1.2 TiO2 immobilization through electrochemical anodization

Immobilized TiO2 can be prepared through electrochemical anodization (Gong et al.,

2001, Paulose et al., 2006, Macak et al., 2005, Wang and Lin, 2008), dipping-coating

(Mikula et al., 1995), sol-gel method (Negishi et al., 1998, Negishi and Takeuchi, 2001,

Yu et al., 2001, Watanabe et al., 2000), chemical vapour deposition (Nakamura et al.,

2001, Kaliwoh et al., 2002, Watanabe et al., 2002), pulsed laser deposition (Yamamoto et

al., 2001) and reactive evaporation (Zeman and Takabayashi, 2002, Mergel et al., 2000).

Among these methods, electrochemical anodization is considered to be superior as it is

economical, convenient, and produce highly photoactive and mechanically durable films

(Li et al., 2013, Natarajan et al., 2011a, Yu et al., 2010, Xie and Li, 2006).

Zwilling et al. (1999) reported the first self-organized anodic TiO2 nanostructure using

electrochemical anodization approach. After that, many studies on fabricating anodic

Page 60: 1.3 Light emitting diode

41

TiO2 nanostructure have been reported. The anodization was carried out using a two-

electrode cell with titanium metals as anode and different materials (Pt, Pd, Ni, Fe, Co,

Al, Carbon and other materials ) as the counter electrode (Allam and Grimes, 2008).

Aqueous/non-aqueous electrolytes (ethylene glycol, glycerol, DMSO or ionic liquids)

containing approximately 0.05 M-0.5 M fluoride ions are used during anodization (Roy et

al., 2011). The applied voltage is between 1-30V for aqueous electrolyte and 5-150 V for

non-aqueous electrolytes. The formation of TiO2 nanostructure in these fluorine

containing electrolytes is the result of competition between the electrochemical oxidation

of Ti and electrical field induced etching of TiO2 as well as chemical etching of TiO2 by

fluorine ions (Wang and Lin, 2009). The fluoride ion can promotes the growth of anatase

TiO2 with high reactive facets such as (001) facets (Yang et al., 2008).

Usually, the TiO2 structure obtained via anodization at room temperature are in an

amorphous form. However, an amorphous form of TiO2 does not show a good

photocatalytic activity (Wu et al., 2011). To obtain TiO2 nanostructure with high

photocatalytic activity, the prepared TiO2 should be converted to crystallized form

(anatase/rutile) with high temperature annealing treatment (eg. 500 oC) (Roy et al., 2011).

2.5.2 Light source

The light source is a key component for photoreactor design. To select the appropriate

lamps, technical and economic considerations should be taken into consideration

(Oppenlander, 2003): firstly, the radiation source should provide the photons that can be

directly or indirectly utilized by the reactant. In TiO2 photocatalysis, only photons with

Page 61: 1.3 Light emitting diode

42

energy equal or larger than band-gap energy of TiO2 can be used, therefore, the

wavelength of chosen lamp should be shorter than 385 nm. Secondly, a high efficiency

light source should be chosen to make sure that, most of input electric energy can be

converted to desired light energy. This requires that unusable wavelengths (emission

wavelengths not aligned with absorption wavelengths) of light should be as little as

possible and the energy loss due to dissipation as heat should be limited. Thirdly, the

geometry and the size of lamps should not limit reactor design.

2.5.2.1 Sunlight

Figure 2-8: Solar spectral irradiance distribution on the surface of earth. [Reproduced from (Hulstrom et al., 1985) with permission]

0

200

400

600

800

1000

1200

1400

1600

1800

0 500 1000 1500 2000 2500

Irra

dian

ce (

W m

-2 μ

m-1

)

Wavelength (nm)

Page 62: 1.3 Light emitting diode

43

The sun is a spherical UV/VIS source with a radiant power of 3.842*1026 W

(Oppenlander, 2003). Only a small part of its radiation can reach the earth through

travelling the 1.4957*1011 m distance. The radiation received on the top of the earth's

atmosphere is around 1400 W m-2 (Ryer and Light, 1997). After passing through the

atmosphere, the solar radiation is attenuated due to the light absorption by the molecules

in the atmosphere. The solar radiation on the earth's surface strongly depends on the

weather condition, e.g. cloud, fog etc. With clear skies, the solar radiation at the earth's

surface is around 1000 W m-2 (Oppenlander, 2003). Also, it can vary with latitude, the

time of day and the season. The average annual solar irradiation on the earth's surface

range from 100 W m-2 (polar region) to about 300 W m-2 (desert regions) (Bolton, 1989).

Figure 2-9: Fractional cumulative integrated irradiance vs. wavelength. [Reproduced from (Hulstrom et al., 1985) with permission]

0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70 0.80 0.90 1.00

0 500 1000 1500 2000 2500 3000 3500 4000 4500

Fra

ctio

n

Wavelength (nm)

Page 63: 1.3 Light emitting diode

44

The spectral distribution of solar radiation received at the earth's surface is shown in

Figure 2-8. At the earth's surface, sunlight contains no VUV and UV-C radiation because

of their efficient absorption by oxygen (absorption at wavelength below 200 nm) and by

ozone ( absorption at wavelength below 330 nm), respectively, in the upper atmosphere,

(Finlayson-Pitts and Pitts Jr, 1986). Less than 5 % of solar radiation reaching on the

earth's surface is UV radiation (below 400 nm) and more than 95% of that is visible or

infrared radiation (Figure 2-9).

2.5.2.2 Mercury lamps

Mercury-vapor lamps are gas discharge lamps that use mercury in an excited state to

produce light. There are two major mercury lamp types based on the mercury pressure

inside the lamp: low pressure (LP) mercury lamp (0.001-13 mbar) and medium pressure

(MP) mercury lamp (~1333 mbar) (Oppenlander, 2003). LP lamps are extensively used

in the field of UV disinfection, which provide almost monochromatic UV radiation at

253.7 nm with an ordinary quartz envelope. The 253.7 nm monochromatic UV radiation

produced by the low pressure mercury lamp can be fairly efficiently converted to broader

band emission at longer wavelengths by coating the lamp envelope with a suitable set of

fluorescent materials, as in the fluorescent lamps used for interior lighting. LP mercury

lamps can convert 40% to 60% of electrical energy into radiant energy (Altena et al.,

2001). To generate a constant radiation output, the maximum electric power input of LP

mercury lamps is usually less than 300 W. MP mercury lamps can be operated with much

higher electrical input power up to 30 kW, but with a reduced UV radiant power

efficiency ( 30%-40%). Instead of monochromatic radiation, MP mercury lamps generate

Page 64: 1.3 Light emitting diode

45

polychromatic emission ranging from the UV (UVC~15-23%, UVB~6-7%, UVA~8%)

over the VIS (~15%) to the IR (47-55%)) (Oppenlander, 2003).

2.5.2.3 Light emitting diode

A new generation of lamps with promising features for photochemical applications has

been developed, the so-called light-emitting diodes (LED). The LED technology shows

many advantages over conventional light sources including energy savings (higher

current to light conversion), less heat production, longer lifetime (up to100 000 hours),

improved robustness, smaller size, faster switching, greater durability and reliability,

besides, it is more environmental friendly as it does not use mercury.

LED is a semiconductor light source, which consists of a chip of semiconducting material

doped with impurities to create a p-n junction (Schubert, 2006). A p-n junction consists

of n-type and p-type semiconductors. P-type semiconductors are a type of semiconductor

which is capable of providing extra positive charge (holes). N-type semiconductors are a

type of semiconductor which can provide an excess of negative electron charge carriers

(electrons). Both types of semiconductor are obtained by doping or adding an electron

acceptor/donor to the semiconductor in order to increase the number of free charge

carriers. The LED working mechanism is shown in Figure 2-10. When a light-emitting

diode is forward biased, the electron in the N part of the junction will be ejected into the

p-part of junction. The combination of an electron-hole pair can then lead to an emission

of a photon. The wavelength of photons released depends on the band gap energy of the

materials forming the p-n junction. The visible spectrum LEDs can be fabricated using

Page 65: 1.3 Light emitting diode

46

GaAsP, GaP, AlGaAs, GaAs, AlGaInP, GaAs, GsAsN or other semiconductors

(Schubert, 2006). The UV LEDs are mainly made of the Group III-nitride-based

semiconductors such as Boron nitride (Mishima et al., 1988), AlGaN (Yasan et al., 2002,

Adivarahan et al., 2004, Yasan et al., 2003, Chitnis et al., 2003), AlN (Khan et al., 2008,

Taniyasu et al., 2006) and AlGaInN (Khan et al., 2008, Kipshidze et al., 2002, Kipshidze

et al., 2003).

Figure 2-10: An inner working on an LED. [Adapted from (Wikipedia, 2011)]

In past few years, the energy efficiency and the output power of UV LEDs have

improved significantly, and the average price has dropped. So far, LED of low intensity

associated with UVA/UVB applications represented 89% of the overall UV-LED market

(Semiconductor Today, 2013). The major UV-LEDs producers include Nichia

Corporation, Lumileds, Cree, Mitsubishi, PARC, NTT, RIKEN, Nitride Semiconductors,

Page 66: 1.3 Light emitting diode

47

SETI, SEMILEDS, etc. Their products mainly emit the photons with wavelengths equal

or above 365nm and with power output ranging from a few micro watts at short

wavelengths to several watts at wavelengths around 365nm.

External quantum efficiency (EQE) reflects the energy efficiency of LED, which is

defined as the ratio of the number of photons emitted from the LED to the number of

electrons passing through the device. An excellent EQE (up to 40%) has been obtained

for LEDs in the UVA part of the spectral range (LEDs Magazine, 2012). However, the

average EQE of UV LED shorter than 365nm is at least one order of magnitude below

the best devices in the near UV wavelength range (Kneissl et al., 2011). Therefore,

applications of UVC-LED are still in their infancy and mainly for R&D purposes and

analytic instruments (Semiconductor Today, 2013). Following the same development

trend of UVA-LED, the journey on the path to efficient UVC-LED has just begun and

there are many optimistic reasons to produce highly efficient UVC-LED in the near

future (Kneissl et al., 2011).

2.5.3 Artificially illuminated photocatalytic reactors

The photocatalytic reactors using conventional UV mercury lamps can be classified as

two major categories; slurry type photocatalytic reactors and immobilized photocatalytic

reactors. The major geometries of slurry type photocatalytic reactors include:

annular reactors (Figure 2-11 (a)) which consists of two co-axial cylinders that

define the reaction zone (Mo et al., 2008, Johnson and Mehrvar, 2008, Chong et

al., 2009, Behnajady et al., 2009, Imoberdorf et al., 2007, Tang and Chen, 2004),

Page 67: 1.3 Light emitting diode

48

immersion well reactors (Figure 2-11 (b) and (c)) where one or more lamps are

immersed in the well stirred reactors (Thiruvenkatachari et al., 2005, Saquib and

Muneer, 2003),

elliptical reactors (Figure-2-11 (d)) of which foci are occupied by lamps and the

tubular reactor (Jacob and Dranoff, 1969),

parabolic reactors (Figure-2-11 (e)) of which foci are occupied by lamps (Alfano

et al., 2000).

Figure 2-11: Various photochemical reactor configurations. [Reproduced from (Pareek et al., 2008) with permission]

All the slurry type photocatalytic reactors should be incorporated with the photocatalyst

separation systems such as membrane filtration. To avoid this process, the immobilized

Page 68: 1.3 Light emitting diode

49

photocatalytic reactor is developed. One of the earliest type of immobilized

photocatalytic reactor is a spiral photoreactor where the photocatalyst is coated onto the

inner walls of a glass spiral which has a UV lamp in the centre (Matthews, 1987). Since

then, immobilized photocatalytic reactors with various designs have been developed. A

variety of immobilized photocatalytic reactors are listed below:

• Thin film reactor (Roselin and Selvin, 2011) in which a thin reacting fluid film

flowing through a surface coated with photocatalyst. The UV irradiation

penetrates the fluid film to reach the photocatalytic surface.

• Multiple tube reactor (Figure 2-12) containing of a cylindrical vessel within

which a number of hollow quartz glass tubes externally coated with photocatalyst

were placed. The liquid flows through the shell-side over the outside surfaces of

the coated tubes while the light travels through the inside of hollow tubes via an

aluminum reflector.

• Fiber optic cable reactor (Figure 2-13) in which a number of fiber optic cables

were coated with TiO2. The UV light can reach the supported TiO2 along the

optic cable.

• Rotating disk reactor (Figure 2-14) composed of a rotating disk coated with

photocatalyst. A thin film of liquid becomes entrained on the disc from the bulk

solution during rotation. Reaction takes place in the head space due to the

illumination.

• Distributive photocatalytic reactor (Figure 2-15) in which light conductors coated

on its outside surface with catalysts are embedded vertically. This configuration

provides a higher illuminated catalyst area per volume.

Page 69: 1.3 Light emitting diode

50

• Taylor vortex reactor (Figure 2-16) consisting of two coaxial cylinders. The inner

cylinder coated with TiO2 is rotated to generate a vortex-induced fluid instability.

• Fluidized bed photocatalytic reactor (Figure 2-17) which has a fluidized bed

consisting of small TiO2-coated particles such as glass beads. This configuration

improves the surface area to reaction volume ratio and mass transfer condition.

Apart from these reactors, there are several novel designs which are useful for specialized

applications, however, two inadequacies limit their use (Pareek et al., 2008). Firstly,

complex mechanical designs of novel photoreactors make construction difficult and

hinder their routine maintenance and cleaning. Secondly, the processing capacity of the

novel reactors is limited.

Figure 2-12: Scheme of a multiple tube reactor. [Reproduced from (Ray and Beenackers, 1998) with permission]

Page 70: 1.3 Light emitting diode

51

Figure 2-13: Scheme of an optical fibre photocatalytic reactor. [Reproduced from (Nguyen and Wu, 2008) with the permission]

Figure 2-14: Scheme of a rotating disk reactor. [Reproduced from (Hamill et al., 2001) with the permission]

Page 71: 1.3 Light emitting diode

52

Figure 2-15: Top view of a distributive photocatalytic reactor. [Reproduced from (Ray and Beenackers, 1998) with permission]

Figure 2-16: Experimental setup of taylor vortex photocatalytic reactor:(1) motor, (2) speed controller, (3) gear coupling, (4) UV lamp (5) sample collection point (6) lamp holder (7) outer cylinder and (8) catalyst-coated inner cylinder. [Reproduced from (Dutta and Ray, 2004) with the permission]

Page 72: 1.3 Light emitting diode

53

Figure 2-17: Scheme of a fluidized photocatalytic reactor. [Reproduced from (Vaisman et al., 2005) with permission]

2.5.4 Solar photocatalytic reactors:

Major designs of solar photocatalyic reactors have been reviewed by several authors

(Braham and Harris, 2009, Lasa et al., 2005, Thiruvenkatachari et al., 2008), including

parabolic trough reactors, compound parabolic reactors, inclined plate photocatalytic

reactors, double-skin sheet photocatalytic reactors, horizontal rotating disk reactors and

water bell reactors. A parabolic trough reactor (Figure 2-18 a) is a light concentrating-

type unit, which uses a long parabolic reflecting trough to concentrate solar radiation on a

transparent tubular reactor placed on the parabolic focal line. Compound parabolic

reactors (Figure 2-18 b) are trough reactors without light concentrating devices. The

reflector in compound parabolic reactor is characterized with a two half-cylinders of

Page 73: 1.3 Light emitting diode

54

Figure 2-18: Solar photocatalytic reactor: (a) parabolic trough reactor (PTR) (b) compound parabolic collector (CPC). [Reproduced from (Braham and Harris, 2009) with permission]

parabolic profile which allow indirect light to be reflected onto the tubular reactor. An

inclined plate photocatalytic reactor (Figure 2-19 a) consists of an inclined surface coated

Page 74: 1.3 Light emitting diode

55

with photocatalyst. The reactant fluid flows through the inclined surface to form a thin

film. An double-skin sheet photoreactor (Figure 2-19 b) uses a double-skin transparent

plexiglass to construct a long, convoluted back and forth channel on a flat plane through

which the reactant fluid flow. A horizontal rotating disk reactor (Figure 2-20 a) has a

rotating disk of which the surface is exposed to sunlight. The reactant fluid is injected up

from the center of the disk and forms a fluid film on the surface of the disk. The rotation

of disk generates a turbulent flow regime in the fluid film. A water bell reactor (Figure 2-

20 b) features a nozzle which sprays a continuous and unsupported thin film of liquid

exposed to solar irradiation.

Figure 2-19. Typical reactor layout for an (a) inclined plate collector and (b) double skin sheet photoreactor. [Reproduced from (Braham and Harris, 2009) with permission]

Page 75: 1.3 Light emitting diode

56

Figure 2-20: Typical reactor layout for (a) horizontal rotating disk reactor and (b) water bell reactor. [Reproduced from (Braham and Harris, 2009) with permission]

2.6 Radiation-field modelling

A uniform distribution of light is necessary for an efficient photocatalytic treatment

system. In this context, a radiation field model is very useful. The model need to compute

the rate of photon absorption at any position within the reactor or 'local volumetric rate of

energy absorption' (LVREA) (Cassano et al., 1995). In homogenous media, the change of

the light intensity along the direction of photon propagation is due to absorption process

in the reaction media, while, in heterogeneous media scattering of radiation by particles

should also be accounted in the variation of light intensity. Various radiation field models

in homogenous or heterogeneous environments have been described in several papers

(Imoberdorf et al., 2008b, Jacob and Dranoff, 1969, Jacobm and Dranoff, 1970, Irazoqui

et al., 1973, Alfano et al., 1986a, Alfano et al., 1986b, Alfano et al., 1986c).

Page 76: 1.3 Light emitting diode

57

2.6.1 The Radiation transport equation (RTE)

In heterogeneous media (slurry TiO2 photocatalysis), the light intensity may change

along its light path due to photon absorption, scattering and emission. In-scattering and

emission from reacting mixture can increase the light intensity, while, the photon

absorption and out-scattering can reduce the light intensity (Figure 2-21).

Figure 2-21: Schematic for photon transport.

In a control volume V (m3), the photon balance equation can be described as (Cassano et

al., 1995):

𝑑𝑄𝑑𝑡

+ 𝐴 ∗ 𝑑𝐼𝜆 = −𝑞𝑎 + 𝑞𝑒 + 𝑞𝑖𝑛 − 𝑞𝑜𝑢𝑡 [2-25]

Where is the number of photon in control volume (photon), t is time (s), A is the cross

section surface area (m2), is the spectral light intensity (photon m-2 s-1), is rate of

Page 77: 1.3 Light emitting diode

58

photon absorption (photon s-1), is rate of photon emission (photon s-1), is rate of

photon in-scattered (photon s-1), is rate of photon out-scattered (photon s-1).

At steady state: =0, then the equation can be written as:

𝐴 ∗ 𝑑𝐼𝜆 = −𝑞𝑎 + 𝑞𝑒 + 𝑞𝑖𝑛 − 𝑞𝑜𝑢𝑡 [2-26]

Equation [2-27] is obtained by dividing Equation [2-26] with V; V=A*ds,

𝐴𝑑𝐼𝜆𝑉

= −𝑞𝑎/V + 𝑞𝑒/v + 𝑞𝑖𝑛/v − 𝑞𝑜𝑢𝑡/v [2-27]

Finally, the photon balance in a control volume can be described by Equation [2-28]

(Pareek et al., 2008, Cassano et al., 1995).

𝑑𝐼𝜆(𝑠,𝛺)𝑑𝑠

= −𝑊𝑎 + 𝑊𝑒 + 𝑊𝑖𝑛 −𝑊𝑜𝑢𝑡 [2-28]

Where is the solid angle (steradian), is volumetric rate of photon absorption

(photon m-3 s-1), is volumetric rate of photon emission (photon m-3 s-1), is

volumetric rate of photon in-scattered (photon m-3 s-1), is volumetric rate of photon

out-scattered (photon m-3 s-1).

At normal or low temperature, the spontaneous radiation emission can be neglected:

𝑊𝑒 = 0 [2-29]

Page 78: 1.3 Light emitting diode

59

In general, linear isotropic constitutive equations (Equation [2-30] &[2-31])can be used

to characterize absorption and out-scattering (Cassano et al., 1995),

𝑊𝑎 = 𝛼𝜆𝐼𝜆(𝑠,𝛺) [2-30]

𝑊𝑜𝑢𝑡 = 𝜎𝜆𝐼𝜆(𝑠,𝛺) [2-31]

Where is the volumetric absorption coefficient (m-1) and is the volumetric

scattering coefficient (m-1), usually, the combination of the absorption coefficient and the

scattering coefficient is defined as the extinction coefficient:

βλ = 𝛼𝜆 + 𝜎𝜆 [2-32]

Where is the extinction coefficient (m-1).

In the RTE, In-scattering is a more complicated term, which can be described by

Equation [2-33] (Cassano et al., 1995). Two assumptions were made here:

• Every scattering is independent of each other.

• The scattering is elastic, which means that the frequency of scattered radiation is

the same as incident radiation.

𝑊𝑖𝑛 =1

4𝜋𝜎𝜆 � 𝑝(𝛺′ ⟶ 𝛺)𝐼𝜆(𝑠,𝛺′)𝑑𝛺′

4𝜋

0 [2-33]

Where is the phase function for the in-scattering of photons, usually, the

phase function is normalized according to Equation [2-34]:

14𝜋

� 𝑝(𝛺′ ⟶ 𝛺)𝑑𝛺′ = 14𝜋

0 [2-34]

Page 79: 1.3 Light emitting diode

60

The phase function is the one that make the RTE difficult to solve. To simulate the model

more efficiently, the Henyey-Greenstein (HG) phase function (Equation [37]) can be an

appropriate choice (Marugán et al., 2006).

𝑃𝐻𝐺(𝜑) =1

4π∗

1 − τ2

(1 + τ2 − 2τ ∗ cos(φ))3/2 [2-35]

Where is the scattering angle and τ is the asymmetry factor of the scattered radiation

distribution. The value of τ varies smoothly from -1 to +1. When τ =0, it represents an

isotropic phase function.

According to Equations [2-25] ~ [2-35], the final RTE can be written as

𝑑𝐼𝜆(𝑠,𝛺)𝑑𝑠

= −βλ𝐼𝜆(𝑠,𝛺) +1

4𝜋𝜎𝜆 � 𝑝(𝛺′ ⟶ 𝛺)𝐼𝜆(𝑠,𝛺′)𝑑𝛺′

4𝜋

0 [2-36]

For homogenous media (when a photocatalyst is immobilized), the scattering term on

Equation [2-36] can be excluded.

Then the incident light intensity from all direction:

Gv(s) = � 𝐼𝜆(𝑠,𝛺)𝑑𝛺4𝜋

0 [2-37]

Where is the incident light intensity (photon m-2 s-1).

And the LVREA at any point is given by

LVREA= αλGv(S) [2-38]

2.6.2 Numerical methods to solve the RTE

The RTE is an integral-differential equation, an exact analytical solution is impossible

except for homogeneous photoreaction systems, where scattering phenomenon is not

taken into account (Pareek et al., 2008). Numerical method offers a viable alternative to

Page 80: 1.3 Light emitting diode

61

solve the RTE. Carvalho and Farias (1998) reviewed a variety of methods developed to

numerically solve the RTE, including Zone method, Monte Carlo (MC) method, flux

method and hybrid method. Among these methods, the MC method is been accepted as

efficient and reliable. (Pareek et al., 2008, Pasquali et al., 1996). MC is a statistical

method, which is based on following the probable trajectories and fates of photons inside

the reaction zone, until their final absorption in the system or existing out of the system.

Both the trajectories and fates are decided with the help of random numbers, which are

generated by a random function through computer (Pareek et al., 2008). Consider a

photon entering into the reaction zone, it may get absorbed by the particle and its

trajectory ends there or it may be scattered by the particle to a new direction and the

trajectory continues until being absorbed by other particles or exiting out of the reaction

zone (Pareek et al., 2008). Whether absorption or scattering is determined by a random

choice based on the absorption coefficient and the scattering coefficient. To solve the

RTE with the Monte Carlo method, the optical properties of the reaction medium like

absorption coefficient scattering coefficient and the phase function should be obtained.

Further, the boundary conditions are needed.

2.6.3 Radiation source models

Radiation source model are the functions predicting the light intensity emitted from the

lamps. When the scattering effect is negligible such as in homogeneous medium, the

radiation source model can be directly used as a radiation filed model. The radiation

source model also play an important role in solving the boundary condition of RTE.

Alfano et al. (1986b) reviewed a number of radiation source models and classified them

Page 81: 1.3 Light emitting diode

62

into two categories: incidence models and emission models. Three incidence models have

been proposed (Pareek et al., 2008, Alfano et al., 1986b), including the radial incidence

model, the partially diffuse incidence model and the diffuse incidence model. The

development of incidence models requires an existing specific radiant energy distribution

to be assumed in the reactor space. Besides, these incidence model always need one or

more experimentally adjustable parameters, which is dependent on the size and the

configuration of the reactor (Alfano et al., 1986b). To overcome this problem, emission

models based on the lamp emission are developed and regarded as a preferred choice for

radiation source modeling. A lamp may be regarded as a point (LED), a line, a surface, or

a volume source. Depending upon the nature of the lamp, different emission models such

as line source model, surface source model, volume source model have been developed

(Pareek et al., 2008, Alfano et al., 1986b). The line source models were considered as

appropriate methods to simulate the light intensity distribution over the photocatalytic

plate when the photocatalytic reactor is equipped with tubular lamps (the geometry of

conventional mercury lamp) (Salvadó-Estivill et al., 2007).

A good radiation field model can accurately predicted the light intensity distribution

within photoreactors. Such model can be used as a tool to figure out the optimal

arrangement of light source and reactor geometry. Furthermore, radiation field model is

very important to the mathematically simulating photochemical treatment processes.

Page 82: 1.3 Light emitting diode

63

Chapter Three: ELECTRON TRANSFER SENSITIZED PHOTODECHLORINATION OF SURFACTANT SOLUBILIZED PCB138

3.1 Introduction

Direct and sensitized photodechlorination of polychlorinated biphenyls (PCBs) dissolved

in organic or organic/water mixtures have been the focus of various investigations (Bunce

et al., 1978, Ruzo et al., 1974, Hawari et al., 1992, Hawari et al., 1991, Miao et al., 1996,

Izadifard et al., 2008, Lin et al., 1996, Jakher et al., 2007). The low solubility of PCBs in

water necessitates the presence of an organic solvent. Surfactants also have been used to

solubilize PCBs in water (Yao et al., 2000, Bunce et al., 1978, Hawari et al., 1992,

Hawari et al., 1991, Miao et al., 1999). Using a surfactant instead of an organic solvent is

advantageous for two reasons: lower cost and minimized side reactions (Chu et al., 1998).

To-date investigations on PCBs dissolved in water by surfactants has been restricted to

direct UV photolysis, which requires high energy photons with wavelengths less than 300

nm (Chu et al., 2005, Chu et al., 1998). The application of sensitized dechlorination can

make use of photons with longer wavelengths, eventually leading to using sunlight for

dechlorination. Besides, sensitized dechlorination of PCBs can also result in a high

photodegradation rate (Dhol, 2005). To the best of our knowledge, no research is reported

on sensitized dechlorination of PCBs dissolved in water using surfactants; though there

are a few reports on sensitized reaction, where an aliphatic amine which cannot function

as a sensitizer is used to enhance the efficiency of the reaction (Chu and Kwan, 2002,

Chu and Kwan, 2003). In this case, PCB itself must be excited so that an electron transfer

from the aliphatic amine to PCBs becomes favorable.

Page 83: 1.3 Light emitting diode

64

This paper presents the results of an investigation on sensitized dechlorination of a PCB

congener - 2,2',3,4,4',5'-hexachlorobiphenyl (PCB 138) and a commercial mixture of

PCBs (Aroclor 1254). PCBs were dissolved in water using an anionic surfactant (sodium

dodecyl sulfate, SDS), a nonionic surfactant (polyoxyethylene (80) sorbitan monooleate,

TWEEN 80) or a cationic surfactant (cetyltrimethylammonium bromide, CTAB). The

sensitizer of choice is leuco-methylene blue (LMB), which has been reported to

effectively dechlorinate PCBs under blue and near UV irradiation (Izadifard et al.,

2010a). LMB, while being stable in an oxygen devoid environment, is produced

efficiently upon reduction of methylene blue (MB). MB can be reduced under red light

and in the presence of an aliphatic amine (such as triethylamine, TEA), or by a thermal

reaction using sodium borohydride (NaBH4). Both approaches are studied in this paper.

Air oxidation of LMB closes a ‘catalytic’ cycle that consumes the reductant.

3.2 Materials and methods

3.2.1 Materials

All PCB congeners and Aroclor 1254 were purchased from Chromatographic Specialties

Inc.; MB, CTAB, and 99.5% pure TEA were obtained from Sigma; 99.8% pure hexane

and 98% pure sodium borohydride were procured from EMD; ultra pure SDS was

obtained from MP Biomedicals and TWEEN 80 was purchased from VWR. All reagents

were used as received. Milli-Q ultrapure water was used in the experiments.

Page 84: 1.3 Light emitting diode

65

3.2.2 Methods

3.2.2.1 PCB 138 solubilization with surfactants

For the selected surfactant (SDS, CTAB and TWEEN 80) a stock solution with 4 g L-1

concentration was prepared. Each surfactant stock solution was prepared by dissolving 4

g of pure surfactant in 1 L milli-Q water with the aid of sonication. The PCB stock

solution was prepared by dissolving 10 mg PCB138 or 10 mg Aroclor 1254 in 10 ml

acetonitrile. For each photolysis experiment, a certain amount of PCB stock solution,

PCB 138 or Aroclor 1254, in acetonitrile (1000 mg L-1) was pipetted into a 20 ml Pyrex

glass vial and, left at room temperature in a fume hood for one day to evaporate the

acetonitrile completely. Then a known amount of surfactant stock solution was added to

the vial and the mixture was left in the sonicator for 4 hours to prepare the PCB

surfactant solution.

3.2.2.2 Photochemical reaction

The sensitizer of choice, LMB, was generated in two ways: (1) reaction between MB and

TEA under visible light irradiation and (2) thermal reaction between MB and sodium

borohydride. In each case, to the prepared PCB surfactant mixture was added a MB

solution along with either TEA or NaBH4. The final solution volume was made equal to

20 ml with milli-Q water. Uniform mixing during irradiation was achieved by placing the

sample vial on a magnetic stirring plate. The samples (contained in a Pyrex glass vial)

were irradiated in a Rayonet photoreactor equipped with either 8 or 14 cool white

fluorescent lamps. The intensity of light (Io) was measured using ferrioxalate actinometry

Page 85: 1.3 Light emitting diode

66

that monitored wavelengths from 254 to 500 nm, covering the region of LMB absorption

(Calvert and Pitts, 1966). Values were 5.2×1016 photon s-1 for 14 lamps and 3.0×1016

photon s-1 for 8 lamps. This convenient actinometer measures the light intensity inside the

reaction vessel and provides an order of magnitude value of intensity in the region of

LMB absorbance. Our irradiation source, the cool white lamps emit light in visible range,

up to 700 nm. MB absorption peak is at 660 nm, so it can be effectively reduced to LMB

with the cool white lamps. It is the actinometry method used here that measures

wavelengths between 254 nm to 500 nm. Since our sensitizer is LMB, whose absorption

is within this range, we used this actinometry.

3.2.2.3 Sampling, extraction and GC analysis

Exactly 0.5 ml illuminated samples were taken at different irradiation times and

transferred into a small glass vial. To each aliquot was added 2 ml of hexane, the vial was

covered by aluminum foil and was left in the wrist shaker for 1 hour to extract PCBs

from the water mixture. Around 80% extraction efficiency was obtained following this

procedure.

The extracted PCBs were analyzed using an Agilent 6890 gas chromatograph equipped

with auto-sampler and electron capture detector (ECD), using a fused silica capillary

column DB608. Helium was used as the carrier gas with a flow rate of 1.5 ml min-1 and

the temperature of the injection port was 280 oC. The GC/ECD temperature programming

was set up as follow: The initial temperature for each run was set at 80 oC, which was

Page 86: 1.3 Light emitting diode

67

ramped up to 180 oC at a rate of 10 oC min-1; the temperature beyond 180 oC was ramped

at a rate of 3 oC until it reached 270 oC where it was held for 15 minutes (USEPA,

1996b). For Aroclor 1254, six major peaks were chosen and a multipoint calibration

curve was made by using those six peaks.

Each experiment was conducted in duplicate and results were reported as an error bar

along with the average.

3.3 Results and Discussion

3.3.1 Selectivity of surfactants

The critical micelle concentration (CMC) in water for SDS, CTAB and TWEEN 80 were,

respectively, 2300 mg L-1 (Mandal et al., 1988), 324 mg L-1 (Paredes et al., 1984) and 15

mg L-1 (Hillgren et al., 2002). The sensitized dechlorination of PCB 138 in these

surfactant solutions are shown in Figure 3-1 and Figure 3-2. Figure 3-1 presents results

where LMB was produced under irradiation, while Figure 3-2 where it was produced

thermally. Based on our previous experiments published elsewhere (Izadifard et al.

2010b), the NaBH4 based reactions are faster than TEA based reactions. Consequently, to

ensure that we can reliably measure the concentration of PCB (above the detection limit)

within the irradiation time, a higher concentration of PCB and a lower light intensity was

applied in the NaBH4 system. We are fully aware that two different initial concentrations

were used for the two experiments, that is why we are careful in the paper not to compare

Page 87: 1.3 Light emitting diode

68

the results of Figure 3-1 with those of Figure 3-2. We have instead compared the

performance of different surfactants amongst themselves, presented in either Figure 3-1

or Figure 3-2. Of the three surfactants investigated, the dechlorination efficiency of PCB

138 solubilized with TWEEN 80 or with CTAB are similar, when TEA is used. However,

in the NaBH4 system the performance of CTAB is better than that of TWEEN 80. In

addition, it takes only half the irradiation intensity to achieve similar results with NaBH4

reduction.

Figure 3-1: Reductive dechlorination of PCB 138 using LMB with TEA as the

reducing agent; [PCB 138] = 6.6 mg L-1, [SDS] = [CTAB] = [TWEEN 80] = 3.2 g L-1,

[MB] = 750 mg L-1, [TEA] = 68 g L-1, Io = 5.2×1016 photon s-1 .

0

0.2

0.4

0.6

0.8

1

1.2

0 5 10 15 20 25 30

C/C

o

Irradiation time (min)

SDS CTAB TWEEN80

Page 88: 1.3 Light emitting diode

69

Figure 3-2: Reductive dechlorination of PCB 138 using LMB with NaBH4 as the

reducing agent; [PCB 138] = 20 mg L-1, [SDS] = [CTAB] = [TWEEN 80] = 3.2 g L-1;

[MB] = 600 mg L-1; [NaBH4] = 20 g L-1, Io = 3.0×1016 photon s-1.

In both cases, dechlorination of PCB 138 is less efficient if SDS is used. It is worth

noting that the different surfactants were all used at the same concentration, which was

above the CMC of all of them. There are three possibilities for these results:

(1) CMC for SDS is much higher than that for CTAB and TWEEN 80, which correlate

to lesser solubilization of the PCBs. Possibly, the concentration of SDS used was

0

0.2

0.4

0.6

0.8

1

1.2

0 5 10 15 20 25 30

C/C

o

Irradiation time (min)

SDS CTAB TWEEN80

Page 89: 1.3 Light emitting diode

70

insufficient to completely solubilize PCB 138, and the extent of PCB 138 solubilization

in surfactant solution influenced the dechlorination efficiency. (This was investigated by

experimenting with higher concentrations of SDS, see below)

(2) The hydrophobic chain of surfactant is the hydrogen source for photodechlorination.

Possibly, CTAB and TWEEN 80, with longer hydrophobic chains than SDS, may orient

better with respect to the aryl radical and prevent aryl Cl⋅ recombination.

(3) Dye molecule incorporation in micelle species is more favorable for the cationic dye

and the anionic surfactant, affecting results (Aksu et al., 2010).

To evaluate the first possibility and study the effect of the concentration of surfactant on

the reaction rate, for the MB-TEA system SDS, CTAB and TWEEN 80 were tested at

higher concentrations than their corresponding CMCs. An apparent first order kinetic rate

coefficient for each situation appears in Table 3-1. As the concentration of SDS

increased, the extent of PCB138 solubilization increased and the declorination efficiency

improved. However, as the concentration of CTAB and TWEEN 80 becomes much

higher than CMC, the rate of dechlorination decreases. Possibly, at higher concentration

of the surfactants, LMB and PCB 138 are bound at sites distant from each other.

Since our ultimate objective is to develop a system that can photodegrade PCBs extracted

from soils and sediments and knowing that cationic surfactants are not good choices for a

surfactant-aided soil washing system as they can strongly adsorb onto soil (Wang and

Page 90: 1.3 Light emitting diode

71

Keller, 2009) and the anionic surfactant (SDS) did not give good results, the nonionic

surfactant (TWEEN 80) was chosen for further investigations except for the pathway

study.

Table 3-1: First order rate coefficients (K) for PCB138 dechlorination in TEA-MB

system with different concentration of surfactants.

[PCB138]

mg L-1

Type of

Surfactant

[Surfactant]

mg L-1

[MB]

mg L-1

[TEA]

g L-1

Io

photon s-1

K

s-1

6.6 SDS 3200 750 68 5.2 ×10-16 3.0 ×10-4

6.6 SDS 700 750 68 5.2 ×10-16 7.9 ×10-4

6.6 CTAB 914 750 68 5.2 ×10-16 3.9 ×10-3

6.6 CTAB 3430 750 68 5.2 ×10-16 1.2 ×10-3

6.6 CTAB 11500 750 68 5.2 ×10-16 9.4 ×10-4

6.6 TWEEN 80 45.7 750 68 5.2 ×10-16 3.0 ×10-3

6.6 TWEEN 80 228 750 68 5.2 ×10-16 4.3 ×10-3

6.6 TWEEN 80 914 750 68 5.2 ×10-16 2.3 ×10-3

6.6 TWEEN 80 2290 750 68 5.2 ×10-16 1.8 ×10-3

Page 91: 1.3 Light emitting diode

72

3.3.2 Dechlorination of PCBs in TEA and NaBH4 systems

As mentioned before, both the TEA and NaBH4 systems can be used to reduce MB to

LMB. Depending on the light sources available either one or both can be used. If using

sunlight is the final goal, both NaBH4 and TEA can be used. In case of using

approximately monochromatic light sources such as LEDs at 436 nm (Izadifard et al.,

2010b), using a strong reductant such as NaBH4 is unavoidable.

3.3.2.1 MB and TEA

Figure 3-3: Dechlorination of PCB 138 solubilized with TWEEN 80 in the presence of MB and different concentrations of TEA, [PCB 138] = 6.6 mg L-1, [MB] = 750 mg L-1, [TWEEN80] = 914 mg L-1, [TEA]1 = 22.6 g L-1 [TEA]2 = 31.5 g L-1, [TEA]3 = 45 g L-1, [TEA]4 = 68 g L-1, Io = 5.2×1016 photon s-1.

0

0.2

0.4

0.6

0.8

1

1.2

0 10 20 30 40 50 60

C/C

o

Irradiation time (min)

TEA₁ TEA₂ TEA₃ TEA₄

Page 92: 1.3 Light emitting diode

73

Figure 3-3 presents the results of an optimization study, wherein PCBs solubilised with

TWEEN 80 are dechlorinated in the presence of MB with varying concentrations of TEA.

The result indicates that when 23 g L-1 of TEA was used, there was no significant

dechlorination after a hour of irradiation; as the concentration of TEA increased to 31.5 g

L-1, more than 90% of PCB138 dechlorinated in 40 minutes, and it only took 20 minutes

to obtain the same removal efficiency at a concentration of 45 g L-1 of TEA. The

increased concentration of TEA in can improve the removal efficiency, however, the

limited solubility of TEA in water provides an upper limit to the TEA concentration that

can be used.

3.3.2.2 MB and NaBH4

For this system, effects of changing the concentrations of both MB and NaBH4 on the

reaction rate were studied (Table 3-2). In the chosen range of concentrations for MB, the

lowest concentration provided the best results, which can be due to the self quenching of

the dye at high concentrations (Turro, 1991). In the chosen range of concentrations for

NaBH4, a higher concentration of NaBH4 led to a higher reaction rate, however, an

experiment where the concentration of NaBH4 was doubled did not contribute to a

significant improvement in photodechlorination rate once the concentration of NaBH4

exceeded 10 g L-1.

Page 93: 1.3 Light emitting diode

74

Table 3-2: First order rate coefficients (K) for PCB138 dechlorination in NaBH4-

MB system.

[PCB138]

mg L-1

Type of

Surfactant

[Surfactant]

mg L-1

[MB]

mg L-1

[NaBH4]

g L-1

Io

photon s-1

K

s-1

20 TWEEN 80 1600 600 2 3.0 ×1016 2.1 ×10-3

20 TWEEN 80 1600 600 5 3.0 ×1016 2.5 ×10-3

20 TWEEN 80 1600 600 8 3.0 ×1016 4.0 ×10-3

20 TWEEN 80 1600 600 10 3.0 ×1016 5.0 ×10-3

20 TWEEN 80 1600 600 20 3.0 ×1016 5.1 ×10-3

20 TWEEN 80 1600 300 20 3.0 ×1016 6.0 ×10-3

20 TWEEN 80 1600 60 20 3.0 ×1016 6.5 ×10-3

3.3.2.3 Photodegradation of Aroclor 1254 with NaBH4 and TEA

A commercial mixure of PCBs (Aroclor 1254) was chosen to explore a practical case

using an LMB-surfactant system. To this end, a 10 mg L-1 Aroclor 1254 was studied

using both the TEA and the NaBH4 systems. The results indicate that Aroclor 1254 was >

95% dechlorinated within 10 minutes in both TEA and NaBH4 systems under the present

white light (see Figure 3-4). In this system, there is no evidence to support the chain

reaction mechanism advanced by Izadifard et al. (2010b).

Page 94: 1.3 Light emitting diode

75

Figure 3-4: Dechlorination of Aroclor1254 solubilized with TWEEN 80 in the

presence of MB and TEA or NaBH4: [Aroclor1254] = 10 mg L-1, [MB] = 600 mg L-1,

[TWEEN80] = 1.6 g L-1, [TEA] = 108 g L-1, [NaBH4] = 20 g L-1, Io = 3.0 ×1016 photon

s-1.

3.3.3 The dechlorination pathways of PCB 138 using CTAB and TWEEN 80

In all CTAB and TWEEN 80 cases, a stepwise dechlorination is evident. PCB 138 loss is

accompanied by appearance of lower chlorinated PCBs. Penta, tetra, tri, di and mono

chloro PCBs, initially increase as PCB138 decreases. These photoproducts were

identified by comparing the retention time of the photoproducts to those of available

standards. Whether pathways were influenced by surfactants was examined by comparing

product distributions after 6 minutes of irradiation (Figure 3-5 & Figure 3-6). The results

0

0.2

0.4

0.6

0.8

1

1.2

0 5 10 15 20 25 30

C/C

o

Irradiation time (min)

TEA NaBH₄

Page 95: 1.3 Light emitting diode

76

show that in MB-TEA case, for both surfactants, the primary photoproduct was PCB 118,

which is produced by losing an ortho chlorine. In case of MB-NaBH4 system, it was PCB

87 which appeared as the primary product by losing a chlorine at the meta position. This

may be attributable to nucleophilic attack by BH4- at the meta position as suggested by

the calculated charge distributions on carbons at different position in PCB congeners

(Chang et al., 2003): carbons at meta positions have lower charge density than those in

ortho positions.

Figure 3-5: The product distribution (after six minutes irradiation) for

dechlorination of PCB 138 solubilized by TWEEN 80 (1.6 g L-1) or CTAB (1.6 g L-1)

in the presence of MB (600 mg L-1) and TEA (68 g L-1), Io = 3.0 ×1016 photon/s, P:

peak area of each congener from GC, Po: the peak area of initial PCB 138.

0%

15%

30%

45%

P/Po

CTAB TWEEN80

Page 96: 1.3 Light emitting diode

77

Figure 3-6: The product distribution (after six minutes irradiation) for

dechlorination of PCB 138 solubilized by TWEEN 80 (1.6 g L-1) or CTAB (1.6 g L-1)

in the presence of MB (600 mg L-1) and NaBH4 (10 g L-1); Io = 3.0 ×1016 photon s-1, P:

peak area of each congener from GC, Po: the peak area of initial PCB 138.

3.4 Conclusions

There are two major conclusions that are drawn from this study: Firstly, it is shown that

PCBs can be dechlorinated in an aqueous medium using sensitized visible light. This

opens an opportunity to exploit sunlight for dechlorinating PCBs, thus significantly

lowering the energy costs. Of course in order to dissolve the PCBs certain surfactants are

necessary. Secondly, amongst the different kinds of surfactants, the non ionic (TWEEN

0%

15%

30%

45%

P/Po

CTAB TWEEN80

Page 97: 1.3 Light emitting diode

78

80) and the cationic (CTAB) surfactants work better than the anionic surfactant (SDS) for

dechlorination, even though the cationic surfactant is not preferred for PCB extraction

from soil. The concentration of the surfactant plays a role in the rate of dechlorination.

The results provides promise to develop a practical method to dechlorinate PCBs in

aqueous solution using surfactants and sensitized visible light.

Page 98: 1.3 Light emitting diode

79

Chapter Four: LED-BASED PHOTOCATALYTIC TREATMENT OF PESTICIDES AND CHLOROPHENOLS

4.1 Introduction

Extensive and sometimes excessive use of pesticides has led to surface and ground water

pollution. Whereas some pesticides may degrade in the environment, others may persist

and pose an ecological risk. This paper focuses on those pesticides that are commonly

used in North America, such as 2-Methyl-4-chlorophenoxyacetic acid (MCPA), 2,4-

Dichlorophenoxyacetic acid (2,4-D), 4-chlorophenol (4-CP) and 2,4-dichlorophenol (2,4-

DCP). Although these are considered to be less harmful than some other environmental

pollutants, their continued use may have long-term consequences. 2-Methyl-4-

chlorophenoxyacetic acid and 2,4-D rank third and fourth, respectively, in the amount of

pesticide used in Canada (Brimble et al., 2005). Chlorophenols are not only widely used

in pesticides, but they are also formed upon chlorination (during disinfection of water and

wastewater) of humic matter (Exon, 1984, Health Canada, 2008). Health Canada

recommends a maximum concentration of 0.1, 0.1 and 0.9 mg L-1, respectively, for 2,4-D,

MCPA and 2,4-DCP in drinking water (Health Canada, 2012).

Advanced oxidative processes (AOPs) have been shown to be successful in degrading

organic contaminants such as pesticides in water (Parsons, 2004, Andreozzi et al., 1999).

Whereas AOPs have a wide range of applications, our focus is to use TiO2-based

photocatalysis, which is known to be broadly applicable. TiO2 has low toxicity, is

biologically and chemically stable, and is economical (Bhatkhande et al., 2002,

Page 99: 1.3 Light emitting diode

80

Linsebigler et al., 1995, Hoffmann et al., 1995). The bandgap energy of TiO2 varies

from 3.0-3.2 eV based on its structure and size. Commercially used TiO2 powder (P25)

has a bandgap energy of 3.2 eV, equal to the energy of photons with wavelength of 385

nm. Photons with this energy or higher can promote electrons in the valence band of such

TiO2 to its conduction band, leaving a positive hole in the valence band (Fox and Dulay,

1993). An electron scavenger, such as oxygen, can capture the electron from the

conduction band and form a superoxide radical ion, whereas the remaining hole is able to

oxidize most organic molecules or oxidize H2O to surface hydroxyl radicals. The holes

and hydroxyl radicals in addition to the superoxide oxygen radicals are reactive species

and can initiate the degradation of pesticides.

Use of mercury discharge lamps to conduct irradiation is the conventional approach in

TiO2 photocatalysis. However, energy costs and lamp life are factors limiting

applications in photocatalysis. In addition, after their service life, the mercury in these

lamps poses an environmental hazard. A light emitted diode (LED), a recent and novel

light source, has a long lifespan, high energy efficiency and small size. LEDs are also

mercury-free and cost effective. These advantages render it an attractive light source for

investigating AOPs. So far, there are only a few papers on LED photocatalysis applied to

the field of environmental engineering. The combination of visible LED and

photocatalysts has been used to treat chlorophenol (Ghosh et al., 2008) and inactivate E.

coli (Chen et al., 2011) in aqueous media. Air purification using ultraviolet (UV) LED

photocatalysis has been reported in several papers (Huang et al., 2009, Shie et al., 2008,

Page 100: 1.3 Light emitting diode

81

Chen et al., 2005). Wang and Ku (2006) successfully used UV LED photocatalysis to

degrade dyes in water sample. To the best of our knowledge, no research focuses on

photocatalytic treatment of pesticides with UV-LED. In this paper, an investigation on

TiO2-based photocatalytic degradation of four pesticides in an LED photoreactor is

reported. Further, a comparison is made between the efficiency of degradation with UV

LED lamps and mercury discharge lamps (black lamps).

4.2 Methods and Materials

4.2.1 Photoreactor

An LED-based photoreactor fitted with UV LED lamps (λmax = 365 nm, full width at half

maximum = 15 nm, NSHU551B) was designed and fabricated. The LED lamps were

procured from Nichia Corporation (Japan). The outer body of the batch reactor is made of

PVC, whereas a reflective material is coated on the inside to minimize loss of light.

Ninety UV LED lamps are arranged in 15 rows with each row having six lamps (Figure

4-1). Every six lamps in series are driven by an LED driver (LT3465). Series connection

of the LEDs can provide identical currents and eliminate the need for ballast resistors. To

minimize the rise of temperature, a small fan was fixed on the wall of the reactor. In

addition, a specific insert containing 15 holes was fabricated, which was used between

the lamps and the sample to partially block the light and consequently vary the intensity.

The number of holes of the insert, through which the light passed, was varied by covering

different number of holes with black tapes. The more holes were covered with black

Page 101: 1.3 Light emitting diode

82

tapes, the lower light intensity was obtained in the center of the insert. As the sample was

continuously stirred, the sample illumination was uniform.

For the irradiation experiments using mercury discharge lamps, a standard Rayonet

photoreactor equipped with 2 black lamps (λmax = 350 nm, full width at half maximum =

50 nm, Hitachi FL8BL-B) was used.

Figure 4-1: LED photoreactor and insert.

Page 102: 1.3 Light emitting diode

83

4.2.2 Chemicals

Ninety-eight percent pure MCPA, 99% pure 4-CP, 99% pure 2,4-D and 99% pure 2,4-

DCP, and 98% pure formic acid were procured from Sigma Aldrich; TiO2 (P25) was

bought from Degussa; and high-pressure liquid chromatography (HPLC)-grade

acetonitrile was obtained from VWR. All of the chemicals were used as received. Milli-Q

water was used in the experiments.

4.2.3 Photocatalytic degradation

Pesticide solutions containing either one pesticide or a pesticide mixture were prepared.

To prepare 20 mg L-1 of a single pesticide solution, 10 mg of its pure product was

dissolved in 500 ml of water using ultrasonication. For pesticide mixtures, 10 mg of each

pesticide was dissolved in 500 ml of water using ultrasonication. The solutions were kept

in dark and stored in a refrigerator. The irradiation experiments were conducted in a

small Pyrex glass vessel [an inner diameter of 2.8 cm], in which 20 ml of the solution was

placed. Different experiments were conducted by adding to the solution varying amounts

of TiO2 photocatalyst. The amounts of TiO2 for different experiments can be founded in

the caption of Figure 4-2 to Figure 4-6. Prior to irradiation, the solution that contained

photocatalyst was stirred for 30 minutes in the dark to ensure that the adsorption of the

pesticide onto the surface of photocatalyst reached equilibrium. During irradiation, a

magnetic stirrer was used to homogenize the solution, and 1-mL samples after different

irradiation periods were collected in a 1.5 ml centrifuge vial.

Page 103: 1.3 Light emitting diode

84

Experiments were conducted using both the LED photoreactor and the Rayonet

photoreactor. Each experiment was conducted in duplicate and results were reported as

an error bar along with the average.

4.2.4 Actinometric Experiment

The amount of radiation entering the reaction vessel was determined using ferrioxalate

actinometry which can monitor wavelengths from 254 nm to 500 nm (Calvert and Pitts,

1966). Table 4-1 shows the varying light intensities measured inside a 20 ml reaction

vessel for the LED and Rayonet photoreactor.

Table 4-1: Light intensity of different photoreactors.

Type of photoreactor Light intensity (×1016 photon s-1)

LED photoreactor 8.55

LED photoreactor with an insert 3.95

LED photoreactor with an insert, 10 holes covered 1.20

LED photoreactor with an insert, 14 holes covered 0.49

Rayonet photoreactor with two black lamps 8.25

Page 104: 1.3 Light emitting diode

85

4.2.5 Analysis of sample

4.2.5.1 HPLC Analysis

All collected samples were initially centrifuged (5,000 resolution min-1) for 5 minutes

with a Fisher Scientific Micro Centrifuge (Model 59A); the supernatant was then passed

through a 0.22 µm filter (Micro Separation) to remove all TiO2 particles. The filtrate was

then stored in a 2-ml amber glass vial in the refrigerator. 2,4-Dichlorophenoxyacetic acid,

MCPA, 2,4-DCP and 4-CP were identified and quantified using a Varian Prostar 210

HPLC instrument equipped with a 325-liquid chromatrography (LC) UV-Visible

detector. A Kinetex pentafluorophenyl (PFP) column (2.6-µM 100 Å) was used to

separate the parent compound and its byproducts. A 20 µL sample was injected and

isocratic elution with a 1.0 ml min-1 flow rate was used in analysis; the eluent was

comprised of 50% acetonitrile (0.1% w/v formic acid) and 50% water (0.1% w/v formic

acid). The UV-Visible detector wavelength was set at 280 nm and the temperature was

controlled at 25 oC. Identification of each compound was achieved by comparing their

retention time with known commercially procured standards. Each compound was

quantified using an external standard. For each compound, a calibration curve was

prepared by using seven different concentrations of standard solutions. The detection

limit for each pesticide is 0.1 mg L-1.

4.2.5.2 TOC Analysis:

Twenty milliliters of samples at different irradiation time were centrifuged at 2, 000

revolution min-1 for 10 minutes with a Centaur 2 centrifuge (Fisons). The supernatant was

Page 105: 1.3 Light emitting diode

86

then passed through a syringe filter of 0.44 µm pore size (Whatman Inc), and analyzed

using an Apollo 9000 Combustion total organic carbon (TOC) analyzer equipped with an

autosampler. The detection limit is 0.1 mg L-1.

4.3 Results and Discussions

4.3.1 Photocatalytic degradation of pesticides and chlorophenols

In the UV-Visible spectrum range from 250-700 nm, 2,4-D, in addition to the other three

pesticides (MCPA, 4-CP, and 2,4-DCP), have single peaks with a maximum at around

280 nm. None of them has absorption at either 350 nm (peak radiation for the mercury

discharge black lamps) or 365 nm (peak wavelength for the LEDs). Our control

experiments also showed that no degradation occurred by direct irradiation without TiO2

photocatalyst. Consequently, direct photolysis does not play a role in the degradation of

these chemicals. Any photodegradation that may occur is a consequence of TiO2

photocatalysis.

Figure 4-2 (a) plots the change in concentration of the four pesticides with 2 g L-1 TiO2

under UV-LED irradiation against the energy dosage. All the data reported here are

reliable and the errors for all duplicates are within 5%. Energy dosage, based on the

number of photons entering the solution, provides a scalable parameter and aids a

comparison between the mercury discharge lamps and LEDs. From a kinetic perspective,

it is linear with time and hence a surrogate for time is recognizing reaction orders.

Page 106: 1.3 Light emitting diode

87

Figure 4-2: Photocatalytic degradation of different pesticides with UV-LED

photoreactor (Io = 8.55×1016 photon s-1, CTiO2=2.0 g L-1, Co=20 mg L-1): (a) loss of

parent pesticides; and (b) loss of total organic carbon.

0.0

0.2

0.4

0.6

0.8

1.0

1.2

0 0.05 0.1 0.15 0.2

C/C

o-Pe

stic

ides

Energy dosage (kJ)

(a) MCPA 2,4-D

2,4-DCP 4-CP

0.0

0.2

0.4

0.6

0.8

1.0

1.2

0 0.1 0.2 0.3 0.4 0.5

C/C

o-TO

C

Energy dosage (kJ)

(b)

MCPA 2,4-D

2,4-DCP 4-CP

Page 107: 1.3 Light emitting diode

88

The results, obtained using a HPLC, show that all four compounds can be catalytically

degraded in the LED reactor. For MCPA, with 0.014 kJ energy dosage, 90% was

removed from solution, and when the energy dosage reached 0.041 kJ more than 99%

was degraded. More than 70% of 2,4-D degraded within 0.014 kJ dosage and it became

undetectable with a 0.041 kJ. Degradation of chlorophenols was slower than those of 2,4-

D and MCPA. To get about 50% degradation of 2,4-DCP and 4-CP, a 0.055 kJ energy

dosage was needed. When the energy dosage increased to 0.11 kJ, they achieved more

than 99% degradation. The first order rate coefficients (Table 4-2) showed that

photocatalytic degradation of MCPA is 10× faster than that of 4-CP. The photocatalytic

degradation of these compounds is caused by the attack of generated surface holes and/or

hydroxyl radicals. Possible degradation pathways attributable to hydroxyl radicals

attacking can be found in several papers (Theurich et al., 1996, Topalov et al., 2001,

Kwan and Chu, 2004). The major pathways of photocatalytic degradation of phenoxy

pesticides are through homolysis of carbon-oxygen bond on its aromatic ring. This step is

very fast and leads to higher disappearance rates for 2,4-D and MCPA than

chlorophenols. In adsorption experiment, respectively, 17%, 15%, 3% and 3% of MCPA,

2,4-D, 4-CP and 2,4-DCP adsorbed on the surface of TiO2 after 30 minutes dark. The

higher percentage of adsorption of MCPA and 2,4-D also contribute to higher

degradation rates based on Langmuir-Hinshelwood mechanism (Fox and Dulay, 1993).

Analysis of intermediates was not conducted, but measurement of TOC at different times

identifies the extent of mineralisation. Figure 4-2 (b) shows the TOC results for different

pesticides. For MCPA, 2,4-D, 4-CP and 2,4-DCP to be completely (not detectable)

mineralized, one needs less than 0.5 kJ energy. Complete mineralization shows that the

Page 108: 1.3 Light emitting diode

89

intermediates produced from the degradation of these four pesticides are not recalcitrant

and can be mineralised to carbon dioxide and other inorganic compounds.

Table 4-2: First order rate coefficients (K) for photocatalytic pegradation of

different pesticides.

Pesticide k (× 10-3 s-1)

MCPA 6.02

2,4-D 4.01

4-CP 0.63

2,4-DCP 0.52

Note: Experiments with 2 g L-1 TiO2 and a light intensity of 8.55×1016 photon s-1.

4.3.2 Photocatalytic degradation of pesticides mixtures

Most contaminated water contains a mixture of different compounds. To investigate how

pesticides compete with each other, three mixtures of pesticides were studied. Table 4-3

shows the composition of the mixtures.

Table 4-3: Mixtures of Pesticides.

Mixture Composition

A 20 mg L-1 of 4-CP, 20 mg L-1 of 2,4-DCP

B 20 mg L-1 of 2,4-D, 20 mg L-1 of 4-CP

C 20 mg L-1 of 2,4-D, 20 mg L-1 of 2,4-DCP

Page 109: 1.3 Light emitting diode

90

Figure 4-3: Photocatalytic degradation of pesticides mixture with UV-LED

photoreactor based on the loss of pesticides detected by HPLC (Io =8.55×1016 photon

s-1, CTiO2=2 g L-1, Co=20 mg L-1): (a) mixture containing 4-CP and 2,4-DCP; (b)

mixture containing 4-CP and 2,4-D; (c) mixture containing 2,4-DCP and 2,4-D.

0.0

0.5

1.0

1.5

0 0.03 0.06 0.09 0.12 0.15 0.18

C/C

o

Energy dosage (kJ)

(a) 4-CP 2,4-DCP 4-CP (mixture) 2,4-DCP (mixture)

0.0

0.5

1.0

1.5

0 0.03 0.06 0.09 0.12 0.15 0.18

C/C

o

Energy dosage (kJ)

(b) 4-CP 2,4-D 4-CP (mixture) 2,4-D (mixture)

0.0

0.5

1.0

1.5

0 0.03 0.06 0.09 0.12 0.15 0.18

C/C

o

Energy dosage (kJ)

(c) 2,4-D 2,4-DCP 2,4-D (mixture) 2,4-DCP (mixture)

Page 110: 1.3 Light emitting diode

91

Figure 4-3 presents the results for the photocatalytic degradation of pesticide mixtures. In

each case, the rate of degradation of a pesticide in the mixture was slower than that in

solution containing a single pesticide. To compare the mixture and solution containing

single pesticide quantitatively for each case, the percentage removal of pesticides at the

same energy dosage (0.028 kJ) for each case was summarized in Table 4-4. For Mixture

A, it was found that only 11% removal for 4-CP and 19% removal for 2,4-DCP were

achieved with 0.028 kJ energy dosage, however, more than 25% removal for 4-CP and

28% removal for 2,4-DCP was attained with this energy dosage when a single pesticide

solution was used. In Mixture B, the percentage removal of 2,4-D was similar to its single

pesticide solution with a 0.028 kJ energy dosage; however, the percentage removal of

Table 4-4: Percentage removal of pesticides at 0.028 kJ energy dosage.

Case

Percentage removal

2,4-D 4-CP 2,4-DCP

Single pesticides 90 25 28

Mixture A _ 11 19

Mixture B 88 11 _

Mixture C 72 _ -24a

Note: Energy dosage =0.028 kJ.

ain the mixture C, the concentration of 2,4-DCP is increased by 24% at the energy dosage

of 0.028 kJ, reflecting production from 2,4-D.

Page 111: 1.3 Light emitting diode

92

4-CP significantly decreased. In Mixture C, the removal rate of 2,4-D is apparently

smaller compared with its single pesticide solution with a 0.028 kJ energy dosage, and

the concentration of 2,4-DCP increased at this energy dosage as it was produced upon the

degradation of 2,4D. In Mixture C, the concentration of 2,4-DCP initially increased and

then decreased after reaching a maximum.

The previous results can possibly be explained by the competition between pesticides for

photons, adsorption sites, and the holes or hydroxyl radicals on the surfaces. In our case,

all of these pesticides neither absorb nor directly use a photon of wavelength of 350 or

365 nm; consequently, the competition for the photons can be neglected in this paper. The

results of pesticides adsorption on the surface of TiO2 after 30 minutes in the dark (Table

4-5) showed that the additional pesticides do not affect the adsorption of previous

existing pesticides on surface of TiO2 in the solution, which indicates 2 g L-1 TiO2 can

provide enough adsorption sites for pesticides in our experimental conditions and the

competition for the adsorption sites is trivial. Therefore, the competition for hydroxyl

radicals/holes may be the real reason causing the slower degradation rate in mixtures. In

mixture A, 4-CP and 2,4-DCP have similar properties, thus the competing effect for

hydroxyl radicals or holes is similar and their corresponding degradation rates were

significantly affected by each other. In mixture B, 2,4-D is a superior competitor than 4-

CP for harvesting hydroxyl radicals, because the carbon-oxygen bond on the aromatic

ring of 2,4-D is more easily broken down by hydroxyl radical. Thus, degradation of 2,4-D

is not affected too much but the degradation of 4-CP is significantly retarded. In mixture

Page 112: 1.3 Light emitting diode

93

C, hydroxyl radicals also favoured attacking 2,4-D, and lead to an accumulation of 2,4-

DCP at beginning and then decreased because of attacking by hydroxyl radicals. In this

case, the accumulation of 2,4-DCP can inversely retard the degradation of 2,4-D to form

2,4-DCP.

Table 4-5: Percentage of pesticides adsorbed on the surface of TiO2 after 30 minutes

of stirring in the dark.

Case

Percentage adsorption

2,4-D 4-CP 2,4-DCP

Single pesticides 15 3 3

Mixture A / 3 3

Mixture B 14 3 /

Mixture C 14 / 3

4.3.3 Effect of Photocatalyst Loading

Five different TiO2 loadings (0.2, 0.5, 1.0, 2.0 and 3.0 g L-1) were investigated for the

photocatalytic degradation of 2,4-D. Figure 4-4(a) shows the result of 2,4-D

concentration against irradiation time. The photocatalytic degradation of 2,4-D in our

system follows approximate first-order kinetics. Figure 4-4(b) plots the first-order rate

constants fitted over the first 15 minutes. The rate constants increased with TiO2 loading

Page 113: 1.3 Light emitting diode

94

(a)

(b)

Figure 4-4: Photocatalytic degradation of 2,4-D with different TiO2 loadings and

LED irradiation (Io =8.55×1016 photon s-1, Co=20 mg L-1).

appearing to approach a limiting value with increased TiO2 loadings, as expected for

light-limiting conditions (Augugliaro et al., 1988). The light-limiting condition is

0.0

0.2

0.4

0.6

0.8

1.0

1.2

0 10 20 30 40 50 60

C/C

o

Irradiation time (min)

[TiO₂]=0.2 g/L

[TiO₂]=0.5 g/L

[TiO₂]=1.0 g/L

[TiO₂]=2.0 g/L

[TiO₂]=3.0 g/L

0.00

0.05

0.10

0.15

0.20

0.25

0.30

0 0.5 1 1.5 2 2.5 3 3.5

r (m

in-1

)

[TiO2] (g L-1)

Page 114: 1.3 Light emitting diode

95

confirmed by the dependence of first-order rate on light intensity. An increase of TiO2

loading at lower levels can significantly improve the photocatalytic degradation rate.

When the TiO2 loading increased to 0.5 g L-1 from 0.2 g L-1, the rate constant doubled;

however, at a high level, increasing TiO2 loading does not contribute too much to the

improvement of photocatalytic degradation rate. An increase in TiO2 loading from 2.0 g

L-1 to 3.0 g L-1 only led to a 5% enhancement in the photocatalytic degradation rate.

Based on this result, a suitable loading of TiO2 was determined to be 2.0 g L-1, which had

a kinetic rate constant of 0.24 min-1.

As chemical reactions occur on the surface of TiO2, the efficiency of photocatalytic

degradation depends on the surface of TiO2 particles, which can simultaneously be in

contact with the target contaminant and illuminated by light. At low TiO2 loadings and

unchanged light intensity the active absorbing surface area is a limiting factor, thus better

degradation efficiency is observed as TiO2 loading increased. Nevertheless, with

increased TiO2 loadings, the light available is utilized efficiently and light becomes the

limiting factor.

4.3.4 Effect of Light Intensity

Investigations were conducted with varying light intensities. With the optimal TiO2

loading of 2.0 g L-1, four different light intensities (4.9×1015, 1.28×1016, 3.95×1016 and

8.55×1016 photon s-1) were investigated in 2,4-D photocatalytic degradation. Figure 4-5

shows the effect of different light intensities on 2,4-D degradation. The smallest first

Page 115: 1.3 Light emitting diode

96

(a)

(b)

Figure 4-5: Photocatalytic degradation of 2,4-D with UV-LED photoreactor under

different light conditions (Co=20 mg L-1, CTiO2=2 g L-1).

order rate constant was 0.0091 min-1 at a light intensity of 4.9×1015 photon s-1, and the

largest rate constant was 0.2404 min-1 at light intensity of 8.55×1016 photon s-1. The

0.00

0.20

0.40

0.60

0.80

1.00

1.20

0 10 20 30 40 50 60

C/C

o

Irradiation time (min)

Iₒ=4.90×10¹⁵ photon s⁻¹ Iₒ=1.28×10¹⁶ photon s⁻¹ Iₒ=3.95×10¹⁶ photon s⁻¹ Iₒ=8.55×10¹⁶ photon s⁻¹

0

0.05

0.1

0.15

0.2

0.25

0.3

0 1 2 3 4 5 6 7 8 9

r(m

in-1

)

Io (*1016 photon s-1)

Page 116: 1.3 Light emitting diode

97

percentage of 2,4-D eliminated was dependent on the light intensity, which is

proportional to the intensity of irradiation. Based on the previous results, the applied

TiO2 loading (2.0 g L-1) reported in this paper is not a limiting factor for the light

intensity investigated. Therefore, it was the light intensity that causes the different

degradation kinetics reported in this paper. Ollis et al. (1991) summarized the relationship

between light intensity and kinetics rate as follows: (1) at low light intensity electron-hole

formation dominates and rate is proportional to the light intensity; (2) as light intensity

increased, electron-hole pair generation competes with electron-hole recombination and

the kinetic rate is proportional to the square root of light intensity; (3) at an even higher

light intensity, TiO2 loading may become a limiting factor. The behaviour reported in this

paper indicates that the applied light intensity is relative low, so electron-hole formation

dominated in the photocatalytic processes.

4.3.5 Comparison between LED and Mercury Lamp Irradiation

Figure 4-6 shows the degradation of 2,4-D in both LED and the Rayonet photoreactor in

terms of concentration versus energy dosage. Photocatalytic degradation of 2,4-D with

LEDs as the light source has a higher energy-efficiency than that with mercury lamps. To

compare the efficiency of LED lamps and mercury lamps, the concept of photon energy

per order (PEPO) was developed and used. PEPO refers to the amount of photonic energy

required to reduce an order of magnitude of contaminant concentration. PEPO is 0.028 kJ

per order for the LED irradiation while it increased to 0.038 kJ per order for mercury

lamps. To understand this phenomenon, two possible hypotheses are provided. First,

photons with longer wavelengths have smaller energies; consequently, for the same

Page 117: 1.3 Light emitting diode

98

energy dosage, more photons enter the reaction vessel in the LED (365 nm) reactor than

the Rayonet reactor equipped with black lamps (350 nm). For example, a 1 J energy

dosage is equal to the total energy of 1.83×1018 photons at a wavelength of 365 nm, or

1.76×1018 photons at a wavelength of 350 nm, accounting for 4% excess photons for the

LED. Second, black lamps also emit more photons of longer wavelengths, which cannot

excite TiO2 and hence become useless. Our calculations based on the emission spectra of

Figure 4-6: Photocatalytic degradation of 2,4-D in the two photoreactors: Co=20 mg

L-1, CTiO2=2 g L-1. (a): LED reactor, Io =8.55×1016 photon s-1; (b): Rayonet reactor,

Io =8.25×1016 photon s-1.

lamps, show that, in the LED lamps, 99.9% of the energy is available for the reaction;

however, a smaller number (95.8%) was obtained for mercury lamps (see Appendix D),

0.00

0.20

0.40

0.60

0.80

1.00

1.20

0 0.02 0.04 0.06 0.08 0.1

C/C

o

Energy dosage (kJ)

Mercury lamps LED

Page 118: 1.3 Light emitting diode

99

another 4% advantage for LEDs. These approximate calculations can account, at least

qualitatively, for the 8% advantage of the LEDs. This shows the advantage of LEDs,

which will be even larger when compared to solar irradiation where the spectrum is broad

and limited in the ultraviolet B range at the Earth’s surface. The essential fact here is that

LEDs can be rendered monochromatic close to the band edge of the photocatalyt so that

photon loss does not occur from non-absorbed wavelengths, nor does energy above the

band edge get wasted as heat.

4.4 Conclusions

In this research, LEDs were shown to be a promising light source in photocatalytic

treatment of pesticides. For all four pesticides investigated, more than 99% degradation

was achieved within a short period of irradiation with 2.0 g L-1 of TiO2 and UV LED

(365 nm) irradiation. The degradation of pesticides in the mixture is slower than when

only one pesticide is investigated because of the competition for surface hydroxyl

radicals between different compounds. When this applied light intensity is 8.55×1016

photon/s, a suitable loading of TiO2 for 2,4-dichlorophenoxyacetic acid degradation was

determined to be 2.0 g L-1. The rate constant at this loading was 0.2404 min-1. The

relationship between light intensity and first order kinetic constants was linear.

Furthermore, the comparison between mercury lamps and LEDs show that LEDs can be

more energy-efficient, and the emission spectrum of LED lamps can be well-matched

with the absorption band of TiO2. If a solar reactor is considered as a competitor, the

photon energy-use advantage of LEDs is even greater.

Page 119: 1.3 Light emitting diode

100

Chapter Five: DESIGN A HOMOGENEOUS RADIATION FIELD IN A UV-LED BASED PHOTOCATALYTIC REACTOR

5.1 Introduction

TiO2 photocatalysis has been widely investigated for its use in water and wastewater

treatment and in air purification (Peral and Ollis, 1992, Hoffmann et al., 1995, Fujishima

et al., 2000). It is based on the principle that ultraviolet (UV) light when incident on the

photocatalyst (TiO2) leads to the generation of strong oxidants such as holes or hydroxyl

radicals. These oxidants are capable of oxidizing most organic compounds and eventually

leading to their mineralisation. Conventional UV light source for photocatalysis use

mercury lamps. With the advent of light emitting diodes (LEDs), they are increasingly

being considered as an attractive alternative to mercury lamps. The advantages of LEDs

include a longer life span, smaller size, higher energy efficiency where technology is

mature, and being mercury free. As LED technology advances further, it will become

more cost effective. The application of UV-LED in photocatalytic treatment of water and

wastewater is still novel and has only been reported by a few researchers (Natarajan et

al., 2011b, Wang and Ku, 2006, Yu et al., 2013).

As cost effectiveness these systems depend on the efficient use of light, an optimized

radiation field such that maximum photons are harvested, is necessary. Some researchers

(Imoberdorf et al., 2008b, Jacob and Dranoff, 1969, Jacobm and Dranoff, 1970, Irazoqui

et al., 1973, Alfano et al., 1986a) have described various radiation field models generated

by conventional light source both in homogenous and heterogeneous environments

Page 120: 1.3 Light emitting diode

101

involving different geometries. Wang et al (2012) studied the radiation field model

generated by the UV-LED array and showed that the homogeneity of the radiation is

affected by the distances between the UV-LED array and photocatalyst plate. However,

no discussion on how to develop an optimized homogenous radiation field was presented.

In this paper, a radiation field model is developed for a UV-LED array in a planar

photoreactor with the intent to use it for designing a homogenous radiation field such as

that photon harvesting can be maximized.

5.2 Advantage of homogeneous radiation field in a photocatalytic reactor

Several papers (Mehrotra et al., 2005, Ollis et al., 1991, Choi et al., 2000, Kim and Hong,

2002) have investigated the effect of light intensity on photocatatyic kinetics, showing

that first order photocatalytic kinetics rate (K) usually follows power-law dependence on

light intensity (I) ( Equation[5-1] ).

𝐾 = 𝜂𝐼𝛾 [5-1]

Where 𝜂, 𝛾 are constant, 𝜂 is a positive value and 𝛾 has a value between 0 and 1. The

performance of a photocatalytic reactor is determined by the average photocatalytic

kinetics rate (Ka) occurring on the photocatalyst plate (Equation [5-2]).

Ka = ∫KdApAp

= ∫𝜂𝐼𝛾dApAp

[5-2]

Page 121: 1.3 Light emitting diode

102

Where Ap is the area of the photocatalyst plate. Once the photocatalyst plate is

homogenously irradiated, the average kinetics rate become:

Ka = ∫𝜂𝐼𝑎𝛾dApAp

[5-3]

Where 𝐼𝑎 is the average light intensity for a homogenous radiation field.

𝐼𝑎 =∫ 𝐼dAp

Ap [5-4]

It is known that 𝑓(𝐼) = 𝜂𝐼𝛾 is a concave function. Hence, according to Jensen's

inequality,

∫𝜂𝐼𝛾dApAp

≤ ∫𝜂𝐼𝑎𝛾dApAp

[5-5]

Equation [5-5] revealed that the average photocatalytic kinetics rate in a homogeneous

radiation field is larger than that in a non-homogenous radiation field. Therefore, a

homogenous radiation field is superior for an ideal photocatalytic reactor. Besides, a

homogeneous radiation field can simplify the photocatalytic process model and require

less computation.

Page 122: 1.3 Light emitting diode

103

5.3 Development of radiation field model

5.3.1 UV-LED array and photocatalyst plate

Figure 5-1 provides the geometry of a UV-LED array and a fixed catalyst surface. The

UV-LED array panel is considered parallel to the photocatalyst plate. The UV-LEDs in

the panel were arranged as a regular array. The distance between UV-LEDs panel and

photocatalyst plate is known as the ''irradiated distance'' (ID). The distance between the

two adjacent LEDs is called as the ''gap''.

Figure 5-1: UV-LED array and photocatalyst plate.

For the experiments to calibrate and validate the model, a panel comprising of 16 UV-

LEDs arranged in a 4 by 4 array was fabricated. UV-LEDs were spaced 2.5 cm and

surface mounted on a 10 cm by 10 cm PCB board. UV-LEDs (λmax=365nm, half width=9

Page 123: 1.3 Light emitting diode

104

nm, NSCU033) were procured from Nichia Corporation. The UV-LEDs were driven by a

high current Quad output LED driver (LT3476). A Pyrex glass plate (thickness=0. 32 cm)

was obtained from Chemglass Life Sciences (Vineland, NJ, US) and used as a shielding

glass. The light transmittance of the glass plate is around 90% at 365 nm.

5.3.2 Radiation field model without shielding glass plate

In most photocatalytic reactors, the UV light source is protected by a glass shield (Figure

5-1.c), so the impact of shielding glass on radiation field model needs to be considered. In

this first instance, a radiation field model without shielding glass (Figure 5-1.b) is

developed. The optical effects such as scattering, reflection and refraction were assumed

to be negligible. All UV-LED lamps were assumed to be identical and considered to be

point light sources with a special directivity. The radiation field model was developed

using a similar method described in Wang et al (2012). Initially, a radiation field model

produced by a single UV-LED was established. Then the radiation field generated by

multiple UV-LEDs was developed by considering the sum of the contribution of each

UV-LED. A stepwise description of the development of the radiation field is given

below:

(1) Obtain the radiation directivity function [Re(θ)] for a single UV-LED. The radiation

directivity function (Equation [5-6]) is the ratio between light intensity [I(θ)] with view

angle θ and the light intensity with a zero view angle. View angle (θ) is defined as the

angle between the radiation direction and the direction perpendicular to the UV-LED.

Page 124: 1.3 Light emitting diode

105

𝑅𝑒(𝛉) = 𝐼(𝛉)𝐼(0)

[5-6]

For most UV-LEDs, the light emitted from UV-LED varies with the radiation directivity.

Figure 5-2 shows the radiation directivity of our UV-LED. The light intensity emitted

from zero view angle has the maximum value, and then it gradually decrease as the view

angle increases. The light intensity approach zero when the view angle is 70o or 1.22

radian. In this study a quadratic equation (Equation [5-7]) was used to simulate the

radiation directivity within 1.22 radian.

𝑅𝑒(θ) = 1 − 0.67θ2 [5-7]

When 1.22 ≤ θ < π/2, 𝑅𝑒(θ)=0, so we have:

𝑅𝑒(θ) = �1 − 0.67θ2, 0 < θ < 1.220, 1.22 ≤ θ < π/2

� [5-8]

Figure 5-2: Directivity of radiation (NICHIA, 2013).

Page 125: 1.3 Light emitting diode

106

(2) Determine the relationship between the light intensities at different radial distances

when view angle is the same. When the view angle is constant, the light intensity is

proportional to the inverse of the square of distance (Equation [5-9]).

𝐼 ∝ 1R2

[5-9]

(3) Develop a function between the light intensity [I(θ,R)] at any position and the

referenced light intensity [I(0,do)]. The referenced light intensity is the light intensity at a

specific distance (do) and with a zero view angle. Based on Equation [5-6] and [5-9], I(θ,

R) can be expressed as a function of I(0,do) as:

𝐼(θ, R) = 𝐼(θ, do) ∗do

2

R2 = 𝐼(0, do) ∗ 𝑅𝑒(θ) ∗do

2

R2 [5-10]

Considering the normal flux density, then:

𝐼𝑛(θ, R) = 𝐼(θ, R) ∗ cos(θ) = 𝐼(0, do) ∗ 𝑅𝑒(θ) ∗do

2

R2 ∗ cos(θ) [5-11]

(4) Convert the polar coordinates in Equation [5-11] to Cartesian coordinates. The

Cartesian coordinate system and the polar coordinate system are shown in Figure 5-3,

where the x-y plane is a photocatalyst plate.

Page 126: 1.3 Light emitting diode

107

Figure 5-3: Cartesian and polar coordinates in radiation system.

cos (θ) = dR [5-12]

R2 = d2 + g2 = d2 + (x − xo)2 + (y − yo)2 [5-13]

θ = arctan (�(x−xo)2+(y−yo)2

d) [5-14]

Incorporating Equations [5-12], [5-13] and [5-14] into Equation [5-11], the normal light

intensity at any point on a photocatalyst plate can be expressed as a function of d and its

x-y coordinates as:

Page 127: 1.3 Light emitting diode

108

In(x, y, d) = 𝐼(0, do) ∗ Re�arctan (�(x− xo)2 + (y − yo)2

d) �

∗do

2 ∗ d(d2 + (x − xo)2 + (y − yo)2)3/2

[5-15]

Therefore, the normal light intensity at any point due to the contribution of multiple LED

lamps is:

𝐼𝑠𝑢𝑚(x, y, d) = � Ini(x, y, d)n

i=1

= � (𝐼(0, do) ∗ Re�arctan��(x − xi)2 + (y − yi)2

d��

n

i=1

∗do

2 ∗ d

(d2 + (x − xi)2 + (y − yi)2)32

)

[5-16]

(5) Measure and estimate the referenced light intensity. The referenced light intensity

can be measured by UV radiometer or be estimated based on its relationship with the

light output of UV-LED (It) shown as Equations [5-17] ~ [5-20].

The light output of a UV-LED is the integral of radiation over the entire view angles:

It = � I(θ)dsθo

0 [5-17]

The following relationships can be obtained from Figure 5-4.

ds ≅ 2πr ∗ dl = 2πr ∗ R ∗ dθ = 2π ∗ R ∗ sinθ ∗ R ∗ dθ = 2πR2 ∗ sinθ ∗ dθ [5-18]

Page 128: 1.3 Light emitting diode

109

Substitute Equation [5-6] and Equation [5-18] into Equation [5-17]:

It = I0 ∗ 2πR2 ∫ Re(θ) ∗ sinθdθθo0 [5-19]

I0 =It

2πR2 ∫ Re(θ) ∗ sinθdθθo0

[5-20]

Figure 5-4: Scheme of UV-LED radiation.

5.3.3 Radiation field model with a shielding glass plate

The light intensity is attenuated due to the absorption by the shielding glass. The loss of

light intensity can be expressed by the Beer-Lambert law:

T = 10−β𝑙 [5-21]

Page 129: 1.3 Light emitting diode

110

Where T is the transmittance, 𝑙 is the light passing length in glass plate, and β is the

attenuation coefficient. After considering the light absorption by shielding glass plate, the

modified directivity function will be:

𝑅𝑒′(𝜃) = 𝑅𝑒(𝜃) ∗ T(𝜃)T(0) = 𝑅𝑒(𝜃) ∗ 10

−β𝑙𝑔𝑐𝑜𝑠 (𝜃)

10−β𝑙𝑔= 𝑅𝑒(𝜃) ∗ 10𝑙𝑔∗β∗(1− 1

𝑐𝑜𝑠 (𝜃)) [5-22]

Where T(𝜃) is the ratio of light passing through the glass plate with a view angle ( 𝜃), 𝑙𝑔

is the thickness of glass plate. Therefore, the modified radiation field model becomes:

𝐼𝑠𝑢𝑚(x, y, d) = � (𝐼(0, do) ∗ Re�arctan��(x − xi)2 + (y − yi)2

d��

n

i=1

∗do

2 ∗ d

(d2 + (x − xi)2 + (y − yi)2)32

)

[5-23]

Where d > 𝑙𝑔.

5.4 Calibration and validation of the radiation field model

For the purpose of model calibration, the light intensity at a distance of 1cm

perpendicular to a single UV-LED was used as referenced light intensity. The model was

validated using the light intensity generated by the UV-LED array at different distances

and locations.

5.4.1 Light intensity measurement

The light intensity was measured using a Silver Line UV Radiometer (M007153, Geneq

Inc. Canada). The measurements range from 0-2000 mW cm-2. The sensor of the

Page 130: 1.3 Light emitting diode

111

radiometer has a radius of 0.25 cm. The average light intensity received by the sensor

was reported in this UV radiometer reader. The sensor surface was kept parallel to the

UV-LED panel when the light intensity was measured.

The readout of the light intensity from the radiometer is the average light intensity

received by the sensor. A relationship between referenced light intensity and the UV

radiometer measured value need to be developed. The measured value received by the

sensor (Figure 5-5) can be expressed as Equation [5-24]:

Imeasured =∫ Idsπrs2

=∫ I ∗ 2πr ∗ drrs0

πrs2=∫ I ∗ 2r ∗ drrs0

rs2 [5-24]

Where rs is the radius of the sensor, Imeasured is the light intensity read from the UV

radiometer, S is the area of the sensor.

Figure 5-5: Geometry of sensor.

Page 131: 1.3 Light emitting diode

112

According to Equation [5-15] and [5-24], we have:

� {𝐼(0, do) ∗ Re �arctan (r

do) � ∗

do2 ∗ do

(do2 + r2)3/2

∗ 2r ∗ drrs

0= rs2 ∗ Imeasured [5-25]

𝐼(0, do) =rs2 ∗ Imeasured

2do3 ∗ ∫ {Re �arctan ( r

do) � ∗ r

(do2 + r2)3/2 ∗ drrs

0

[5-26]

The average light intensity received by the sensor at 1 cm distance for a single UV-LED

lamp is measured to be 103.9 mW cm-2. Equation [5-26] provides the referenced light

intensity to be 110 mW cm-2.

5.4.2 Model light intensities vs measured light intensities

To validate the model developed above, it was necessary to compare the theoretically

calculated values with measured ones. As the UV radiometer provides the average light

intensity received by the sensor, the average model light intensity received by the sensor

were calculated and compared. The radiation field model without shielding glass and

with shielding glass were respectively calculated using Equation [5-16], Equation [5-23]

and parameters provided in Table 5-1.

Page 132: 1.3 Light emitting diode

113

Table 5-1: Parameters used for radiation field model calculation. Parameters Value

do 1 cm

I(0, do ) 110 mW cm-2

lg 0. 32 cm

θo 1.22 radian

β 0.143 cm-1

The difference between the model value and measured value was used to evaluate the

relative errors as:

𝑅𝑒𝑙𝑎𝑡𝑖𝑣𝑒 𝑒𝑟𝑟𝑜𝑟 = | 𝐼𝑚𝑜𝑑𝑒𝑙−𝐼𝑚𝑒𝑎𝑠𝑢𝑟𝑒𝐼𝑚𝑒𝑎𝑠𝑢𝑟𝑒

| ∗ 100% [5-27]

Results presented in Table 5-2. show that the relative errors are small, which indicates

that the radiation filed model reliably captures the light intensity.

Table 5-2: Comparison of modeled light intensity and measured light intensity.

Position Without Shielding Glass With Shielding Glass

x (m) y (m) d (m) Imodel (mW cm-2)

Imeasure (mW cm-2)

Relative Error

Imodel (mW cm-2)

Imeasure (mW cm-2)

Relative Error

0.072 0.06 0.036 31.9 31.4 1.5% 27.9 26.2 6.6%

0.056 0.06 0.036 34.0 34.4 1.1% 29.8 27.8 7.1%

0.041 0.06 0.036 33.8 33.2 1.9% 29.6 28.6 3.6%

0.027 0.06 0.036 31.5 29.6 6.5% 27.6 28.0 1.3%

0.062 0.06 0.054 25.9 26.0 0.3% 22.9 22.0 4.0%

0.049 0.06 0.054 26.5 26.8 1.0% 23..4 23.2 1.0%

0.033 0.06 0.054 25.2 26.4 4.5% 22.3 22.6 1.4%

Page 133: 1.3 Light emitting diode

114

5.5 Design of a homogenous radiation filed

The concept of Max Error (Me) was extended to evaluate the homogeneity of the

radiation field.

𝑴𝒆 = max [ �Imax−IaIa

� ∗ 100%, �Ia−IminIa

� ∗ 100%] [5-28]

Where 𝐼𝑚𝑎𝑥 is the maximum light intensity, 𝐼𝑚𝑖𝑛 is the minimum light intensity and Ia is

the average light intensity.

To achieve a high degree of homogeneity, the calculated Me should be small. If Me is

zero, the radiation filed is completely homogenous. Here the radiation field is defined to

be homogenous when Me is smaller than 1%.

5.5.1 The effect of ID on the homogeneity of radiation field for a fixed gap

To investigate the impact of ID/gap ratio on the homogeneity of radiation field, a 2 m by

2 m square photocatalyst plate was studied. To ignore the edge effects, the degree of

homogeneity is only calculated within 10 cm by 10 cm square area. Equation [5-16] was

used to determine the radiation field without shielding glass. The gap between two

adjacent lamps is fixed at 2.5 cm, and the Max Errors of radiation field for different

distances were calculated. The radiation field results for different irradiated distance are

shown in Figure 5-6 and the Max Error results are provided in Figure 5-7. The results

show that the degree of homogeneity is low when distance between the UV-LED panel

and photocatalyst plate approaches zero and the degree of homogeneity is increased as

''ID'' increased. However, the light intensity decreases as the ID increases. An ideal

radiation field has a high degree of homogeneity and a minimal loss of light intensity.

Page 134: 1.3 Light emitting diode

115

Therefore, the optimal "ID" is the smallest distance which generates a homogenous

radiation field.

(a)

(b)

Figure 5-6: The radiation field with different ID: (a) ID=0.01 m, gap=0.025 m; (b) ID= 0.04 m, gap=0.025 m.

Page 135: 1.3 Light emitting diode

116

Figure 5-7: The effect of irradiated distance (ID) on Maximum Error.

5.5.2 Optimal combination of ID and gap

To determine the optimal combination of ID and gap, gap is varied from 0.01 m to 0.1

m, and the optimal ID for each gap is investigated using the method described in part

5.5.1. The result (Figure 5-8) shows that the optimal ID is proportional to gap and the

optimal ID is 1.26 times of gap.

Figure 5-8: Optimal combination of ID and gap.

0%

50%

100%

150%

200%

250%

0 0.02 0.04 0.06 0.08 0.1

Me

ID (m)

ID = 1.26 *gap

0

0.03

0.06

0.09

0.12

0.15

0 0.02 0.04 0.06 0.08 0.1

ID (m

)

gap (m)

Page 136: 1.3 Light emitting diode

117

5.5.3 Selection of the output of the UV-LED

In section 5.5.2, the optimal combination of ID and gap has been determined. The next

step is to choose the light output of LED lamps which generate a specific homogenous

radiation field. The selection of the light out of LED can use the following methods:

(1) Specify the dimension of the gap of UV-LEDs array and specify the designed light

intensity received by the photocatalyst plate.

(2) Calculate the irradiated distance based on the optimal ID/gap ratio to achieve a

homogenous radiation field.

(3) Calculate the required referenced light intensity based on the radiation field model.

(4) Calculate the light output of LED lamp based on Equation [5-19].

The required light out of LED for different gaps and different designed light intensity

were shown in Figure 5-9.

Figure 5-9: Selection of light output of UV-LED.

0 500

1000 1500 2000 2500 3000 3500 4000 4500 5000 5500

0 0.02 0.04 0.06 0.08 0.1

Ligh

t out

put p

er la

mp

(mW

)

gap(m)

Designed light intensity = 50 mW/ cm² Designed light intensity = 20 mW/ cm² Designed light intensity = 10 mW/ cm² Designed light intensity = 5 mW/ cm² Designed light intensity = 2 mW/ cm² Designed light intensity = 1 mW/ cm² Designed light intensity = 0.5 mW/ cm²

Page 137: 1.3 Light emitting diode

118

5.6 Conclusions

Radiation field model for a UV-LEDs array has been developed, which can predict the

normal light intensity at any location of a photocatalyst plate with any ID. Based on the

model, the degree of homogeneity of the radiation field is significantly affected by the ID

when the gap is fixed. Homogenous radiation field can be achieved by choosing an

optimal ID/gap ratio. The ratio is found to be 1.26 for Nichia UV-LED (NCSU033).

Besides, the method of selecting the light output of UV-LED for different gaps to achieve

a desired homogenous light intensity was developed and evaluated.

Page 138: 1.3 Light emitting diode

119

Chapter Six: A NOVEL LIGHT EMITTING DIODE BASED PHOTOCATALYTIC REACTOR FOR WATER TREATMENT

6.1 Introduction

In recent decades, "emerging contaminants" such as pharmaceuticals, personal care

products, pesticides have been frequently detected in domestic wastewater and surface

water (Petrovic et al., 2008). Conventional treatment processes do not specifically target

these emerging contaminants, and their presence in aquatic environments has led to

adverse ecological effects. TiO2 based photocatalysis has been shown to be an efficient

method in dealing with these contaminants (Herrmann, 1999, Miranda-García et al.,

2011, Belgiorno et al., 2007). This technology utilizes strong oxidants such as hydroxyl

radicals or holes or reactive oxygen species formed upon electron capture by O2 to

oxidize most organic compounds. Usually this leads to conversion of parent compounds

to harmless compounds and in many cases complete mineralisation.

The successful application of TiO2 photocatalysis in water treatment requires designing

simple and efficient photocatalytic reactors. Generally, photocatalytic reactors are

classified as (a) slurry reactors and (b) immobilized reactors. In slurry reactors, TiO2

powder is dispersed in water and continual mixing ensures it has good contact with light

source. Commercial Degussa P25 is acknowledged as the best readily available TiO2

powder with high photocatalytic activity. In contrast, in immobilized reactors,

photocatalysts are immobilized on inert substrates and become less photon-efficient due

to their low active surface area to volume ratio. However, immobilized reactors are more

Page 139: 1.3 Light emitting diode

120

practical as they do not require the secondary step of separation of TiO2 particles from

treated waters. Direct anodization of titanium is recognized as one of the better methods

to prepare immobilized TiO2 film due to its simplicity and being able to control the

thickness and morphology of a nanotubular TiO2 film (Li et al., 2009). One way to

improve the surface area to reaction volume ratio in an immobilized photocatalytic

reactor is to coat TiO2 on 250 micron to millimeter sized particles that can be easily

separated (Vega et al., Geng and Cui, 2010, Imoberdorf et al., 2008a, Vaisman et al.,

2005, Pozzo et al., 2005, Kanki et al., 2005, Chiovetta et al., 2001, Haarstrick et al.,

1996).

Most conventional photocatalytic reactors obtain their ultraviolet (UV) irradiation from

mercury arc lamps which are not environmental friendly and can be less energy efficient

than light emitting diodes (LEDs). Recently, the development of UV-LED technology

has made UV-LEDs lamps a promising replacement for mercury lamps in TiO2

photocatalysis application (Yu et al., 2013, Natarajan et al., 2011b). The advantages of

UV-LEDs include smaller size, higher durability, longer life, narrower spectrum, energy

efficiency, and fast switching.

In this paper, the details of design and fabrication of a novel immobilized UV-LED

photocatalytic reactor is presented. The reactor have been evaluated by testing two

phenoxy pesticides [2,4-dichlorophenoxyacetic acid (2,4-D), 2-methyl-4-

chlorophenoxyacetic acid (MCPA)] and chlorophenols. The operational parameters of

reactor were optimized using the 2,4-D degradation as an example.

Page 140: 1.3 Light emitting diode

121

6.2 Experimental details

6.2.1 Chemicals

Ninety-nine percent pure 2,4-D , 98% pure MCPA, 99% pure 4-CP, 99% pure 2,4-DCP,

98% pure formic acid, 98% pure ammonium fluoride as well as 99.7% pure Ti foil were

procured from Sigma Aldrich; HPLC grade acetonitrile and 99% pure ethylene glycol

were obtained from VWR. TiO2 (P25) was obtained from Degussa. Hollow glass

microspheres coated with anatase TiO2 (HGMT) were obtained from Cospheric, which

has a median diameter of 45 µm. All chemicals were used as received and Milli-Q water

was used in all experiments.

6.2.2 Design and fabrication of an LED based photocatalytic reactor

6.2.2.1 Preparation of anodized TiO2 photocatalytic plate.

In this research, the immobilized TiO2 photocatalytic plate was prepared by

electrochemical anodization. Prior to anodization, a 15 cm by 15 cm titanium foil was

dipped in a mixture of acetone, methanol and methylene chloride and sonicated for 30

minutes for degreasing (Wang and Lin, 2009). Then the titanium foil was rinsed and

cleaned with water and dried in fume hood. The anodization was conducted in a two

electrode cell setup with Ti foil as an anode and an aluminum foil as a cathode. The

electrolyte used was an ethylene glycol solution containing 2 % water and 0.5 %

ammonium fluoride. The anodization was performed with a static potential (30V) using a

Lambda Regulated Power supply (Model: LE 104-FM), at a room temperature for 24

hours. The anodized Ti foil was washed with water and dried in the fume hood.

Page 141: 1.3 Light emitting diode

122

Thereafter, the dried anodized Ti foil was annealed at 450oC in air for 3 hours to induce

crystallization to give a better photocatalytic ability (Chang et al., 2011) . The surface

structure of anodized TiO2 was examined using scan electron microscopy (SEM). The

SEM result (Figure 6-1) showed that TiO2 nanotubes were synthesized on the surface of

the titanium foil. The diameters of nanotubes were approximately 100 nm.

Figure 6- 1: SEM image of anodized TiO2 nanostructrure.

6.2.2.2 UV-LEDs module

The UV-LEDs module composed of 16 UV-LED lamps (NSCU330B, Nichia

Corporation, Japan). The lamp has a sharp peak at λ=365nm with the half width band of 9

Page 142: 1.3 Light emitting diode

123

nm. Individual lamps were assembled into a four by four array and mounted on a 10 cm

by 10 cm square circuit board. The distance between adjacent lamps was 2.5 cm. The

UV-LEDs were driven by a high current Quad output LED driver (LT3476). To generate

different light intensities a dual output power supply (Model TW5005D) was used to

provide direct currents ranging from 10 mA to 500 mA.

6.2.2.3 Photocatalytic system

The UV-LED module and photocatalytic plate were mounted parallel in a polyvinyl

chloride (PVC) based casing (Figure 6-2). A Hydro H55 CPU Cooler on the back of

LED module was used to control the temperature. A Pyrex borosilicate glass plate (CG-

1904-16, Chemglass life Sciences) was used to shield electronic parts. UVA light passes

efficiently through the Pyrex glass plate to reach the rectangular reaction zone.

Figure 6-2: Scheme of an LED based photocatalytic reactor.

Page 143: 1.3 Light emitting diode

124

The rectangular reaction zone has a width of 10 cm, a length of 10 cm and a depth of 5

cm. The distance between the shielding glass and the photocatalyst plate (Ds-p) can be

adjusted by inserting the photocatalyst plate in different slots. The distances between

different slots and the shielding glass plate were respectively 1 cm, 3 cm and 5 cm.

The inlets were on the sidewall perpendicular to the photocatalytic plate. The inlet

manifold system was composed of twelve small tubes. Openings on the side of each tube

face the photocatalytic plate. Such design enhanced the contact between influent

contaminants and the photocatalytic plate. Three small holes located at the top of the

opposite sidewall of reactor were used as outlets.

6.2.3 Radiation field and light intensity estimation

Table 6-1: Average light intensity received by the photocatalytic plate.

Input current (mA)

Ds-p (cm)

DL-P

(cm) Ia

(mw/cm2)

500 1 1.4 29.4 500 3 3.4 22.6 500 5 5.4 17.3 250 5 5.4 8.6 125 5 5.4 4.3 62 5 5.4 2.2

note: DL-P is larger than Ds-p due to the thickness of shielding glass plate and the gap

between lamps and the shielding glass.

Page 144: 1.3 Light emitting diode

125

Figure 6-3: Radiation field on a photocatalyst plate under different conditions; (a) DL-P = 0.014 m, 4 by 4 LEDs panel; (b) DL-P = 0.034 m, 4 by 4 LEDs panel; (c) DL-P = 0.054 m, 4 by 4 LEDs panel.

Page 145: 1.3 Light emitting diode

126

The average light intensity received by photocatalyst plate was estimated based on an

emission radiation field model developed for this design (see Chapter Five). The average

light intensities (Ia) for different situations are listed in Table 6-1. And the light intensity

distribution for different distance between the shielding glass and the photocatalyst plate

(DL-P ) with an input current of 500 mA is shown in Figure 6-3. Light intensity received

by the centre of photocatalyst plate was verified by a Silver Line UV radiometer

(M007153, Geneq Inc. Canada). When DL-P is 5.4 cm and the current is 500 mA, the

value reading from radiometer was 23 mW cm-2, which is in agreement with the model

data (Figure 6-3c).

6.2.4 Experimental set-up and sample analysis

A stock solution of 1.50 L of 20 mg L-1 pesticides or chlorophenol solution was prepared

by dissolving 30 mg of pure compound in water using ultrasonication. A peristaltic pump

(LaSalle Scientific Inc, Model: 400-205) circulated the solution containing the target

compounds between a reservoir covered with aluminum foil and the photocatalytic

reactor. The solution in the reservoir is continuously mixed with a magnetic bar.

Experiments were conducted at variable flow rate, variable light intensity, variable DL-P

and different photocatalyst configuration. All experimental conditions are described in

the captions of the figures. Prior to irradiation, the solution containing pesticides or

chlorophenols was circulated for 30 minutes to ensure that the adsorption of the

investigated compound onto the surface of photocatalyst reached equilibrium. After

different irradiation periods, a 1.0 mL sample from the reservoir was collected and

analyzed using a Varian Prostar 210 high performance liquid chromatography equipped

Page 146: 1.3 Light emitting diode

127

with a 325 LC UV-Vis detector (Yu et al., 2013). Variations in the obtained data in are

shown by error bars.

6.3 Result and discussion

6.3.1 Degradation of phenoxy pesticides and chlorophenols in a flow-through LED based photocatalytic reactor

To better understand the reactor performance from an energy perspective, the results are

expressed as the change of normalized concentration of parent compound versus the

energy dosage per unit volume. Energy dosage per unit volume is defined as the energy

of photons entering the reaction zone divided by the volume of sample treated (1.50 L).

The degradation of two phenoxy pesticides (2,4-D, MCPA) and two chlorophenols (4-

CP, 2,4-DCP) in the photocatalytic reactor are shown in Figure 6-4. The normalized

concentrations of parent compounds decreased as the energy dosage increased. With an

energy dosage of 25 kJ L-1, 91% of 2,4-D 85% of MCPA, 85% of 2,4-DCP and 81% of

4-CP were removed from the solution (approximately one log reduction). As reported in

Yu et al. (2013), these four compound can be efficiently photocatalytically decomposed

with slurry TiO2 in a UVA-LED batch reactor. In our immobilized photocatalytic reactor,

the TiO2 nanotubes growing on the titanium plate can also efficiently capture the UV

photons (365nm). The results show that both phenoxy pesticides and chlorophenol can be

degraded efficiently under the operational conditions shown in the caption of Figure 6-4.

Page 147: 1.3 Light emitting diode

128

Figure 6-4: Photodegradation of MCPA, 2,4-D, 2,4-DCP and 4-CP in a UV-LED photoreactor: flow rate = 2.03 L min-1; DL-P = 0.54 cm; Ia=17.3 mW cm-2.

6.3.2 Degradation of 2,4-D with different combination of (UV, TiO2 photocatalyst plate, H2O2 and O2) in the UV-LED photoreactor .

Photodegradation of 2,4-D under different experimental conditions is shown in Figure 6-

5. With only UV-LED irradiation, 13% of 2,4-D was removed at an energy dosage of

6.22 kJ L-1. In the UV-visible spectrum range from 250 nm-700 nm, 2,4-D has a single

peak with a maximum at around 280 nm and does not have an absorption at 365 nm

(peak wavelength for the LEDs). However, there is still a small amount of photons in the

UV-LED emission spectrum, leading to a direct photolysis of 2,4-D. The presence of

H2O2 (0.1%) in LED reactor did improve the degradation efficiency and 54% of 2,4-D

was removed from bulk solution with the same energy dosage. H2O2 has weak

absorption on the emission spectrum of LED, which may cause the photolysis of

hydrogen peroxide and generate hydroxyl radicals.

0.00

0.20

0.40

0.60

0.80

1.00

1.20

0 5 10 15 20 25 30

C/C

o

Energy dosage per volume (kJ/L)

2,4-DCP 4-CP

MCPA 2,4-D

Page 148: 1.3 Light emitting diode

129

Figure 6-5: Photodegradation of 2,4-D in a flow-through UV-LED photoreactor: flow rate = 2.03 L min-1; DL-P = 0.54 cm ; Ia=17.3 mW cm-2.

The reactor was considered as a photocatalytic reactor while being mounted with a TiO2

photocatalyst plate. Such photocatalytic reactor can eliminate 40 % of 2,4-D at an energy

dosage of 6.22 kJ L-1. Bubbling oxygen in this photocatalytic reactor did not significantly

improve the degradation efficiency. In the experiments, the solution in reservoir is

thoroughly mixed using a magnetic stirrer. The aeration led to the sample getting

saturated with oxygen activity near 0.2 atm. This provided enough oxygen for

photocatalytic reactions. Therefore, further addition of oxygen into the system did not

boost the photocatalytic degradation. Apparently, the presence of 0.1% H2O2 in the

photocatalytic system resulted in a better degradation efficiency. At an energy dosage of

0.00

0.10

0.20

0.30

0.40

0.50

0.60

0.70

0.80

0.90

1.00

0 1 2 3 4 5 6 7

C/C

o

Energy dosage per volume (kJ/L)

UV only UV+H2O2 (0.1%) UV+TiO2 UV+TiO2+bubbling Oxygen UV+TiO2+ H2O2 (0.1%)

Page 149: 1.3 Light emitting diode

130

6.22 kJ L-1, 80% of 2,4-D is degraded. Hydrogen peroxide serves as a good electron

scavenger and accelerates the photocatalytic reaction.

6.3.3 Effect of DL-P

DL-P is a key factor for system scale-up. To study its effect on photocatalytic degradation,

experiments were conducted at three DL-P (1.4 cm, 3.4 cm and 5.4 cm). The results (see

Figure 6-6) show that the degradation of 2,4-D is relatively slow when DL-P is set to be

1.4 cm. The degradation efficiency improved as DL-P was increased to 3.4 cm, while

further increment of DL-P to 5.4 cm did not enhance the degradation efficiency. DL-P can

impact the photocatalytic degradation in two ways: (1) for the same photon energy input,

a uniform radiation field can result in more efficient distribution of activity over the

photocatalyst. In this reactor, a less uniform radiation field is obtained at shorter DL-P

(Figure 6-3); (2) for an immobilized photocatalytic reactor, the photocatalytic

degradation is limited by the mass transfer of the contaminants between the photocatalyst

surface and the bulk solution (Chen et al., 2001). At the same flow rate, the Reynolds

number decreases with DL-P, and hinders mass transfer. Therefore, from a kinetics

perspective, an optimal DL-P should make the light intensity received on the photocatalyst

plate uniform and not inhibit mass transfer.

One way to scale-up this system is to use a baffle reactor design which contains multiple

modules composed of UV-LED plate and photocatalytic plate. Each module has a

reaction zone and dead zone accommodating the electronics. The reaction zone volume

can be adjusted by changing DL-P, while the dead zone volume is limited to the electronic

Page 150: 1.3 Light emitting diode

131

part. The advantage of larger DL-P is that fewer modules are required for the same

reaction zone volume, and the total volume of reactor occupied is reduced. Therefore, in

this research, DL-P of 5.4 cm is superior to a DL-P of 3.4 cm.

Figure 6- 6: The effect of DL-P on 2,4-D degradation: flow rate =2.03 L min-1, Ia=17.3 mW cm-2.

6.3.4 Effect of flow rates on the photocatalytic degradation of 2,4-D.

To investigate the effect of flow rate on performance of LED photocatalytic reactor,

experiments were conducted at four different flow rates (0.72 L min-1, 1.50 L min-1, 2.03

L min-1 and 2.87 L min-1) and the results are shown in Figure 6-7. At the lowest flow rate

(0.72 L min-1), only 28 % 2,4-D removal was achieved with an energy dosage of 6.22

kJ/L. The removal percentage was improved as the flow rate increased and 40% of 2,4-D

0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70 0.80 0.90 1.00

0 2 4 6 8 10 12

C/C

o

Energy dosage per volume (kJ L-1)

D = 1.4 cm

D = 3.4 cm

D = 5.4 cm

Page 151: 1.3 Light emitting diode

132

was eliminated at a flow rate of 1.5 L min-1. The enhancement of degradation efficiency

due to the increase of flow rate was less significant at high flow rates.

Figure 6-7: The effect of flow rate on degradation of 2,4-D: DL-P = 5.4 cm, Iaverage=17.3 mW cm-2.

In a flow-through immobilized photocatalytic reactor, the flow rate impacts the mass

transfer of reactants between the photocatalyst surface and bulk solution. Higher flow

rate lead to a higher mass transfer rate and a faster overall reaction rate is expected. Flow

rate, along with reaction zone volume, determine the residence time of contaminants in

the reactor. In this study, the experiments were carried out in the circulated mode,

therefore, the residence time did not depend on flow rate but on the total operational time

and the ratio of reaction zone volume/total volume.

0.40

0.50

0.60

0.70

0.80

0.90

1.00

0 1 2 3 4 5 6 7

C/C

o

Energy dosage per volume (kJ L-1)

flow rate=0.72L/min

flow rate=1.50L/min

flow rate=2.03 L/min

flow rate=2.87 L/min

Page 152: 1.3 Light emitting diode

133

6.3.5 Effect of UV light intensity

The photocatalytic reactor performance at four different light intensities (2.2, 4.3, 8.6 and

17.3 mw cm-2) were investigated. Figure 6-8(a) summarized the first order kinetic rate

constants (K) at different light intensities. The lowest k (1.3×10-4 s-1) was obtained at a

light intensity of 2.2 mW cm-2 and the highest k (2.9×10-4 s-1) was obtained at a light

intensity of 17.3 mW cm-2. The first order kinetic rate constants increase with light

intensity fitting a power law relationship described by Equation [6-1].

𝑘 = 1.02 × 10−4 × 𝐼𝑎0.3753 [6-1]

The photocatalytic degradation rate kinetics depend on the efficiency of electron-hole

generation and recombination (Ollis et al., 1991). At lower light intensity range, the

electron-hole generation dominates and the reaction rate increases linearly with absorbed

irradiation intensity to a critical value. At a relatively higher light intensity, an increase of

the electron-hole recombination dominates and a power law relationship is obtained, as

observed in this study.

The power law relationship with an exponent less than one indicates a lower quantum

efficiency at a higher light intensity. Figure 6-8b reports results as a function of energy

dosage and showed that in our studied light intensity range, the low light intensity

condition is favored for energy efficiency.

Page 153: 1.3 Light emitting diode

134

(a)

(b)

Figure 6-8: The effect of light intensity on degradation of 2,4-D: DL-P = 5.4 cm, Flow rate=2.03 L min-1.

0.00E+00

1.00E-04

2.00E-04

3.00E-04

4.00E-04

0 5 10 15 20

k (s

-1)

Ia (mW cm-2)

0.00

0.20

0.40

0.60

0.80

1.00

1.20

0 1 2 3 4 5 6 7

C/C

o

Energy dosage per volume (kJ L-1)

I=17.3 mW/cm² I=8.6 mW/cm²

I=4.3 mW/cm² I=2.2 mW/cm²

Page 154: 1.3 Light emitting diode

135

6.3.6 Comparison of three different photocatalyst configurations

2,4-D photocatalytic degradation experiments were conducted with three different

photocatalyst configurations. They are: type (a), anodized TiO2 photocatalytic plate (10

cm by 10 cm); type (b), 2 g L-1 of P25, with an average diameter of 20 nm; type (c), 5 g

L-1 of hollow glass microspheres coated with anatase TiO2 (HGMT), of median particle

diameter of 45 µm.

Table 6-2: First order kinetic rate constants for different photocatalyst configurations

Photocatalyst configuration k (*10-4 s-1)

Type (a) 2.9 Type(b) 31.8 Type (c) 13.2

The first order rate constants for each case were reported in Table 6-2. The results show

that reaction rates in slurry type [type (b)] is ten times faster than that in immobilized

type [type (a)], and the performance of HGMT [type (c)] in removing 2,4-D is between

these two types. Note that the loading of HGMT is higher than that of P25.

The configuration of photocatalyst is a key factor for performance. The access to catalytic

surface by the photons and the reactants determines rate. Larger available catalytic

surface results in a higher rate. Among these three configurations, type (a) has the least

available surface area. Moreover, mass transfer of reactants becomes a limiting factor in

an immobilized catalyst type reactor. Type (b) is better than Type (c) possibly due to

Page 155: 1.3 Light emitting diode

136

higher photocatalyst loading, higher surface area and easier access to the surface of the

photocatalyst.

Kinetics in a slurry reactor is much superior to that in an immobilized type whereas the

operation of an immobilized reactor requires no further separation step. A modified

configuration-HGMT with a suitable concentration can result in a reaction rate

comparable to P25. Besides, the low density (0.22 g cm-3) of HGMT can make it float on

the surface of water and be conveniently recovered.

6.4 Conclusions

This paper presents the design and fabrication of a novel UV-LED based photocatalytic

reactor. The reactor shows its capability to decompose phenoxy pesticides and

chlorophenols. The study on different operational parameters, such as DL-P, flow rate,

light intensity and external electron scavenger provide useful information for system

scale-up. In this reactor, optimal DL-P was determined to be 5.4 cm and 1.5 L min-1 was

chose as a suitable flow rate. The power law relationship with an exponent 0.4 between

first order kinetics rate constants and the studied light intensities indicate increasing light

intensity to reduce reaction time is not energy efficiency at high power input. Adding

hydrogen peroxide is a good option to boost the reactor performance. Furthermore, a

modified photocatalyst (hollow glass microsphere coated with anatase TiO2) can be a

promising photocatalyst configuration, considering the reaction rate and operational

convenience.

Page 156: 1.3 Light emitting diode

137

Chapter Seven: CONCLUSION AND RECOMMENDATION FOR FUTURE RESEARCH

7.1 Conclusions

The photochemical technologies (photosensitization and TiO2 based photocatalysis)

developed in this research have successfully treated PCBs and pesticides in aqueous

medium. LEDs were shown to be a promising light source in TiO2 photocatalytic

application. As an important step for designing an efficient photo-reactor, a radiation

model was developed and validated. Finally, based on these work, an LED based flow-

through photocatalytic reactor were designed, fabricated and optimized. The reactor has

shown its capacity to efficiently treat water based contaminants like pesticides and

chlorophenols. The overall conclusion of this thesis can be further divided into four sub-

conclusions as given below:

7.1.1 Photosensitized dechlorination of PCBs solubized in surfactant solution

It is possible to dechlorinate PCBs in an aqueous medium using longer

wavelength ( visible light). The usage of visible light opens an opportunity to

utilize sunlight for PCBs treatment, thus significantly reducing the energy costs.

The types and the concentrations of surfactant can impact the PCBs

dechlorination rate. The cationic (CTAB) and non ionic (TWEEN 80) surfactants

work better than the anionic surfactant (SDS) for dechlorination, even though the

cationic surfactant is not preferred for PCB extraction from soil.

Page 157: 1.3 Light emitting diode

138

7.1.2 LED based photocatalytic treatment of pesticides and chlorophenols

Complete decomposition of the studied phenoxy pesticides and chlorophenols

was achieved within a short period of irradiation with a slurry TiO2 in a batch

UV-LED (365nm) reactor.

Due to the competition for surface hydroxyl radicals between different

compounds, the degradation rate of pesticides become slower as a second

pesticide is introduced into the solution.

A suitable loading of TiO2 (2g/L) for 2,4-dichlorophenoxyacetic acid degradation

was determined at the applied light intensity (8.55×1016 photon s-1). The rate

constant at this loading and this light intensity was found to be 0.2404 min-1.

The first order kinetic rate constants were proportional to the studied light

intensities.

The comparison between mercury lamps and UVA-LEDs show that UVA-LEDs

is more energy-efficient, since the emission spectrum of UVA-LED lamps can be

well-matched with the absorption band of TiO2. If a solar reactor is considered as

a competitor, the photon energy-use advantage of UVA-LEDs is even greater.

7.1.3 Design a homogenous radiation field model for photocatalytic reactor

Radiation field model for a UV-LEDs array has been developed, which can

predict the light intensity at any location of a photocatalyst plate with any ID.

Based on the model, the degree of homogeneity of the radiation field is

significantly affected by the ID when the gap is fixed.

Page 158: 1.3 Light emitting diode

139

A homogenous radiation field can be achieved by choosing an optimal ID/gap

ratio. The ratio is found to be 1.26 for Nichia UV-LED (NCSU033).

The method of selecting the light output of UV-LED for different gaps to achieve

a desired homogenous light intensity was developed and evaluated.

7.1.4 A novel light emitting diode photocatalytic reactor for water treatment

A novel LED based photocatalytic reactor was designed and fabricated. The novel

reactor is a combination of an environmental friendly light source and an

immobilized nanostructure photocatalyst.

The reactor shows its ability to treat water contaminated with phenoxy pesticides

and chlorophenols under different experimental conditions.

The study on the DL-P, flow rate, light intensity and external electron scavenger

provide useful information for reactor scale-up. Optimal DL-P was determined to

be 5 cm and 2 L/min was chose as a suitable flow rate for current reactor. The

power law relationship with a exponent 0.4 between kinetics and light intensity

was examined, indicating lower energy efficiency is reduced when the light

intensity is increased.

Page 159: 1.3 Light emitting diode

140

7.2 Recommendations for Future Research

7.2.1 Incorporating PCBs extraction using surfactants and PCBs photodechlorination using sensitized visible light

Our research shows that PCBs in surfactant solution can be dechlorinated using sensitized

visible light. Therefore, a study of PCBs extraction using surfactants followed by

dechlorination using sensitized visible light should be conducted in the future.

7.2.2 UVC-LED

Currently, the lower quantum efficiency and high price of UVC-LED limit its

application. Therefore, our research was focused on UVA-LED. Once high intensity

UVC-LED become commercially available as the development of the UVC-LED

technology, other AOPs requiring deep ultraviolet light, such as UV/H2O2, can be

investigated using UVC-LED.

7.2.3 The decay of photocatalytic activity and its life time

The activity of photocatalyst may decay with time, which can reduce the reactor

performance. Thus, factors causing the inactivation of photocatalyst and the methods for

regenerating the photocatalyst need to be studied. In addition, the decay of photocatalytic

activity can be used to predict the life time of photocatalyst.

7.2.4 Hollow microsphere coated with TiO2 (HGMT)

HGMT open a promising future for photocatalytic applications. Since the density of

HGMT is much lower than water, mixing of HGMT with water will become an issue. A

way with less energy to mix HGMT with water is needed to be studied.

Page 160: 1.3 Light emitting diode

141

7.2.5 Scale-up of the reactor

The reactor can be scaled up using a baffle design (Figure 7-1) with multiple LED panels

and photocatalyst plates. The hydraulic conditions and the mass transfer of reactants and

products in such system need to be studied.

Figure 7-1: Scheme of a scale-up LED based photocatalytic reactor.

Page 161: 1.3 Light emitting diode

142

REFERENCES

ACHARI, G., GUPTA, C., DHOL, A., LANGFORD, C. H. & JAKHER, A. 2003.

Photochemical dechlorinated of highly chlorinated PCBs. University of

Calgary(report submitted to TransCanada).

ADESINA, A. A. 2004. Industrial exploitation of photocatalysis: progress, perspectives

and prospects. Catalysis Surveys from Asia, 8, 265-273.

ADIVARAHAN, V., SUN, W., CHITNIS, A., SHATALOV, M., WU, S., MARUSKA,

H. & KHAN, M. A. 2004. 250 nm AlGaN light-emitting diodes. Applied Physics

Letters, 85, 2175.

AHMED, S., RASUL, M. G., BROWN, R. & HASHIB, M. A. 2011. Influence of

parameters on the heterogeneous photocatalytic degradation of pesticides and

phenolic contaminants in wastewater: A short review. Journal of Environmental

Management, 92, 311-330.

AKPAN, U. G. & HAMEED, B. H. 2009. Parameters affecting the photocatalytic

degradation of dyes using TiO2-based photocatalysts: A review. Journal of

Hazardous materials, 170, 520-529.

AKSU, Z., ERTUGRUL, S. & DÖNMEZ, G. 2010. Methylene Blue biosorption by

Rhizopus arrhizus: Effect of SDS (sodium dodecylsulfate) surfactant on

biosorption properties. Chemical Engineering Journal, 158, 474-481.

ALFANO, O. M., BAHNEMANN, D., CASSANO, A. E., DILLERT, R. & GOSLICH,

R. 2000. Photocatalysis in water environments using artificial and solar light.

Catalysis Today, 58, 199-230.

Page 162: 1.3 Light emitting diode

143

ALFANO, O. M., ROMERO, R. L. & CASSANO, A. E. 1986a. A cylindrical

photoreactor irradiated from the bottom--II. Models for the local volumetric rate

of energy absorption with polychromatic radiation and their evaluation. Chemical

Engineering Science, 41, 1155-1161.

ALFANO, O. M., ROMERO, R. L. & CASSANO, A. E. 1986b. Radiation field

modelling in photoreactors--I. homogeneous media. Chemical Engineering

Science, 41, 421-444.

ALFANO, O. M., ROMERO, R. L. & CASSANO, A. E. 1986c. Radiation field

modelling in photoreactors--II. Heterogeneous media. Chemical Engineering

Science, 41, 1137-1153.

ALLAM, N. K. & GRIMES, C. A. 2008. Effect of cathode material on the morphology

and photoelectrochemical properties of vertically oriented TiO2 nanotube arrays.

Solar Energy Materials and Solar Cells, 92, 1468-1475.

ALONSO-TELLEZ, A., MASSON, R., ROBERT, D., KELLER, N. & KELLER, V.

2012. Comparison of Hombikat UV100 and P25 TiO2 performance in gas-phase

photocatalytic oxidation reactions. Journal of Photochemistry and Photobiology

A: Chemistry, 250, 58-65.

ALTENA, F., VAN OVERVELD, J. B. J. & GILLER, H. Technological advances in

disinfection lamps leading to more compact UV sources. Conference Proceedings

of the First International Congress on Ultraviolet Technologies, International

Ultraviolet Assoc.(IUVA), Washigton DC, electronic release, 2001.

Page 163: 1.3 Light emitting diode

144

ANDREOZZI, R., CAPRIO, V., INSOLA, A. & MAROTTA, R. 1999. Advanced

oxidation processes (AOP) for water purification and recovery. Catalysis Today,

53, 51-59.

AUGUGLIARO, V., PALMISANO, L., SCLAFANI, A., MINERO, C. & PELIZZETTI,

E. 1988. Photocatalytic degradation of phenol in aqueous titanium dioxide

dispersions. Toxicological and Environmental Chemistry, 16, 89-109.

BAHNEMANN, W., MUNEER, M. & HAQUE, M. M. 2007. Titanium dioxide-

mediated photocatalysed degradation of few selected organic pollutants in

aqueous suspensions. Catalysis Today, 124, 133-148.

BEHNAJADY, M. A., MODIRSHAHLA, N., SHOKRI, M. & VAHID, B. 2009. Design

equation with mathematical kinetic modeling for photooxidative degradation of

C.I. Acid Orange 7 in an annular continuous-flow photoreactor. Journal of

Hazardous Materials, 165, 168-173.

BELGIORNO, V., RIZZO, L., FATTA, D., DELLA ROCCA, C., LOFRANO, G.,

NIKOLAOU, A., NADDEO, V. & MERIC, S. 2007. Review on endocrine

disrupting-emerging compounds in urban wastewater: occurrence and removal by

photocatalysis and ultrasonic irradiation for wastewater reuse. Desalination, 215,

166-176.

BERGEN, B. J., RAHN, K. A. & NELSON, W. G. 1998. Remediation at a Marine

Superfund Site:  Surficial Sediment PCB Congener Concentration, Composition,

and Redistribution. Environmental Science & Technology, 32, 3496-3501.

Page 164: 1.3 Light emitting diode

145

BHATKHANDE, D. S., PANGARKAR, V. G. & BEENACKERS, A. A. C. M. 2002.

Photocatalytic degradation for environmental applications – a review. Journal of

Chemical Technology and Biotechnology, 77, 102-116.

BHATTACHARYA, A. 2006. Remediation of Pesticide-Polluted Waters Through

Membranes. Separation and Purification Reviews, 35, 1-38.

BICKLEY, R. I., GONZALEZ-CARRENO, T., LEES, J. S., PALMISANO, L. &

TILLEY, R. J. D. 1991. A structural investigation of titanium dioxide

photocatalysts. Journal of Solid State Chemistry, 92, 178-190.

BOLTON, J. 1989. Economics of solar-energy conversion using a product-value analysis.

Chimia, 43, 226-227.

BOUSSAHEL, R., BOULAND, S., MOUSSAOUI, K. M. & MONTIEL, A. 2000.

Removal of pesticide residues in water using the nanofiltration process.

Desalination, 132, 205-209.

BRAHAM, R. J. & HARRIS, A. T. 2009. Review of Major Design and Scale-up

Considerations for Solar Photocatalytic Reactors. Industrial and Engineering

Chemistry Research, 48, 8890-8905.

BRIMBLE, S., BACCHUS, P. & CAUX, P. Y. 2005. Pesticide Utilization in Canada: A

compilation of current sales and use data. In: CANADA, E. (ed.). Gatineau, QC.

BUNCE, N. J., KUMAR, Y., RAVANAL, L. & SAFE, S. 1978. Photochemistry of

chlorinated biphenyls in iso-octane solution. Journal of the Chemical Society,

Perkin Transactions 2, 880-884.

Page 165: 1.3 Light emitting diode

146

BURROWS, H. D., CANLE L, M., SANTABALLA, J. A. & STEENKEN, S. 2002.

Reaction pathways and mechanisms of photodegradation of pesticides. Journal of

Photochemistry and Photobiology B: Biology, 67, 71-108.

CALVERT, J. & PITTS, J. 1966. Photochemistry, New York, John Wiley.

CAREY, J. H., LAWRENCE, J. & TOSINE, H. M. 1976. Photodechlorination of PCB's

in the presence of titanium dioxide in aqueous suspensions. Bulletin of

Environmental Contamination and Toxicology, 16, 697-701.

CARVALHO, M. G. & FARIAS, T. L. 1998. Modelling of Heat Transfer in Radiating

and Combusting Systems. Chemical Engineering Research and Design, 76, 175-

184.

CASSANO, A. E., MARTIN, C. A., BRANDI, R. J. & ALFANO, O. M. 1995.

Photoreactor Analysis and Design: Fundamentals and Applications. Industrial

and Engineering Chemistry Research, 34, 2155-2201.

CHANG, F. C., CHIU, T. C., YEN, J. H. & WANG, Y. S. 2003. Dechlorination

pathways of ortho-substituted PCBs by UV irradiation in n-hexane and their

correlation to the charge distribution on carbon atom. Chemosphere, 51, 775-784.

CHANG, Y. J., LEE, J. W., WENG, G. J., LEE, C. K. & HUANG, Y. C. 2011. The

Influence of Annealing Temperatures on the Crystalline and Photocatalytic

Abilities of Anodized TiO2 Nanotube Arrays. Advanced Materials Research, 261,

623-627.

CHEN, D., LI, F. & RAY, A. K. 2001. External and internal mass transfer effect on

photocatalytic degradation. Catalysis Today, 66, 475-485.

Page 166: 1.3 Light emitting diode

147

CHEN, D. H., YE, X. & LI, K. 2005. Oxidation of PCE with a UV LED Photocatalytic

Reactor. Chemical Engineering and Technology, 28, 95-97.

CHEN, H. Y., ZAHRAA, O. & BOUCHY, M. 1997. Inhibition of the adsorption and

photocatalytic degradation of an organic contaminant in an aqueous suspension of

TiO2 by inorganic ions. Journal of Photochemistry and Photobiology A:

Chemistry, 108, 37-44.

CHEN, S. & LIU, Y. 2007. Study on the photocatalytic degradation of glyphosate by

TiO2 photocatalyst. Chemosphere, 67, 1010-1017.

CHEN, Y., LU, A., LI, Y., YIP, H. Y., AN, T., LI, G., JIN, P. & WONG, P.-K. 2011.

Photocatalytic inactivation of Escherichia coli by natural sphalerite suspension:

Effect of spectrum, wavelength and intensity of visible light. Chemosphere, 84,

1276-1281.

CHIARENZELLI, J., SCRUDATO, R., WUNDERLICH, M., RAFFERTY, D., JENSEN,

K., OENGA, G., ROBERTS, R. & PAGANO, J. 1995. Photodecomposition of

PCBs absorbed on sediment and industrial waste: implications for photocatalytic

treatment of contaminated solids. Chemosphere, 31, 3259-3272.

CHIOVETTA, M. G., ROMERO, R. L. & CASSANO, A. E. 2001. Modeling of a

fluidized-bed photocatalytic reactor for water pollution abatement. Chemical

Engineering Science, 56, 1631-1638.

CHITNIS, A., ZHANG, J., ADIVARAHAN, V., SHATALOV, M., WU, S.,

PACHIPULUSU, R., MANDAVILLI, V. & KHAN, M. A. 2003. Improved

performance of 325-nm emission AlGaN ultraviolet light-emitting diodes.

Applied Physics Letters, 82, 2565-2567.

Page 167: 1.3 Light emitting diode

148

CHOI, W., HONG, S. J., CHANG, Y.-S. & CHO, Y. 2000. Photocatalytic Degradation of

Polychlorinated Dibenzo-p-dioxins on TiO2 Film under UV or Solar Light

Irradiation. Environmental Science & Technology, 34, 4810-4815.

CHONG, M. N., JIN, B., ZHU, H. Y., CHOW, C. W. K. & SAINT, C. 2009. Application

of H-titanate nanofibers for degradation of Congo Red in an annular slurry

photoreactor. Chemical Engineering Journal, 150, 49-54.

CHU, W., CHAN, K. H., KWAN, C. Y. & JAFVERT, C. T. 2005. Acceleration and

Quenching of the Photolysis of PCB in the Presence of Surfactant and Humic

Materials. Environmental Science & Technology, 39, 9211-9216.

CHU, W., JAFVERT, C. T., DIEHL, C. A., MARLEY, K. & LARSON, R. A. 1998.

Phototransformations of Polychlorobiphenyls in Brij 58 Micellar Solutions.

Environmental Science & Technology, 32, 1989-1993.

CHU, W. & KWAN, C. Y. 2002. The direct and indirect photolysis of 4,4'-

dichlorobiphenyl in various surfactant/solvent-aided systems. Water Research, 36,

2187-2194.

CHU, W. & KWAN, C. Y. 2003. Reactor design and kinetics study of 4,4'-

dichlorobiphenyl photodecay in surfactant solution by using a photosensitizer and

hydrogen source. Water Research, 37, 2442-2448.

CHU, W. & WONG, C. C. 2004. The photocatalytic degradation of dicamba in TiO2

suspensions with the help of hydrogen peroxide by different near UV irradiations.

Water Research, 38, 1037-1043.

Page 168: 1.3 Light emitting diode

149

CRANK, G. & MURSYIDI, A. 1992. Oxidations of thioureas with photochemically

generated singlet oxygen. Journal of Photochemistry and Photobiology A:

Chemistry, 64, 263-271.

D'OLIVEIRA, J. C., GUILLARD, C., MAILLARD, C. & PICHAT, P. 1993.

Photocatalytic destruction of hazardous chlorine‐ or nitrogen‐containing

aromatics in water. Journal of Environmental Science and Health . Part A:

Environmental Science and Engineering and Toxicology, 28, 941-962.

DEMPSEY, B. A. & O'MELIA, C. R. 1984. Removal of naturally occurring compounds

by coagulation and sedimentation. Critical Reviews in Environmental Control, 14,

311-331.

DENNY, F., SCOTT, J., PAREEK, V., DING PENG, G. & AMAL, R. 2009. CFD

modelling for a TiO2-coated glass-bead photoreactor irradiated by optical fibres:

Photocatalytic degradation of oxalic acid. Chemical Engineering Science, 64,

1695-1706.

DEVIPRIYA, S. & YESODHARAN, S. 2005. Photocatalytic degradation of pesticide

contaminants in water. Solar Energy Materials and Solar Cells, 86, 309-348.

DHOL, A. S. 2005. An investigation of a photochemical approach for the remediation of

PCB-contaminated soils. Master of Science, University of Calgary.

DUTTA, P. K. & RAY, A. K. 2004. Experimental investigation of Taylor vortex

photocatalytic reactor for water purification. Chemical Engineering Science, 59,

5249-5259.

Page 169: 1.3 Light emitting diode

150

ELLIS, D. A. & MABURY, S. A. 2000. The Aqueous Photolysis of TFM and Related

Trifluoromethylphenols. An Alternate Source of Trifluoroacetic Acid in the

Environment. Environmental Science & Technology, 34, 632-637.

ERICKSON, M. D. 1997. Analytical chemistry of PCBs, Boca Raton, FL., CRC/Lewis

Publishers.

EXON, J. H. 1984. A review of chlorinated phenols. Veterinary and Human Toxicology,

26, 508-20.

FIEDLER, H. 2007. National PCDD/PCDF release inventories under the Stockholm

Convention on Persistent Organic Pollutants. Chemosphere, 67, S96-S108.

FINLAYSON-PITTS, B. J. & PITTS JR, J. N. 1986. Atmospheric chemistry.

Fundamentals and experimental techniques, New York, Wiley.

FOO, K. Y. & HAMEED, B. H. 2010. Detoxification of pesticide waste via activated

carbon adsorption process. Journal of Hazardous Materials, 175, 1-11.

FOX, M. A. & DULAY, M. T. 1993. Heterogeneous photocatalysis. Chemical Reviews,

93, 341-357.

FUJISHIMA, A., RAO, T. N. & TRYK, D. A. 2000. Titanium dioxide photocatalysis.

Journal of Photochemistry and Photobiology C: Photochemistry Reviews, 1, 1-21.

GAL, E., AIRES, P., CHAMARRO, E. & ESPLUGAS, S. 1992. Photochemical

degradation of parathion in aqueous solutions. Water Research, 26, 911-915.

GALADI, A., BITAR, H., CHANON, M. & JULLIARD, M. 1995. Photosensitized

reductive dechlorination of chloroaromatic pesticides. Chemosphere, 30, 1655-

1669.

Page 170: 1.3 Light emitting diode

151

GALADI, A. & JULLIARD, M. 1996. Photosensitized oxidative degradation of

pesticides. Chemosphere, 33, 1-15.

GARABRANT, D. H. & PHILBERT, M. A. 2002. Review of 2,4-Dichlorophenoxyacetic

Acid (2,4-D) Epidemiology and Toxicology. Critical Reviews in Toxicology, 32,

233-257.

GARCIA, J. C. & TAKASHIMA, K. 2003. Photocatalytic degradation of imazaquin in

an aqueous suspension of titanium dioxide. Journal of Photochemistry and

Photobiology A: Chemistry, 155, 215-222.

GAYA, U. I. & ABDULLAH, A. H. 2008. Heterogeneous photocatalytic degradation of

organic contaminants over titanium dioxide: A review of fundamentals, progress

and problems. Journal of Photochemistry and Photobiology C: Photochemistry

Reviews, 9, 1-12.

GENG, Q. & CUI, W. 2010. Adsorption and Photocatalytic Degradation of Reactive

Brilliant Red K-2BP by TiO2/AC in Bubbling Fluidized Bed Photocatalytic

Reactor. Industrial & Engineering Chemistry Research, 49, 11321-11330.

GHOSH, J. P., LANGFORD, C. H. & ACHARI, G. 2008. Characterization of an LED

Based Photoreactor to Degrade 4-Chlorophenol in an Aqueous Medium Using

Coumarin (C-343) Sensitized TiO2. The Journal of Physical Chemistry A, 112,

10310-10314.

GHOSH, J. P., SUI, R., LANGFORD, C. H., ACHARI, G. & BERLINGUETTE, C. P.

2009. A comparison of several nanoscale photocatalysts in the degradation of a

common pollutant using LEDs and conventional UV light. Water Research, 43,

4499-4506.

Page 171: 1.3 Light emitting diode

152

GIERTHY, J. F., ARCARO, K. F. & FLOYD, M. 1997. Assessment of PCB

estrogenicity in a human breast cancer cell line. Chemosphere, 34, 1495-1505.

GONÇALVES, M. S. T., OLIVEIRA-CAMPOS, A. M. F., PINTO, E. M. M. S.,

PLASÊNCIA, P. M. S. & QUEIROZ, M. J. R. P. 1999. Photochemical treatment

of solutions of azo dyes containing TiO2. Chemosphere, 39, 781-786.

GONG, D., GRIMES, C., VARGHESE, O. K., HU, W., SINGH, R., CHEN, Z. &

DICKEY, E. C. 2001. Titanium oxide nanotube arrays prepared by anodic

oxidation. Journal of Materials Research, 16, 3331-3334.

HAARSTRICK, A., KUT, O. M. & HEINZLE, E. 1996. TiO2-Assisted Degradation of

Environmentally Relevant Organic Compounds in Wastewater Using a Novel

Fluidized Bed Photoreactor. Environmental Science & Technology, 30, 817-824.

HAMILL, N. A., WEATHERLEY, L. R. & HARDACRE, C. 2001. Use of a batch

rotating photocatalytic contactor for the degradation of organic pollutants in

wastewater. Applied Catalysis B: Environmental, 30, 49-60.

HARDELL, L. 1981. Relation of soft-tissue sarcoma, malignant lymphoma and colon

cancer to phenoxy acids, chlorophenols and other agents. Scand. J. Work Environ.

Health, 7, 119-110.

HAWARI, J., DEMETER, A., GREER, C. & SAMSON, R. 1991. Acetone-induced

photodechlorination of Aroclor 1254 in alkaline 2-propanol: Probing the

mechanism by thermolysis in the presence of di-t- butyl peroxide. Chemosphere,

22, 1161-1174.

HAWARI, J., DEMETER, A. & SAMSON, R. 1992. Sensitized photolysis of

polychlorobiphenyls in alkaline 2-propanol: dechlorination of Aroclor 1254 in

Page 172: 1.3 Light emitting diode

153

soil samples by solar radiation. Environmental Science & Technology, 26, 2022-

2027.

HEALTH CANADA. 2008. Chlorophenols [Online]. Available: http://www.hc-

sc.gc.ca/ewh-semt/pubs/water-eau/chlorophenols/index-eng.php [Accessed March

11 2012].

HEALTH CANADA 2010 Guidelines for Canadian Drinking Water Quality-Guideline

Technical Document: 2-Methyl-4-chlorophenoxyacetic Acid (MCPA).

HEALTH CANADA. 2012. Guidelines for Canadian Drinking Water Quality - Summary

Table [Online]. Available: http://www.hc-sc.gc.ca/ewh-semt/pubs/water-

eau/2012-sum_guide-res_recom/index-eng.php#t2 [Accessed March 4 2013].

HERRMANN, J.-M. 1999. Heterogeneous photocatalysis: fundamentals and applications

to the removal of various types of aqueous pollutants. Catalysis Today, 53, 115-

129.

HERRMANN, J. M. 2005. Heterogeneous photocatalysis: state of the art and present

applications In honor of Pr. R.L. Burwell Jr. (1912–2003), Former Head of

Ipatieff Laboratories, Northwestern University, Evanston (Ill). Topics in

Catalysis, 34, 49-65.

HERWEH, J. E. & HOYLE, C. E. 1980. Photodegradation of some alkyl N-

arylcarbamates. The Journal of Organic Chemistry, 45, 2195-2201.

HILLGREN, A., LINDGREN, J. & ALDÉN, M. 2002. Protection mechanism of Tween

80 during freeze–thawing of a model protein, LDH. International Journal of

Pharmaceutics, 237, 57-69.

Page 173: 1.3 Light emitting diode

154

HOFFMANN, M. R., MARTIN, S. T., CHOI, W. & BAHNEMANN, D. W. 1995.

Environmental Applications of Semiconductor Photocatalysis. Chemical Reviews,

95, 69-96.

HUANG, X., WANG, H., YIN, S., CHEN, X., CHEN, W. & YANG, H. 2009.

Sterilization system for air purifier by combining ultraviolet light emitting diodes

with TiO2. Journal of Chemical Technology and Biotechnology, 84, 1437-1440.

HULSTROM, R., BIRD, R. & RIORDAN, C. 1985. Spectral solar irradiance data sets

for selected terrestrial conditions. Solar Cells, 15, 365-391.

HUMBURG, N. E. (ed.) 1989. Herbicide handbook of the weed science society of

america, Champaign, Illinois: Weed Science Society of America.

HURUM, D. C., AGRIOS, A. G., GRAY, K. A., RAJH, T. & THURNAUER, M. C.

2003. Explaining the Enhanced Photocatalytic Activity of Degussa P25 Mixed-

Phase TiO2 Using EPR. The Journal of Physical Chemistry B, 107, 4545-4549.

HUSSAIN, S., SIDDIQUE, T., ARSHAD, M. & SALEEM, M. 2009. Bioremediation

and Phytoremediation of Pesticides: Recent Advances. Critical Reviews in

Environmental Science and Technology, 39, 843-907.

IARC 1987. Overall evaluations of carcinogenicity: an updating of IARC monographs

volumes 1 to 42, World Health Organization.

IMOBERDORF, G. E., CASSANO, A. E., IRAZOQUI, H. A. & ALFANO, O. M. 2007.

Optimal design and modeling of annular photocatalytic wall reactors. Catalysis

Today, 129, 118-126.

Page 174: 1.3 Light emitting diode

155

IMOBERDORF, G. E., TAGHIPOUR, F., KESHMIRI, M. & MOHSENI, M. 2008a.

Predictive radiation field modeling for fluidized bed photocatalytic reactors.

Chemical Engineering Science, 63, 4228-4238.

IMOBERDORF, G. E., TAGHIPOUR, F. & MOHSENI, M. 2008b. Radiation field

modeling of multi-lamp, homogeneous photoreactors. Journal of Photochemistry

and Photobiology A: Chemistry, 198, 169-178.

IRAZOQUI, H. A., CERDÁ, J. & CASSANO, A. E. 1973. Radiation profiles in an empty

annular photoreactor with a source of finite spatial dimensions. Aiche Journal, 19,

460-467.

IZADIFARD, M., ACHARI, G. & LANGFORD, C. H. 2008. The pathway of

dechlorination of PCB congener by a photochemical chain process in 2-propanol:

The role of medium and quenching. Chemosphere, 73, 1328-1334.

IZADIFARD, M., LANGFORD, C. H. & ACHARI, G. 2010a. Photocatalytic

dechlorination of PCB 138 using leuco-methylene blue and visible light; reaction

conditions and mechanisms. Journal of Hazardous materials, 181, 393-398.

IZADIFARD, M., LANGFORD, C. H. & ACHARI, G. 2010b. Photocatalytic

Dechlorination of Polychlorinated Biphenyls Using Leuco-methylene Blue

Sensitization, Broad Spectrum Visible Lamps, or Light Emitting Diodes.

Environmental Science & Technology, 44, 9075-9079.

JACOB, S. M. & DRANOFF, J. S. 1969. Light intensity profiles in an elliptical

photoreactor. Aiche Journal, 15, 141-144.

JACOBM, S. M. & DRANOFF, J. S. 1970. Light intensity profiles in a perfectly mixed

photoreactor. Aiche Journal, 16, 359-363.

Page 175: 1.3 Light emitting diode

156

JAKHER, A., ACHARI, G. & LANGFORD, C. H. 2007. Photodechlorination of Aroclor

1254 in a Pilot-Scale Flow through Photoreactor. Journal of Environmental

Engineering, 133, 646-654.

JOHNSON, M. B. & MEHRVAR, M. 2008. Aqueous Metronidazole Degradation by

UV/H2O2 Process in Single-and Multi-Lamp Tubular Photoreactors: Kinetics and

Reactor Design. Industrial and Engineering Chemistry Research, 47, 6525-6537.

KALIWOH, N., ZHANG, J.-Y. & BOYD, I. W. 2002. Characterisation of TiO2

deposited by photo-induced chemical vapour deposition. Applied Surface Science,

186, 241-245.

KAMRIN, M. A. 2000. Pesticide profiles, New York, Lewis Publishers.

KANECO, S., LI, N., ITOH, K.-K., KATSUMATA, H., SUZUKI, T. & OHTA, K. 2009.

Titanium dioxide mediated solar photocatalytic degradation of thiram in aqueous

solution: Kinetics and mineralization. Chemical Engineering Journal, 148, 50-56.

KANKI, T., HAMASAKI, S., SANO, N., TOYODA, A. & HIRANO, K. 2005. Water

purification in a fluidized bed photocatalytic reactor using TiO2-coated ceramic

particles. Chemical Engineering Journal, 108, 155-160.

KENNEPOHL, E., MUNRO, I. C. & BUS, J. S. 2010. Chapter 84 - Phenoxy Herbicides

(2,4-D). In: KRIEGER, R. (ed.) Hayes' Handbook of Pesticide Toxicology (Third

Edition). New York: Academic Press.

KHAN, A., BALAKRISHNAN, K. & KATONA, T. 2008. Ultraviolet light-emitting

diodes based on group three nitrides. Nat Photon, 2, 77-84.

Page 176: 1.3 Light emitting diode

157

KIM, S. B. & HONG, S. C. 2002. Kinetic study for photocatalytic degradation of volatile

organic compounds in air using thin film TiO2 photocatalyst. Applied Catalysis

B: Environmental, 35, 305-315.

KIMURA, K., YOKOYAMA, K., SATO, H., NORDIN, R. B., NAING, L., KIMURA,

S., OKABE, S., MAENO, T., KOBAYASHI, Y., KITAMURA, F. & ARAKI, S.

2005. Effects of Pesticides on the Peripheral and Central Nervous System in

Tobacco Farmers in Malaysia Studies on Peripheral Nerve Conduction, Brain-

Evoked Potentials and Computerized Posturography. Industrial Health, 43, 285-

294.

KIPSHIDZE, G., KURYATKOV, V., BORISOV, B., HOLTZ, M., NIKISHIN, S. &

TEMKIN, H. 2002. AlGaInN-based ultraviolet light-emitting diodes grown on Si

(111). Applied Physics Letters, 80, 3682-3684.

KIPSHIDZE, G., KURYATKOV, V., ZHU, K., BORISOV, B., HOLTZ, M., NIKISHIN,

S. & TEMKIN, H. 2003. AlN/AlGaInN superlattice light-emitting diodes at 280

nm. Journal of Applied Physics, 93, 1363-1366.

KNEISSL, M., KOLBE, T., CHUA, C., KUELLER, V., LOBO, N., STELLMACH, J.,

KNAUER, A., RODRIGUEZ, H., EINFELDT, S., YANG, Z., JOHNSON, N. M.

& WEYERS, M. 2011. Advances in group III-nitride-based deep UV light-

emitting diode technology. Semiconductor Science and Technology, 26, 014036.

KONSTANTINOU, I. K. & ALBANIS, T. A. 2004. TiO2-assisted photocatalytic

degradation of azo dyes in aqueous solution: kinetic and mechanistic

investigations: A review. Applied Catalysis B: Environmental, 49, 1-14.

Page 177: 1.3 Light emitting diode

158

KUMAR, K. V., PORKODI, K. & ROCHA, F. 2008. Langmuir–Hinshelwood kinetics –

A theoretical study. Catalysis Communications, 9, 82-84.

KWAN, C. Y. & CHU, W. 2004. A study of the reaction mechanisms of the degradation

of 2,4-dichlorophenoxyacetic acid by oxalate-mediated photooxidation. Water

Research, 38, 4213-4221.

LASA, H., SERRANO, B. & SALAICES, M. 2005. Novel Photocatalytic Reactors for

Water and Air Treatment. Photocatalytic Reaction Engineering. Springer US.

LAWS, E. R. & HAYES, W. J. 1991. Handbook of Pesticide Toxicology, New York,

Academic press.

LEA, J. & ADESINA, A. A. 1998. The photo-oxidative degradation of sodium dodecyl

sulphate in aerated aqueous TiO2 suspension. Journal of Photochemistry and

Photobiology A: Chemistry, 118, 111-122.

LEDS MAGAZINE. 2012. SemiLEDs achieves 40% external quantum efficiency for

ultraviolet LED chips [Online]. Available:

http://www.ledsmagazine.com/ugc/2012/02/semileds-achieves-40-external-

quantum-efficiency-for-ultraviolet-led-chips.html [Accessed August 8 2013].

LI, G., LIU, Z.-Q., LU, J., WANG, L. & ZHANG, Z. 2009. Effect of calcination

temperature on the morphology and surface properties of TiO2 nanotube arrays.

Applied Surface Science, 255, 7323-7328.

LI, Y., YU, H., ZHANG, C., SONG, W., LI, G., SHAO, Z. & YI, B. 2013. Effect of

water and annealing temperature of anodized TiO2 nanotubes on hydrogen

production in photoelectrochemical cell. Electrochimica Acta, 107, 313-319.

Page 178: 1.3 Light emitting diode

159

LIN, Y., GUPTA, G. & BAKER, J. 1996. Photodegradation of Aroclor 1254 Using

Simulated Sunlight and Various Sensitizers. Bulletin of Environmental

Contamination and Toxicology, 56, 566-570.

LINSEBIGLER, A. L., LU, G. & YATES, J. T. 1995. Photocatalysis on TiO2 Surfaces:

Principles, Mechanisms, and Selected Results. Chemical Reviews, 95, 735-758.

LIU, W., CHEN, S., ZHAO, W. & ZHANG, S. 2009. Study on the photocatalytic

degradation of trichlorfon in suspension of titanium dioxide. Desalination, 249,

1288-1293.

MACAK, J. M., TSUCHIYA, H., TAVEIRA, L., ALDABERGEROVA, S. &

SCHMUKI, P. 2005. Smooth anodic TiO2 nanotubes. Angewandte Chemie

International Edition, 44, 7463-7465.

MANCHESTER-NEESVIG, J. B., ANDREN, A. W. & EDGINGTON, D. N. 1996.

Patterns of Mass Sedimentation and of Deposition of Sediment Contaminated by

PCBs in Green Bay. Journal of Great Lakes Research, 22, 444-462.

MANDAL, A. B., NAIR, B. U. & RAMASWAMY, D. 1988. Determination of the

critical micelle concentration of surfactants and the partition coefficient of an

electrochemical probe by using cyclic voltammetry. Langmuir, 4, 736-739.

MARUGÁN, J., VAN GRIEKEN, R., ALFANO, O. M. & CASSANO, A. E. 2006.

Optical and physicochemical properties of silica-supported TiO2 photocatalysts.

AIChE Journal, 52, 2832-2843.

MATTHEWS, R. W. 1987. Solar-electric water purification using photocatalytic

oxidation with TiO2 as a stationary phase. Solar Energy, 38, 405-413.

Page 179: 1.3 Light emitting diode

160

MATTHEWS, R. W. 1988. Kinetics of photocatalytic oxidation of organic solutes over

titanium dioxide. Journal of Catalysis, 111, 264-272.

MATTHEWS, R. W. 1991. Photooxidative degradation of coloured organics in water

using supported catalysts. TiO2 on sand. Water Research, 25, 1169-1176.

MATTHEWS, R. W. & MCEVOY, S. R. 1992. Destruction of phenol in water with sun,

sand, and photocatalysis. Solar Energy, 49, 507-513.

MEHROTRA, K., YABLONSKY, G. S. & RAY, A. K. 2005. Macro kinetic studies for

photocatalytic degradation of benzoic acid in immobilized systems. Chemosphere,

60, 1427-1436.

MEHRVAR, M., ANDERSON, W. A. & MOO-YOUNG, M. 2001. Photocatalytic

degradation of aqueous organic solvents in the presence of hydroxyl radical

scavengers. International Journal of Photoenergy, 3, 187-191.

MERGEL, D., BUSCHENDORF, D., EGGERT, S., GRAMMES, R. & SAMSET, B.

2000. Density and refractive index of TiO2 films prepared by reactive

evaporation. Thin Solid Films, 371, 218-224.

MIAO, X.-S., CHU, S.-G. & XU, X.-B. 1999. Degradation pathways of PCBs upon UV

irradiation in hexane. Chemosphere, 39, 1639-1650.

MIAO, X. S., CHU, S. G. & XU, X. B. 1996. Photodegradation of 2,2′,5,5′-

Tetrachlorobiphenyl in Hexane. Bulletin of Environmental Contamination and

Toxicology, 56, 571-574.

MIILLE, M. J. & CROSBY, D. G. 1983. Pentachlorophenol and 3,4-dichloroaniline as

models for photochemical reactions in seawater. Marine Chemistry, 14, 111-120.

Page 180: 1.3 Light emitting diode

161

MIKULA, M., BREZOVÁ, V., CĚPPAN, M., PACH, L. & KARPINSKÝ, Ĺ. 1995.

Comparison of photocatalytic activity of sol-gel TiO2 and P25 TiO2 particles

supported on commercial fibreglass fabric. Journal of Materials Science Letters,

14, 615-616.

MILLS, A. & MORRIS, S. 1993. Photomineralization of 4-chlorophenol sensitized by

titanium dioxide: a study of the initial kinetics of carbon dioxide photogeneration.

Journal of Photochemistry and Photobiology A: Chemistry, 71, 75-83.

MIRANDA-GARCÍA, N., SUÁREZ, S., SÁNCHEZ, B., CORONADO, J., MALATO,

S. & MALDONADO, M. I. 2011. Photocatalytic degradation of emerging

contaminants in municipal wastewater treatment plant effluents using

immobilized TiO2 in a solar pilot plant. Applied Catalysis B: Environmental, 103,

294-301.

MISHIMA, O., ERA, K., TANAKA, J. & YAMAOKA, S. 1988. Ultraviolet

light‐emitting diode of a cubic boron nitride pn junction made at high pressure.

Applied Physics Letters, 53, 962-964.

MO, J., ZHANG, Y., YANG, R. & XU, Q. 2008. Influence of fins on formaldehyde

removal in annular photocatalytic reactors. Building and Environment, 43, 238-

245.

MOZIA, S. 2010. Photocatalytic membrane reactors (PMRs) in water and wastewater

treatment. A review. Separation and Purification Technology, 73, 71-91.

MUNEER, M. & BOXALL, C. 2008. Photocatalyzed Degradation of a Pesticide

Derivative Glyphosate in Aqueous Suspensions of Titanium Dioxide.

International Journal of Photoenergy, 7.

Page 181: 1.3 Light emitting diode

162

MUNEER, M., QAMAR, M., SAQUIB, M. & BAHNEMANN, D. W. 2005.

Heterogeneous photocatalysed reaction of three selected pesticide derivatives,

propham, propachlor and tebuthiuron in aqueous suspensions of titanium dioxide.

Chemosphere, 61, 457-468.

MURUGANANDHAM, M. & SWAMINATHAN, M. 2006. Photocatalytic

decolourisation and degradation of Reactive Orange 4 by TiO2-UV process. Dyes

and Pigments, 68, 133-142.

NAKAMURA, M., KATO, S., AOKI, T., SIRGHI, L. & HATANAKA, Y. 2001.

Formation mechanism for TiOx thin film obtained by remote plasma enhanced

chemical vapor deposition in H2–O2 mixture gas plasma. Thin Solid Films, 401,

138-144.

NATARAJAN, M. R., WU, W. M., NYE, J., WANG, H., BHATNAGAR, L. & JAIN,

M. K. 1996. Dechlorination of polychlorinated biphenyl congeners by an

anaerobic microbial consortium. Applied Microbiology and Biotechnology, 46,

673-677.

NATARAJAN, T. S., NATARAJAN, K., BAJAJ, H. C. & TAYADE, R. J. 2011a.

Energy Efficient UV-LED Source and TiO2 Nanotube Array-Based Reactor for

Photocatalytic Application. Industrial & Engineering Chemistry Research, 50,

7753-7762.

NATARAJAN, T. S., THOMAS, M., NATARAJAN, K., BAJAJ, H. C. & TAYADE, R.

J. 2011b. Study on UV-LED/TiO2 process for degradation of Rhodamine B dye.

Chemical Engineering Journal, 169, 126-134.

Page 182: 1.3 Light emitting diode

163

NEGISHI, N. & TAKEUCHI, K. 2001. Preparation of TiO2 Thin Film Photocatalysts by

Dip Coating Using a Highly Viscous Solvent. Journal of Sol-Gel Science and

Technology, 22, 23-31.

NEGISHI, N., TAKEUCHI, K. & IBUSUKI, T. 1998. Preparation of the TiO2 Thin Film

Photocatalyst by the Dip-Coating Process. Journal of Sol-Gel Science and

Technology, 13, 691-694.

NGUYEN, T. V. & WU, J. C. S. 2008. Photoreduction of CO2 in an optical-fiber

photoreactor: Effects of metals addition and catalyst carrier. Applied Catalysis A:

General, 335, 112-120.

NICHIA. 2013. Specifications for UV LED-NCSU033B(T) [Online]. Available:

www.nichia.co.jp/specification/products/led/NCSU033B-E.pdf [Accessed June 6

2013].

OBEE, T. N. & BROWN, R. T. 1995. TiO2 Photocatalysis for Indoor Air Applications:

Effects of Humidity and Trace Contaminant Levels on the Oxidation Rates of

Formaldehyde, Toluene, and 1,3-Butadiene. Environmental Science &

Technology, 29, 1223-1231.

OHKO, Y., HASHIMOTO, K. & FUJISHIMA, A. 1997. Kinetics of Photocatalytic

Reactions under Extremely Low-Intensity UV Illumination on Titanium Dioxide

Thin Films. The Journal of Physical Chemistry A, 101, 8057-8062.

OHTANI, B. 2008a. Preparing articles on photocatalysis—beyond the illusions,

misconceptions, and speculation. Chemistry Letters, 37, 216-229.

OHTANI, B. 2008b. Preparing Articles on Photocatalysis&mdash;Beyond the Illusions,

Misconceptions, and Speculation. Chemistry Letters, 37, 216-229.

Page 183: 1.3 Light emitting diode

164

OHTANI, B., PRIETO-MAHANEY, O. O., LI, D. & ABE, R. 2010. What is Degussa

(Evonik) P25? Crystalline composition analysis, reconstruction from isolated pure

particles and photocatalytic activity test. Journal of Photochemistry and

Photobiology A: Chemistry, 216, 179-182.

OKAMURA, H., AOYAMA, I., LIU, D., MAGUIRE, J., PACEPAVICIUS, G. J. &

LAU, Y. L. 1999. Photodegradation of Irgarol 1051 in water. Journal of

Environmental Science and Health, Part B, 34, 225-238.

OLLIS, D. F. 2005. Kinetics of Liquid Phase Photocatalyzed Reactions:  An Illuminating

Approach†. The Journal of Physical Chemistry B, 109, 2439-2444.

OLLIS, D. F., PELIZZETTI, E. & SERPONE, N. 1991. Photocatalyzed destruction of

water contaminants. Environmental Science & Technology, 25, 1522-1529.

OPPENLANDER, T. 2003. Photochemical Purification of Water and Air,

Winheim,Germany, WILLY-VCH.

PAREDES, S., TRIBOUT, M. & SEPULVEDA, L. 1984. Enthalpies of micellization of

the quaternary tetradecyl- and -cetyl ammonium salts. The Journal of Physical

Chemistry, 88, 1871-1875.

PAREEK, V., CHONG, S., TADÉ, M. & ADESINA, A. A. 2008. Light intensity

distribution in heterogenous photocatalytic reactors. Asia-Pacific Journal of

Chemical Engineering, 3, 171-201.

PARSONS, S. (ed.) 2004. Advanced Oxidation Processes for Water and Wastewater

Treatment, London, UK: IWA Publishing.

PASQUALI, M., SANTARELLI, F., PORTER, J. F. & YUE, P.-L. 1996. Radiative

transfer in photocatalytic systems. AIChE Journal, 42, 532-537.

Page 184: 1.3 Light emitting diode

165

PAULOSE, M., SHANKAR, K., YORIYA, S., PRAKASAM, H. E., VARGHESE, O.

K., MOR, G. K., LATEMPA, T. A., FITZGERALD, A. & GRIMES, C. A. 2006.

Anodic Growth of Highly Ordered TiO2 Nanotube Arrays to 134 μm in Length.

The Journal of Physical Chemistry B, 110, 16179-16184.

PELTON, R., GENG, X. & BROOK, M. 2006. Photocatalytic paper from colloidal

TiO2—fact or fantasy. Advances in Colloid and Interface Science, 127, 43-53.

PERAL, J. & OLLIS, D. F. 1992. Heterogeneous photocatalytic oxidation of gas-phase

organics for air purification: Acetone, 1-butanol, butyraldehyde, formaldehyde,

and m-xylene oxidation. Journal of Catalysis, 136, 554-565.

PETROVIC, M., RADJENOVIC, J., POSTIGO, C., KUSTER, M., FARRE, M., ALDA,

M. & BARCELÓ, D. 2008. Emerging Contaminants in Waste Waters: Sources

and Occurrence. In: BARCELÓ, D. & PETROVIC, M. (eds.) Emerging

Contaminants from Industrial and Municipal Waste. Springer Berlin Heidelberg.

PIZARRO, P., GUILLARD, C., PEROL, N. & HERRMANN, J. M. 2005. Photocatalytic

degradation of imazapyr in water: Comparison of activities of different supported

and unsupported TiO2-based catalysts. Catalysis Today, 101, 211-218.

POLYRAKIS, I. T. 2009. Environmental Pollution from Pesticides. In: COSTA, R. &

KRISTBERGSSON, K. (eds.) Predictive Modeling and RiskAssessment. Springer

US.

PORTIER, R. J., ZOELLER, A. L. & FUJISAKI, K. 1990. Bioremediation of pesticide-

contaminated groundwater. Remediation Journal, 1, 41-60.

POZZO, R. L., BRANDI, R. J., GIOMBI, J. L., BALTANÁS, M. A. & CASSANO, A.

E. 2005. Design of fluidized bed photoreactors: Optical properties of

Page 185: 1.3 Light emitting diode

166

photocatalytic composites of titania CVD-coated onto quartz sand. Chemical

Engineering Science, 60, 2785-2794.

QAMAR, M., MUNEER, M. & BAHNEMANN, D. 2006. Heterogeneous photocatalysed

degradation of two selected pesticide derivatives, triclopyr and daminozid in

aqueous suspensions of titanium dioxide. Journal of Environmental Management,

80, 99-106.

RAHMAN, A. M., QAMAR, M., MUNEER, M. & BAHNEMANN, D. 2006.

Semiconductor mediated photocatalysed degradation of a pesticide derivative,

acephate in aqueous suspensions of titanium dioxide. Journal of Advanced

Oxidation Technologies, 9, 103-109.

RAHMAN, M. A. & MUNEER, M. 2005. Photocatalysed degradation of two selected

pesticide derivatives, dichlorvos and phosphamidon, in aqueous suspensions of

titanium dioxide. Desalination, 181, 161-172.

RAY, A. K. & BEENACKERS, A. A. C. M. 1998. Development of a new photocatalytic

reactor for water purification. Catalysis Today, 40, 73-83.

ROSELIN, L. S. & SELVIN, R. 2011. Photocatalytic Degradation of Reactive Orange 16

Dye in a ZnO Coated Thin Film Flow Photoreactor. Science of Advanced

Materials, 3, 251-258.

ROY, P., BERGER, S. & SCHMUKI, P. 2011. TiO2 Nanotubes: Synthesis and

Applications. Angewandte Chemie International Edition, 50, 2904-2939.

RUZO, L. O., ZABIK, M. J. & SCHUETZ, R. D. 1974. Photochemistry of bioactive

compounds. Photochemical processes of polychlorinated biphenyls. Journal of the

American Chemical Society, 96, 3809-3813.

Page 186: 1.3 Light emitting diode

167

RYER, A. & LIGHT, V. 1997. Light measurement handbook, Newburyport,

Massachusetts, International Light Inc.

SAFE, S. H. 1994. Polychlorinated Biphenyls (PCBs): Environmental Impact,

Biochemical and Toxic Responses, and Implications for Risk Assessment.

Critical Reviews in Toxicology, 24, 87-149.

SAKTHIVEL, S., NEPPOLIAN, B., SHANKAR, M. V., ARABINDOO, B.,

PALANICHAMY, M. & MURUGESAN, V. 2003. Solar photocatalytic

degradation of azo dye: comparison of photocatalytic efficiency of ZnO and

TiO2. Solar Energy Materials and Solar Cells, 77, 65-82.

SALVADÓ-ESTIVILL, I., HARGREAVES, D. M. & LI PUMA, G. 2007. Evaluation of

the Intrinsic Photocatalytic Oxidation Kinetics of Indoor Air Pollutants.

Environmental Science & Technology, 41, 2028-2035.

SAMANIDOU, V., FYTIANOS, K., PFISTER, G. & BAHADIR, M. 1988.

Photochemical decomposition of carbamate pesticides in natural waters of

northern Greece. Science of the Total Environment, 76, 85-92.

SAQUIB, M. & MUNEER, M. 2003. TiO2-mediated photocatalytic degradation of a

triphenylmethane dye (gentian violet), in aqueous suspensions. Dyes and

Pigments, 56, 37-49.

SCHUBERT, E. F. 2006. Light-Emitting Diodes, Cambridge UK, Cambridge University

Press.

SEMICONDUCTOR TODAY 2013. Market focus: UV LEDs. Semiconductor Today

Compouns&AdvancedSilicon.

Page 187: 1.3 Light emitting diode

168

SERRANO, B. & DE LASA, H. 1997. Photocatalytic Degradation of Water Organic

Pollutants. Kinetic Modeling and Energy Efficiency. Industrial and Engineering

Chemistry Research, 36, 4705-4711.

SHIE, J.-L., LEE, C.-H., CHIOU, C.-S., CHANG, C.-T., CHANG, C.-C. & CHANG, C.-

Y. 2008. Photodegradation kinetics of formaldehyde using light sources of UVA,

UVC and UVLED in the presence of composed silver titanium oxide

photocatalyst. Journal of Hazardous materials, 155, 164-172.

SHOURONG, Z., QINGGUO, H., JUN, Z. & BINGKUN, W. 1997. A study on dye

photoremoval in TiO2 suspension solution. Journal of Photochemistry and

Photobiology A: Chemistry, 108, 235-238.

SINGH, H., MUNEER, M. & BAHNEMANN, D. 2003. Photocatalysed degradation of a

herbicide derivative, bromacil, in aqueous suspensions of titanium dioxide.

Photochemical and Photobiological Sciences, 2, 151-156.

SINGH, H. K. & MUNEER, M. 2004. Photodegradation of a herbicide derivative, 2,4-

dichlorophenoxy acetic acid in aqueous suspensions of titanium dioxide. Research

on Chemical Intermediates, 30, 317-329.

SINGH, H. K., SAQUIB, M., HAQUE, M. M., MUNEER, M. & BAHNEMANN, D. W.

2007. Titanium dioxide mediated photocatalysed degradation of phenoxyacetic

acid and 2,4,5-trichlorophenoxyacetic acid, in aqueous suspensions. Journal of

Molecular Catalysis A: Chemical, 264, 66-72.

SO, C. M., CHENG, M. Y., YU, J. C. & WONG, P. K. 2002. Degradation of azo dye

Procion Red MX-5B by photocatalytic oxidation. Chemosphere, 46, 905-912.

Page 188: 1.3 Light emitting diode

169

SOTELO, J. L., OVEJERO, G., DELGADO, J. A. & MARTı́NEZ, I. 2002. Comparison

of adsorption equilibrium and kinetics of four chlorinated organics from water

onto GAC. Water Research, 36, 599-608.

STANGROOM, S. J., MACLEOD, C. L. & LESTER, J. N. 1998. Photosensitized

transformation of the herbicide 4-chloro-2-methylphenoxy acetic acid (MCPA) in

water. Water Research, 32, 623-632.

SULLIVAN, J. R., DELFINO, J. J., BUELOW, C. R. & SHEFFY, T. B. 1983.

Polychlorinated biphenyls in the fish and sediment of the Lower Fox River,

Wisconsin. Bulletin of Environmental Contamination and Toxicology, 30, 58-64.

TADEO, J. L. 2008. Analysis of pesticides in food and environmental samples, New

York, CRC Press.

TANG, C. & CHEN, V. 2004. The photocatalytic degradation of reactive black 5 using

TiO2/UV in an annular photoreactor. Water Research, 38, 2775-2781.

TANIYASU, Y., KASU, M. & MAKIMOTO, T. 2006. An aluminium nitride light-

emitting diode with a wavelength of 210 nanometres. Nature, 441, 325-328.

THEURICH, J., LINDNER, M. & BAHNEMANN, D. W. 1996. Photocatalytic

Degradation of 4-Chlorophenol in Aerated Aqueous Titanium Dioxide

Suspensions:  A Kinetic and Mechanistic Study. Langmuir, 12, 6368-6376.

THIRUVENKATACHARI, R., OUK KWON, T. & SHIK MOON, I. 2005. Application

of Slurry Type Photocatalytic Oxidation‐Submerged Hollow Fiber Microfiltration

Hybrid System for the Degradation of Bisphenol A (BPA). Separation Science

and Technology, 40, 2871-2888.

Page 189: 1.3 Light emitting diode

170

THIRUVENKATACHARI, R., VIGNESWARAN, S. & MOON, I. 2008. A review on

UV/TiO&lt;sub&gt;2&lt;/sub&gt; photocatalytic oxidation process (Journal

Review). Korean Journal of Chemical Engineering, 25, 64-72.

TOPALOV, A., ABRAMOVIĆ, B., MOLNÁR-GÁBOR, D., CSANÁDI, J. & ARCSON,

O. 2001. Photocatalytic oxidation of the herbicide (4-chloro-2-

methylphenoxy)acetic acid (MCPA) over TiO2. Journal of Photochemistry and

Photobiology A: Chemistry, 140, 249-253.

TRILLAS, M., PERAL, J. & DOMÈNECH, X. 1995. Redox photodegradation of 2,4-

dichlorophenoxyacetic acid over TiO2. Applied Catalysis B: Environmental, 5,

377-387.

TSENG, D. H., JUANG, L. C. & HUANG, H. H. 2012. Effect of Oxygen and Hydrogen

Peroxide on the Photocatalytic Degradation of Monochlorobenzene in Aqueous

Suspension. International Journal of Photoenergy, 2012, 9.

TURCHI, C. S. & OLLIS, D. F. 1988. Comment. Photocatalytic reactor design: an

example of mass-transfer limitations with an immobilized catalyst. The journal of

Physical Chemistry, 92, 6852-6853.

TURRO, N. J. 1991. Modern Molecular Photochemistry, Sausalito,CA, University

Science Books.

USEPA 1980. Ambient Water Quality Criteria for Chlorinated Phenols.

Washingtong,D.C.

USEPA 1996a. USEPA Method 3540C soxhlet extraction. Washingtong, D.C.

USEPA 1996b. USEPA method 8082, polychlorinated biphenyls by gas chromatography.

Washingtong, D.C.

Page 190: 1.3 Light emitting diode

171

USEPA. 2000. Lakewide Management Plans [Online]. Available:

http://www.epa.gov/lakemich/ [Accessed April 1 2012].

USEPA 2004. Reregistration Eligibility Decision (RED) for MCPA (2-methyl-4-

chlorophenoxyacetic acid) List A Case 0017. Washington,D.C.

USEPA 2005. Reregistration Eligibility Decision for 2,4-D. Washington, D.C.

USEPA. 2011. About Pesticide [Online]. Available: http://www.epa.gov/pesticides/about/

[Accessed March 11 2011].

USEPA. 2012. Aroclor and Other PCB Mixtures [Online]. Available:

http://www.epa.gov/osw/hazard/tsd/pcbs/pubs/aroclor.htm [Accessed June 15

2012].

USEPA. 2013a. Health Effects of PCBs [Online]. Available:

http://www.epa.gov/osw/hazard/tsd/pcbs/pubs/effects.htm [Accessed December

16 2013].

USEPA. 2013b. Polychlorinated Biphenyls (PCBs)-Basic Information [Online].

Available: http://www.epa.gov/epawaste/hazard/tsd/pcbs/pubs/about.htm

[Accessed 12-16 2013].

VAISMAN, E., KABIR, M., KANTZAS, A. & LANGFORD, C. 2005. A fluidized bed

photoreactor exploiting a supported photocatalyst with adsorption pre-

concentration capacity. Journal of Applied Electrochemistry, 35, 675-681.

VEGA, A. A., IMOBERDORF, G. E. & MOHSENI, M. Photocatalytic degradation of

2,4-Dichlorophenoxyacetic acid in a fluidized bed photoreactor with composite

template-free TiO2 photocatalyst. Applied Catalysis A: General.

Page 191: 1.3 Light emitting diode

172

VELLANKI, B. P., BATCHELOR, B. & ABDEL-WAHAB, A. 2013. Advanced

reduction processes: a new class of treatment processes. Environmental

engineering science, 30, 264-271.

WANG, J. & LIN, Z. 2008. Freestanding TiO2 nanotube arrays with ultrahigh aspect

ratio via electrochemical anodization. Chemistry of Materials, 20, 1257-1261.

WANG, J. & LIN, Z. 2009. Anodic Formation of Ordered TiO2 Nanotube Arrays:

Effects of Electrolyte Temperature and Anodization Potential. The Journal of

Physical Chemistry C, 113, 4026-4030.

WANG, P. & KELLER, A. A. 2009. Partitioning of hydrophobic pesticides within a soil-

water-anionic surfactant system. Water Research, 43, 706-714.

WANG, W. Y. & KU, Y. 2006. Photocatalytic degradation of Reactive Red 22 in

aqueous solution by UV-LED radiation. Water Research, 40, 2249-2258.

WANG, Y. & HONG, C.-S. 2000. TiO2-mediated photomineralization of 2-

chlorobiphenyl: the role of O2. Water Research, 34, 2791-2797.

WANG, Z., LIU, J., DAI, Y., DONG, W., ZHANG, S. & CHEN, J. 2012. CFD modeling

of a UV-LED photocatalytic odor abatement process in a continuous reactor.

Journal of Hazardous materials, 215–216, 25-31.

WATANABE, A., TSUCHIYA, T. & IMAI, Y. 2002. Selective deposition of anatase and

rutile films by KrF laser chemical vapor deposition from titanium isopropoxide.

Thin Solid Films, 406, 132-137.

WATANABE, T., FUKAYAMA, S., MIYAUCHI, M., FUJISHIMA, A. &

HASHIMOTO, K. 2000. Photocatalytic Activity and Photo-Induced Wettability

Page 192: 1.3 Light emitting diode

173

Conversion of TiO2 Thin Film Prepared by Sol-Gel Process on a Soda-Lime

Glass. Journal of Sol-Gel Science and Technology, 19, 71-76.

WAYNE, R. P. 1988. Principles and Applications of Photochemistry New York, Oxford

Science publications.

WEI, L., SHIFU, C., WEI, Z. & SUJUAN, Z. 2009. Titanium dioxide mediated

photocatalytic degradation of methamidophos in aqueous phase. Journal of

Hazardous materials, 164, 154-160.

WIKIPEDIA. 2011. Light-emitting diode [Online]. Available:

http://en.wikipedia.org/wiki/File:PnJunction-LED-E.svg [Accessed Sep. 12

2011].

WILSON, R. I. & MABURY, S. A. 2000. Photodegradation of Metolachlor:  Isolation,

Identification, and Quantification of Monochloroacetic Acid. Journal of

Agricultural and Food Chemistry, 48, 944-950.

WORLD HEALTH ORGANIZATION 2004. Guidelines for drinking-water quality:

recommendations, World Health Organization.

WU, C. Y., CHIANG, B. S., CHANG, S. & LIU, D. S. 2011. Determination of

photocatalytic activity in amorphous and crystalline titanium oxide films prepared

using plasma-enhanced chemical vapor deposition. Applied Surface Science, 257,

1893-1897.

WU, R. J., CHEN, C. C., CHEN, M. H. & LU, C. S. 2009. Titanium dioxide-mediated

heterogeneous photocatalytic degradation of terbufos: Parameter study and

reaction pathways. Journal of Hazardous materials, 162, 945-953.

Page 193: 1.3 Light emitting diode

174

WU, R. J., CHEN, C. C., LU, C. S., HSU, P. Y. & CHEN, M. H. 2010. Phorate

degradation by TiO2 photocatalysis: Parameter and reaction pathway

investigations. Desalination, 250, 869-875.

WÜRTELE, M. A., KOLBE, T., LIPSZ, M., KÜLBERG, A., WEYERS, M., KNEISSL,

M. & JEKEL, M. 2011. Application of GaN-based ultraviolet-C light emitting

diodes – UV LEDs – for water disinfection. Water Research, 45, 1481-1489.

XIE, Y. B. & LI, X. Z. 2006. Preparation and characterization of TiO2/Ti film electrodes

by anodization at low voltage for photoelectrocatalytic application. Journal of

Applied Electrochemistry, 36, 663-668.

YAMAMOTO, S., SUMITA, T., SUGIHARUTO, MIYASHITA, A. & NARAMOTO,

H. 2001. Preparation of epitaxial TiO2 films by pulsed laser deposition technique.

Thin Solid Films, 401, 88-93.

YANG, H. G., SUN, C. H., QIAO, S. Z., ZOU, J., LIU, G., SMITH, S. C., CHENG, H.

M. & LU, G. Q. 2008. Anatase TiO2 single crystals with a large percentage of

reactive facets. Nature, 453, 638-641.

YAO, Y., KAKIMOTO, K., OGAWA, H. I., KATO, Y., HANADA, Y., SHINOHARA,

R. & YOSHINO, E. 1997. Photodechlorination Pathways of Non-Ortho

Substituted PCBs by Ultraviolet Irradiation in Alkaline 2-Propanol. Bulletin of

Environmental Contamination and Toxicology, 59, 238-245.

YAO, Y., KAKIMOTO, K., OGAWA, H. I., KATO, Y., KADOKAMI, K. &

SHINOHARA, R. 2000. Further study on the photochemistry of non-ortho

substituted PCBs by UV irradiation in alkaline 2-propanol. Chemosphere, 40,

951-956.

Page 194: 1.3 Light emitting diode

175

YASAN, A., MCCLINTOCK, R., MAYES, K., DARVISH, S., KUNG, P. & RAZEGHI,

M. 2002. Top-emission ultraviolet light-emitting diodes with peak emission at

280 nm. Applied Physics Letters, 81, 801-802.

YASAN, A., MCCLINTOCK, R., MAYES, K., SHIELL, D., GAUTERO, L.,

DARVISH, S., KUNG, P. & RAZEGHI, M. 2003. 4.5 mW operation of AlGaN-

based 267 nm deep-ultraviolet light-emitting diodes. Applied Physics Letters, 83,

4701-4703.

YING, G. G. & WILLIAMS, B. 1999. The degradation of oxadiazon and oxyfluorfen by

photolysis. Journal of Environmental Science and Health, Part B, 34, 549-567.

YOUNES, M. & GALAL-GORCHEV, H. 2000. Pesticides in drinking water—A case

study. Food and Chemical Toxicology, 38, Supplement 1, S87-S90.

YU, J., DAI, G. & CHENG, B. 2010. Effect of Crystallization Methods on Morphology

and Photocatalytic Activity of Anodized TiO2 Nanotube Array Films. The

Journal of Physical Chemistry C, 114, 19378-19385.

YU, J., ZHAO, X. & ZHAO, Q. 2001. Photocatalytic activity of nanometer TiO2 thin

films prepared by the sol–gel method. Materials Chemistry and Physics, 69, 25-

29.

YU, J. C., KWONG, T. Y., LUO, Q. & CAI, Z. 2006. Photocatalytic oxidation of

triclosan. Chemosphere, 65, 390-399.

YU, L., ACHARI, G. & LANGFORD, C. H. 2013. LED-Based Photocatalytic Treatment

of Pesticides and Chlorophenols. Journal of Environmental Engineering, 139,

1146-1151.

Page 195: 1.3 Light emitting diode

176

ZEMAN, P. & TAKABAYASHI, S. 2002. Effect of total and oxygen partial pressures on

structure of photocatalytic TiO2 films sputtered on unheated substrate. Surface

and Coatings Technology, 153, 93-99.

ZEPP, R. G. & CLINE, D. M. 1977. Rates of direct photolysis in aquatic environment.

Environmental Science & Technology, 11, 359-366.

ZERTAL, A., MOLNÁR-GÁBOR, D., MALOUKI, M. A., SEHILI, T. & BOULE, P.

2004. Photocatalytic transformation of 4-chloro-2-methylphenoxyacetic acid

(MCPA) on several kinds of TiO2. Applied Catalysis B: Environmental, 49, 83-

89.

ZHANG, J., LAN, W., QIAO, C. & JIANG, H. 2004. Bioremediation of

Organophosphorus Pesticides by Surface-Expressed Carboxylesterase from

Mosquito on Escherichia coli. Biotechnology Progress, 20, 1567-1571.

ZWILLING, V., AUCOUTURIER, M. & DARQUE-CERETTI, E. 1999. Anodic

oxidation of titanium and TA6V alloy in chromic media. An electrochemical

approach. Electrochimica Acta, 45, 921-929.

Page 196: 1.3 Light emitting diode

177

APPENDIX A: INVESTIGATION OF ULTRTRASONIC EXTRACTION OF

POLYCHORINATED BIPHENYLS FROM SOIL

A.1. Experimental

A.1.1. Chemicals

Aroclor1254 standard (1000 mg/L in hexane) was purchased from AccuStandard,

decachlorobiphenyl (200 mg/L in acetone) was procured from Sigma-aldrich, 99% purity

of 2-propanol (IPA), 99.9% purity of acetone and 99.9% purity of hexane were obtained

from EMD. The water used in this experiment is ultrapure water.

A.1.2. Pre-Processing of contaminated soil

500 g of wet contaminated soil was dried at 50 ℃ for two days. The dried contaminated

soil was then passed through a size-20 sieve and homogenized using a spatula, mortar

and pestle.

A.1.3. Ultrasonic extraction of PCBs

Ten gram of dried soil samples were placed in 40-ml glass vials. Each vial was filled with

a known amount of IPA and be sonicated for different time. After sonication, each vial

was centrifuged for 15 minutes at 1500 rpm. Then, one and half ml of the supernatant

was transferred to a 1.5-ml centrifuge vial for a second centrifugation (5 minutes at 1500

rpm) to remove the trace particles. After second centrifugation, 0.1 ml of secondary

Page 197: 1.3 Light emitting diode

178

centrifugation supernatant was diluted to 1 ml with IPA and analyzed using GC-ECD

(USEPA, 1996b).

A.1.4. Soxhlet extraction of the remaining PCBs in soil :

The remaining solid in part A.1.3 was holed with a Pasteur pipette and washed with 30 ml

water to remove residual IPA extract that might have lingered in the soil. The washing

was conducted by gently shaking the vials for ten seconds. After washing procedure, the

vials were again centrifuged for 5 minutes at 1500 rpm. The separated solid in the vial

was subjected to Soxhlet extraction (USEPA, 1996a). After Soxhlet extraction, the

concentrate extracts were carefully transferred into 250-ml beakers. The beakers

containing extract were then placed inside of a well-ventilated fume hood, at room

temperature, and evaporated to dryness. Ten ml of the hexane was used to redissolve the

dried solid in each beaker and was centrifuged for 5 minutes at 1500 rpm. 0.1-ml of the

supernatant was then diluted to 1ml with hexane and analyzed using GC-ECD. All the

experiments were conducted in duplicates

A.1.5. Calculation of ultrasonic extraction efficiency

The ultrasonic extraction efficiency (η) can be calculated using Equation [A-1]

η = ms−PCBsms−PCBs+ mr−PCBs

× 100% [A-1]

Page 198: 1.3 Light emitting diode

179

Where ms−PCBs is the mass of PCBs extracted using ultrasonication and is calculated

using Equation [A-2]; mr−PCBs is remaining PCBs in the soil after ultrasonication and is

calculated based on Equation [A-3]

ms−PCBs(mg) = C1 × V1 × 10 [A-2]

Where C1 (mg/L) is PCBs concentration obtained from GC equipment in part A.1.4,

V1(mL) is the volume of IPA used for ultrasonic extraction

mr−PCBs (mg) =C2 × 10 mL × 10

R

[A-3]

Where C2 (mg/L) is PCBs concentration obtained from GC equipment in part A.1.5, R is

the recovery rate of surrogate.

A.2. Results and Discussions

The ultrasonic extraction efficiencies of PCBs under different experimental conditions

were shown in Figure A-A-1. The ultrasonic extraction efficiency ranges from 15% to

30% under different conditions. When the IPA/soil ratio is 1:1, the increase of sonication

time from 15 min to 90 min did improve the extraction efficiency. Whereas, at higher

IPA/soil ratio (2:1 or 3:1), the extraction efficiency is not impacted by the investigated

Page 199: 1.3 Light emitting diode

180

Figure A-A-1: Ultrasonic extraction efficiency of PCBs from 10 g soil under different experimental conditions.

sonication time. For the same sonication time, the increase of IPA/soil ratio resulted in a

decrease of extraction efficiency. A possible reason is that at higher IPA/soil ratio

experiment, the volume of IPA is increased, therefore, the volumetric energy captured by

soil and surrounding IPA decrease, which leads to a lower extraction efficiency.

A.3. Reference

USEPA 1996a. USEPA Method 3540C soxhlet extraction. Washingtong, D.C.

USEPA 1996b. USEPA method 8082, polychlorinated biphenyls by gas chromatography.

Washingtong, D.C.

3:1

2:1

1:1

0

5

10

15

20

25

30

15 30

60 90

Extra

ctio

n ef

ficie

ncy

(%)

Sonication time (min)

Page 200: 1.3 Light emitting diode

181

APPENDIX B: INVESTIGATION OF PHOTODEGRADATION OF BIPHENYL

IN ULTRAVIOLET WATER PURIFICATION SYSTEMS

B.1. Experimental

B.1.1. Chemicals

Ninety nine percent purity of biphenyl was obtained from sigma aldrich, 99% purity of

isopropopanol (IPA) was procured from VWR.

B.1.2. Photoreactor

The testing UV water purification system (UVS238S) was purchased from Neotech Aqua

solutions. It is an annular photoreactor (Figure A-B-1) which is equipped with a medium

mercury lamp. The total power output of the medium mercury lamp is 150W and the

power output of that at 254 nm is 45 W. The capacity of the reactor is 6.5 L.

Figure A-B-1: Ultraviolet water purification system.

B.1.3. Photodegradation of biphenyl in IPA

500 mg/L biphenyl solution was prepared by dissolving 20 g biphenyl crystals in 40 L

IPA. The prepared biphenyl solution was then stored in a 250L reservoir. An air pump

Page 201: 1.3 Light emitting diode

182

was used to pump the biphenyl solution into photoreactor from the reservoir at a high

flow rate, and the exit of the photoreactor was connected to reservoir. The

photodegradation experiments were conducted in a circulated mode with three different

flow rates (6 gallon/min, 10 gallon/min, 14 gallon/min). One milliliter sample was

collected at different irradiation time. And the collected samples were analyzed using

GC/FID.

B.2. Results and discussions

The results were plotted as the change of biphenyl concentration verse the exposure

irradiation time in Figure A-B-2. The exposure irradiation equal to the total experimental

time multiplied by the ratio between the reaction zone volume and total treated volume

(6.5L/40L). It showed that biphenyl can be degraded in IPA solvent with a medium

mercury lamp. However, the direct photolysis of biphenyl in such system is relatively

slow, only 15% degradation was observed within 2 hour exposure irradiation time. To

complete remove the biphenyl from the system, a long exposure irradiation time is

required. Within the investigated flow rate range (6, 10, 14 gallon/min), the flow rate

does not impact the biphenyl degradation. To obtain a high degradation efficiency, flow

rate should be high enough to achieve complete mixing of solution in the reactor

chamber. The lowest flow rate studied is able to create enough mixing in the reactor

chamber. The plot of log of normalization of biphenyl concentration verse exposure

irradiation time in Figure A-B-3 indicate that direct photolysis of biphenyl in IPA follow

first order kinetics. The observed first order kinetics constant under different flow rate

varied from 0.075 h-1 to 0.077 h-1.

Page 202: 1.3 Light emitting diode

183

Figure A-B-2: Degradation of biphenyl under different flow rate.

Figure A-B-3: The pseudo first order kinetics of biphenyl degradation under different flow rate.

0.84

0.88

0.92

0.96

1

0 0.5 1 1.5 2 2.5

C/C

o

Exposure irradiation time (h)

Flow rate=6 gallon/min Flow rate=10 gallon/min Flow rate=14 gallon/min

-0.18

-0.16

-0.14

-0.12

-0.1

-0.08

-0.06

-0.04

-0.02

0 0 0.5 1 1.5 2 2.5

ln(C

/Co)

Exposure Irradation time ( h)

Low flow

medium flow

high flow

Page 203: 1.3 Light emitting diode

184

APPENDIX C: UV VIS ABSORPTION SPECTRUM OF DIFFERENT

PESTICIDES

Figure A-C-1: UV-Vis absorption spectra of 40 mg/L of 2,4-D in water.

Figure A-C-2: UV-Vis absorption spectra of 40 mg/L of 2,4-DCP in water.

0

0.1

0.2

0.3

0.4

0.5

0.6

250 300 350 400 450

Abs

orba

nce

Wavelength(nm)

0

0.1

0.2

0.3

0.4

0.5

0.6

250 300 350 400 450

Abs

orba

nce

Wavelength(nm)

Page 204: 1.3 Light emitting diode

185

Figure A-C-3: UV-Vis absorption spectra of 40 mg/L of 4-CP in water.

Figure A-C-4: UV-Vis absorption spectra of 40 mg/L of MCPA in water.

0

0.1

0.2

0.3

0.4

0.5

0.6

250 300 350 400 450

Abs

orba

nce

Wavelength(nm)

0

0.1

0.2

0.3

0.4

0.5

0.6

250 300 350 400 450

Abs

orba

nce

Wavelength(nm)

Page 205: 1.3 Light emitting diode

186

APPENDIX D: THE CALCULATION OF PERCENTAGE OF AVAILABLE

PHOTONIC ENERGY FOR PHOTOCATALYTIC REACTION

Assume the emission spectra of UVA-LED and mercury black lamps follow Gaussian

distribution as Equation [A-D-1].

𝑓(λ) = 𝑒−(λ−𝑏)2/𝑎 [A-D-1]

Where: a, b are constants; λ (nm) is the wavelength of photon, 𝑓(λ) is the relative light

intensity (the ratio between the emission light intensity at λ and the maximum emission

light intensity) .

For UVA-LED lamp (NSHU551B), wavelength corresponding to maximum emission

(λmax) is 365 nm, and the full width at half maximum of emission spectra is 15 nm; for

mercury black lamp (FL8BL-B), wavelength corresponding to maximum emission (λmax)

is 350 nm, and the full width at half maximum of emission spectra is 50 nm. These

conditions can be expressed as follow:

LED lamp �𝜆 = 365𝑛𝑚,𝑓(𝜆) = 1

𝜆 = 365 ± 152𝑛𝑚,𝑓(𝜆) = 0.5

�;

Mercury black lamp �𝜆 = 350 𝑛𝑚,𝑓(𝜆) = 1

𝜆 = 350 ± 502𝑛𝑚,𝑓(𝜆) = 0.5

Based on the conditions above, the spectral function for both UVA-LED and Mercury

black lamps were obtained as Equation [A-D-2] and [A-D-3]:

𝑓𝐿𝐸𝐷(𝜆) = 𝑒−(𝜆−365)2/81.5 [A-D-2]

Page 206: 1.3 Light emitting diode

187

𝑓𝑀𝑒𝑟𝑐𝑢𝑟𝑦(𝜆) = 𝑒−(𝜆−350)2/901 [A-D-3]

The energy of a single photon at wavelength λ can be calculated by Equation [A-D-4]

𝐸(𝜆) = hcλ

[A-D-4]

Where h is Planck`s constant (6.62× 10−34 J.s), c is the light speed in vacuum (3× 108

m/s), and E(λ) (J) is the energy of photon at wavelength λ (m).

Figure A-D-1: The emission spectrum and TiO2 band edge.

In TiO2 photocatalysis, the photon of wavelength above 385 nm is unavailable for the

reaction (Figure A-D-1). Therefore, the percentage of available energy for reaction in

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

250 300 350 400 450

Rel

ativ

e lig

ht in

tens

ity

Wavelength (nm)

UVA-LED Mercury black lamp TiO2 band edge

Page 207: 1.3 Light emitting diode

188

each case is equal to the sum of energy of photon (0 nm < λ<385nm) divided by the sum

of energy of photon (0 nm < λ ).

The percentage of available photonic energy for each case is calculated as follows:

𝑃𝐿𝐸𝐷 =∫ 𝑓𝐿𝐸𝐷(𝜆) ∗ 𝐸(𝜆)𝑑𝜆3850

∫ 𝑓𝐿𝐸𝐷(𝜆) ∗ 𝐸(𝜆)𝑑𝜆+∞0

= 99.9%

[A-D-3]

𝑃𝑀𝑒𝑟𝑐𝑢𝑟𝑦 =∫ 𝑓𝑚𝑒𝑟𝑐𝑢𝑟𝑦(𝜆) ∗ 𝐸(𝜆)𝑑𝜆3850

∫ 𝑓𝑚𝑒𝑟𝑐𝑢𝑟𝑦(𝜆) ∗ 𝐸(𝜆)𝑑𝜆+∞0

= 95.8%

[A-D-3]

Where PLED is the percentage of available photonic energy for LED; PMercury is the

percentage of available photonic energy for LED.