Anda di halaman 1dari 172

ELECTROCHEMICAL WASTEWATER TREATMENT FOR DENITRIFICATION AND TOXIC ORGANIC DEGRADATION USING Ti-BASED SnO2 AND RuO2 ELECTRODES

Xie, Zhaoming Ph. D. THESIS

DEPARTMENT OF CIVIL ENGINEERING THE UNIVERSITY OF HONG KONG


2006

Abstract of thesis entitled

Electrochemical Wastewater Treatment for Denitrification and Toxic Organic Degradation Using Ti-based SnO2 and RuO2 Electrodes
submitted by

Xie, Zhaoming
for the degree of Doctor of Philosophy at the University of Hong Kong in December 2006

Many wastewater pollutants, including toxic organic compounds and nutrients at a high concentration, are difficult to treat with conventional treatment processes. Wastewater treatment technologies require continuous advances to ensure that the effluent continues to comply with more stringent discharge standards and to improve the cost-effectiveness of the practice. The electrochemical (EC) method is an advanced oxidation process that has shown promise as an effective treatment alternative for wastewater containing toxic substances. This study focused on the development of the electrochemical technology for degradation of aromatic and chlorinated aromatic organic pollutants and for nitrogen removal from landfill leachate and sewage sludge liquor. A double-layer coating technique was employed to make titanium-based catalytic active electrodes in Ti/SnO2-Sb-Al for EC organic oxidation. The experimental results show that the Ti/SnO2-Sb-Al anode is much more effective for electrolysis of phenol and chlorophenol than other conventional anodes of Pt, Ti/RuO2 and Ti/IrO2. For phenol (100 mg/L) treatment, complete organic removal was

achieved after 5 hrs on anode Ti/SnO2-Sb-Al at 20 mA/cm2, compared with only 36% and 12% organic reductions provided by the Pt and Ti/RuO2 anodes, respectively. Similarly, for chlorophenol electrolysis, the best treatment performance was obtained with anode Ti/SnO2-Sb-Al, followed by Pt and then Ti/RuO2. Analysis with the High Performance Liquid Chromatography (HPLC) revealed many intermediate products during the EC process, from aromatic compounds to aliphatic acids, which were oxidized rapidly by Ti/SnO2-Sb-Al but accumulated in the Pt and Ti/RuO2 cells. In addition, formation of colored polymeric materials occurred in the Pt and Ti/RuO2 electrolyzed solutions. Based on the EC degradation performance and the HPLC results, modified pathways of phenol and chlorophenol electrolysis at different types of electrodes were proposed. The different electrodes were characterized for their electrochemical behavior and surface properties. Compared to the Pt and Ti/RuO2 anodes, Ti/SnO2-Sb-Al had a considerably higher overpotential of oxygen evolution according to the voltammograms, which would facilitate the production of hydroxyl radicals. Direct phenol and chlorophenol oxidation could take place on the Ti/SnO2-Sb-Al surface but not on the Pt and Ti/RuO2 anodes. The Ti/SnO2-Sb-Al electrode also had apparently a higher affinity with organic substances than Ti/RuO2. It is argued that the anodic property has a significant influence on the kinetics, pathway and mechanism of organic electrolysis. Specific surface treatment such as the SnO2-Sb-Al coating provides the anode with a catalytic function for rapid degradation of toxic organics. In addition, electrochemical denitrification experiments were performed using a lab-scale system with Ti/RuO2-based mesh electrode plates for treating saline sewage sludge liquor and raw landfill leachate. Complete ammonia nitrogen removal was achieved in 1 hour of electrolysis for the sludge liquor (500 mg/L) and in 3 hours for the leachate (1600 mg/L). Electro-chlorination is considered to be the major mechanism of EC ammonia oxidation. Additional electro-flocculation with a pair of iron needle electrodes was able to increase the organic removal by sedimentation from below 30% to about 70%. The experimental results suggest that the EC treatment can be developed as a cost-effective method of denitrification for high-strength wastewater.

Electrochemical Wastewater Treatment for Denitrification and Toxic Organic Degradation Using Ti-based SnO2 and RuO2 Electrodes
By

Xie, Zhaoming

( )

BEng Wuhan Univ H & E Eng, China MEng Wuhan University, China

A thesis submitted in partial fulfillment of the requirements for the Degree of Doctor of Philosophy at The University of Hong Kong

Hong Kong, December 2006

Declaration

I hereby declare that this thesis represents my own work, expect where due acknowledgement is made, and that it has not been previously included in a thesis, dissertation or report submitted to this University or to any other institution for a degree, diploma, or other qualification.

Signed Xie, Zhaoming

Acknowledgements
I am sincerely grateful to my supervisor, Dr. Xiaoyan Li, for his guidance, inspiration, and enthusiasm for academic research that has greatly affected my attitude towards the acquisition of knowledge. His giving of concern for my professional and personal development in these years is appreciated. I also wish to express my gratitude to my labmates, for their technical assistance and enthusiastic discussion during my studies. I am especially grateful to the help of Mr. Keith Wong during daily experiments and Mr. Yunhai Wang in carrying out the electrochemical analysis. I am indebted to all the staffs and postgraduate students in the Department of Civil Engineering and Department of Chemistry, for their friendship and making my PhD study enjoyable. I would like to thank Prof. K.Y. Chan, Prof. Herbert Fang and Dr. JiDong Gu for that they so generously shared with me and for the assistance in using their instruments. Lastly, I would express my deepest appreciation to my dear great mother for her accompany in the road of pursuing the degree, and also to my families, for their understanding, patience, love and support during the course of this work at HKU.

iii

Table of Contents
Declaration ............................................................................................................ Acknowledgements................................................................................................. Table of Contents ................................................................................................... Abbreviations ......................................................................................................... i ii iii xv

List of Illustrations/Figures/Tables........................................................................ xi

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


1.1 Problem statement................................................................................ 1.1.1 1.1.2 1.1.3 Advance oxidation processes and electrochemical method for wastewater treatment........................................................ 1 Electrochemical treatment of toxic organic pollutants .......... Electrochemical denitrification for high-strength 4 5 6 7 7 7 wastewater.............................................................................. 1.2 Objectives of this study........................................................................ 1.3 Scope of this study ............................................................................... 1.3.1 1.3.2 Electrochemical treatment of saline sludge centrate and landfill leachate...................................................................... Preparation of catalytic active anodes for organic oxidation. different electrodes................................................................. 1.3.4 Characterization of the electrochemical and surface properties of different electrodes and their effects on organic electrolysis ................................................................ 8 1.3.3 Degradation of phenol and chlorophenol by electrolysis at 2 1

Chapter 2 Literature review........................................................... 9


2.1 Wastewater treatment for removal of toxic organics and nutrients ..... 2.1.1 2.1.2 Toxic organic pollutants and nutrients in wastewater discharge ................................................................................ Conventional wastewater treatment technologies.................. iii 9 10 9

2.1.3

Wastewater treatment............................................................. 2.1.3.1 Biological treatment processes .............................. 2.1.3.2 Physical-chemical treatment methods....................

10 10 11 12 12 13 13 14 15 16 16 16 17 17 17 18 19 19 20 21 22 23 24 25 25 26

2.1.4

Advanced oxidation processes (AOPs).................................. 2.1.4.1 Ozonation............................................................... 2.1.4.2 Potolysis................................................................. 2.1.4.3 Hydrogen peroxide/Ultraviolet light...................... 2.1.4.4 Limitations of AOPs ..............................................

2.2 Electrochemistry in wastewater treatment ........................................... 15 2.2.1 2.2.2 Development of electrochemical wastewater treatment ........ Application of electrochemical technologies in wastewater treatment ................................................................................ 2.2.2.1 Electro-coagulation, electro-flocculation and electro-flotation...................................................... 2.2.2.2 Electro-chlorination for wastewater treatment....... 2.2.2.3 Electro-chlorination for wastewater disinfection... 2.3 Electro-catalytic oxidation of toxic organic pollutants........................ 2.3.1 2.3.2 2.3.3 Electrochemistry for toxic organic degradation..................... Dimensionally stable anodes (DSA) for organic oxidation ... 2.3.3.1 The substrate base materials: Titanium.................. 2.3.3.2 Antimony doped tin oxide ..................................... 2.3.3.3 Sol-gel coating and thermo-decomposition ........... 2.4 Reaction pathways and mechanisms of electrochemical degradation of toxic organics................................................................................... 2.4.1.1 EC degradation of phenol ...................................... 2.4.1.2 EC degradation of chlorophenol ............................ 2.4.2 Mechanisms of electro-catalytic organic oxidation ............... 2.4.2.1 OH generation....................................................... 2.4.2.2 Electron transfer..................................................... 2.4.2.3 Direct oxidation ..................................................... 2.4.1 Degradation pathway of aromatic compounds ...................... 21

Preparation of catalytic active anodes.................................... 19

Chapter 3 Preparation of Ti-based electrodes.............................. 30

iv

3.1 Metal substrate and the coating materials............................................ 3.1.1 3.1.2 3.1.3 Metal substrate: Titanium ...................................................... Dimensionally stable anodes (DSA) ..................................... Coating materials ................................................................... 3.1.3.1 RuO2 coating film ........................................................ 3.2 Pretreatment of titanium plates ............................................................ 3.3 Preparation of Ti/RuO2 and Ti/IrO2 electrodes.................................... 3.4 Electroplating for the inner layer of Ti/SnO2-Sb-Al electrodes........... 3.4.1 3.4.2 Electroplating solution ........................................................... Cathodic electro-plating.........................................................

30 30 30 31 31 32 34 35 35 36

3.1.3.2 SnO2-based coating film .............................................. 31

3.4.3 Annealing at a high temperature ............................................ 36 3.5 Thermal decomposition of the outer layer of the Ti/SnO2-Sb-Al electrodes ............................................................................................. 3.5.1 3.5.2 3.5.3 3.5.4 3.6.1 3.6.2 Coating solution ..................................................................... Dipping and drying procedures.............................................. Annealing............................................................................... Procedure of the Ti/SnO2-Sb-Al electrode preparation ......... Platinium electrodes............................................................... Ti/RuO2 mesh electrodes used for electrochemical 40 40 40 41 41 41 42 42 denitrification......................................................................... 3.7 Characterization of the electrodes........................................................ 3.7.1 3.7.2 3.7.3 Linear sweep voltammetry (LSV) ......................................... Surface contact angle measurement....................................... Surface morphology............................................................... 3.7.3.1 Microscopic observations ............................................ 3.7.3.2 Scanning electron microscopy (SEM) ......................... 3.7.3.4 Energy Dispersive X-ray (EDX) analysis.................... 37 37 38 38 39 39 39

3.6 Other electrodes used for the experimental research ...........................

3.7.3.3 X-ray Photoelectron Spectroscope (XPS).................... 42

Chapter 4 Experimental materials and methods ......................... 44


4.1 Electrochemical denitrification............................................................ 44

4.1.1

Wastewater samples............................................................... 4.1.1.1 Saline sludge centrate .................................................. 4.1.1.2 Landfill leachate...........................................................

44 44 44 46 48 48 49 49 49 49 50 51 51 51 51 53 55 55 55 55 55 56 56 57 58 58 59 60

4.1.2 4.1.3

Electrochemical denitrification set-up .................................. Experimental conditions ........................................................ 4.1.3.1 Batch denitrification tests for sludge liquor................. 4.1.3.2 Denitrification plus electro-flocculation for organic removal ........................................................................ 4.1.3.3 Continous flow stirred tank reactor (CFSTR) for denitrification of landfill leachate................................

4.1.4

Electro-denitrification and electro-flocculation reactions ..... 4.1.4.1 Electro-chlorination reactions...................................... 4.1.4.2 Electro-flocculation reactions ......................................

4.1.5

Analytical methods ................................................................ 4.1.5.1 Chlorine compounds and chloride ............................... 4.1.5.2 Nitrogen (NH3-N, NO2--N, NO3--N) concentrations.... 4.1.5.3 Trihalomathane (THM) formation potential measurement ................................................................ 4.1.5.4 Other water quality parameters.................................... 4.1.5.5 Current efficiency of EC denitrification ......................

4.2 Electrochemical degradation of aromatic organic pollutants............... 55 4.2.1 Wastewater samples............................................................... 4.2.1.1 Phenol solution............................................................. 4.2.1.2 Chlorophenol solution.................................................. 4.2.1.3 Solutions of the intermediate organics of EC phenol degradation................................................................... 4.2.2 Experimental set-up and operational conditions.................... 4.2.2.1 Experimental set-up ..................................................... 4.2.2.2 Operation and sampling ............................................... 4.3 Analytical methods .............................................................................. 4.3.1 4.3.2 4.3.3 High performance liquid chromatography (HPLC)............... Current efficiency and TOC removal efficiency ................... Rate constants for EC oxidation of organic chemicals ..........

vi

Chapter 5 Nitrogen

removal

from

wastewater

by

electrochemical denitrification .................................... 61


5.1 General characteristics of the waste liquor samples and EC treatment .............................................................................................. 5.2 Electrochemical denitrification of sludge centrate liquor.................... 5.2.1 5.2.2 5.2.3 Electrochemical denitrification.............................................. Electro-chlorination for denitrification of the sludge centrate................................................................................... Power consumption and current efficiency............................ 64 65 68 68 70 71 73 61 62 62

5.2.4 Electro-flocculation for organic removal............................... 66 5.3 Electrochemical denitrification of landfill leachate............................. 5.3.1 5.3.2 5.3.3 Batch EC leachate treatment tests.......................................... EC denitrification during the continuous flow process.......... Power consumption and current efficiency............................

5.4 Summary ..............................................................................................

Chapter 6 Electrochemical degradation kinetics of phenol and its intermediate products.............................................. 74


6.1 Performances of selected anodes in EC phenol degradation ............... 6.1.2 6.1.3 6.2.1 6.2.2 6.2.3 6.2.4 Solution pH during the EC process........................................ Current efficiency and TOC removal efficiency ................... Direct products of phenol oxidation ...................................... Aliphatic acids from phenol oxidation................................... Polymeric compounds produced during the EC process ....... Grouping of intermediate organic compounds ...................... electrolysis ............................................................................. 6.3 EC phenol degradation pathways on different types of electrodes...... 6.3.1 6.3.2 6.3.3 Major phenol electrolysis reactions ....................................... Aromatic ring cleavage vs. polymerization ........................... Pathways of EC phenol degradation ...................................... 74 76 78 78 78 83 83 84 84 86 86 86 87 6.1.1 Phenol degradation and TOC removal................................... 74

6.2 Intermediate products and reaction routs for different electrodes .......

6.2.5 EC destruction of the intermediate products of phenol

vii

6.4 Summary ..............................................................................................

88

Chapter 7 Surface and electrochemical properties of the electrodes and EC phenol degradation mechanisms . 93
7.1 Characterization of the electrodes by linear sweep voltammetry (LSV) ................................................................................................... 7.1.1 LSV and oxidation potentials................................................. 7.1.1.1 Two electrochemical reactions: phenol oxidation and oxygen evolution.......................................................... 7.1.1.2 LSV of the electrodes in the electrolyte without phenol........................................................................... 7.1.2 LSV curves in the solutions with phenol addition ................. 7.1.2.1 Anode Ti/SnO2-Sb-Al.................................................. 7.1.2.2 Anode Pt....................................................................... 7.1.2.3 Anode Ti/RuO2 ............................................................ 7.1.3 Effect of the solution pH on the phenol oxidation overpotential.................................................................................. 7.2 Electrochemical phenol oxidation mechanisms................................... 7.2.1 7.2.2 7.2.3 EC phenol oxidation process and mass transfer ................... EC phenol degradation steps.................................................. Oxygen evolution................................................................... 98 99 100 101 102 102 95 95 95 95 97 94 93 93

7.1.2.4 Effect of phenol concentration on the LSV curve ....... 97

7.3 Summary ..............................................................................................

Chapter 8 Electrochemical

degradation

of

chlorophenol
104 104 106 106 108 109

kinetics and pathways ................................................... 104


8.1 EC chlorophenol oxidation .................................................................. 8.1.1 8.1.2 8.1.3 8.2.1 8.2.2 Chlorophenol oxidation and TOC removal by electrolysis ... The solution pH during electrolysis....................................... Current efficiency .................................................................. HPLC analysis of intermediate compounds........................... Grouping of the intermediate organic compounds.................

8.2 Intermediates and pathways of EC chlorophenol degradation ............ 108

viii

8.2.3 8.2.4 8.3.1 8.3.2 8.4.1 8.4.2

Transformation and degradation of chlorophenol during electrolysis ............................................................................. Basic pathways of EC chlorophenol degradation .................. LSV of the electrodes in chlorophenol solution .................... Peak current density vs. LSV scan rate.................................. Electrolysis mechanisms........................................................ EC chlorophenol degradation pathways ................................ 110 115 116 116 119 119 119 120 121

8.3 Voltammetric study of anodes in monichlorophenol solutions ...........

8.4 EC chlorophenol degradation mechanisms..........................................

8.5 Summary ..............................................................................................

Chapter 9 Discussion: comparison of the surface morphology and properties between different electrodes............... 123
9.1 Surface properties ................................................................................ 9.1.1 9.1.2 Coating density ...................................................................... 9.1.2.1 The mechanism of deactivation ................................... 9.1.2.2 Imagines of the titanium plates ................................... 9.1.3 Coating surface components .................................................. 9.1.3.2 EDX analysis ............................................................... 9.1.4 9.1.5 9.2.1 9.2.2 9.2.3 9.2.4 9.2.5 9.3.1 9.3.2 Contact angle and hydropilicity ............................................. Organic adsorption onto the electrodes.................................. EC degradation of organic matters ........................................ Organic degradation pathways............................................... Surface electrochemical properties ........................................ EC oxidation mechanisms...................................................... Stability of the electrodes....................................................... Improvement of electrode preparation techniques................. 123 123 123 124 126 126 128 129 131 131 132 132 133 134 134 134

Deactivation of electrode Ti/SnO2-Sb-Al .............................. 123

9.1.2.3 The process of deactivation ......................................... 124 9.1.3.1 XPS analysis ................................................................ 126

9.2 Comparison of different electrodes for organic electrolysis................

9.3 Future work..........................................................................................

Investigation of EC organic oxidation mechanisms .............. 133

ix

Chapter 10 Conclusions .................................................................... 136


10.1 Electrochemical treatment of saline sludge centrate and landfill leachate for denitrification and organic removal ................................. 10.2 Degradation of phenol and chlorophenol by electrolysis with the Ti/SnO2-Sb-Al and other electrodes .................................................... 10.3 Surface properties of electrode Ti/SnO2-Sb-Al and their effects on EC organic degradation........................................................................ 138 137 136

References........................................................................................... 140

List of Illustrations

Figures
Figure 2.1. Figure 2.2. Figure 3.1. Figure 3.2. Figure 3.3. Figure 3.4. Figure 4.1. Figure 4.2. Figure 4.3. Summary of the reaction mechanisms of electrochemical phenol degradation....................................................................................... Reaction mechanism of anodic organic oxidation ........................... Surface morphology of Titanium plate ........................................... Electro plating set-up ....................................................................... Well treated electrode Ti/SnO2-Sb-Al ............................................. Flow chart of the Ti/SnO2-Sb-Al electrode preparation procedures Landfill facility in New West Territories of Hong Kong ................ (A) Electrochemical denitrification system illustration; (B) Electrochemical Figure 4.4. Figure 4.5. Figure 4.6. Figure 4.7. Figure 4.8. Figure 4.9. denitrification set-up; (C) Ti/RuO2-ZrO2 47 50 52 54 54 electrode mesh plates ....................................................................... Landfill leachate CFSTR set-up....................................................... UV/Visible spectrometer (Lambda 12, Perkin Elmer) .................... TOC analyzer (TOC-5000A, Shimadzu) ......................................... Turbidimeter (2100N, HACH)......................................................... 22 28 33 36 39 43 45

Dewatering systems in Shatin Wasterwater Treatment Works........ 45

Portable datalogging spectrophotometer (DR/2010, HACH).......... 52

Electrochemical organic degradation experiment set-up................. 56 (HP 6890 )........................................................................................ 57 58 59 62

Figure 4.10. Gas chromatograph mass spectrum (GC-MS) for HTM analysis Figure 4.11. High performance liquid chromatography (HPLC) (Waters 2695 2996) ................................................................................................ Figure 4.12. Characterization of different anodes................................................ Figure 5.1. Figure 5.2. TIN removal as a function of contact time and current density with a gap distance of 8 mm between the electrodes....................... (a) Nitrogen removal as a function of treatment time; (b) Chlorine compounds during the EC process at a current density of 125 mA/cm2 with an electrode gap of 3 mm .......................................... 63

xi

Figure 5.3. Figure 5.4. Figure 5.5.

Current efficiency as a function of current input for different electrode gaps in EC denitrification................................................. TIN removal as a function of power consumption for different electrode gaps at the same current intensity of 75 mA/cm2 ............. Comparison in removals of TOC, COD and SS, sludge production and TIN degradation for conditions of (A) ECdenitrification + EC-flocculation; (B) EC-denitrification only; (C) EC-flocculation only; (D) EC-denitrification + chemical 68 69 70 71 72 72 73 75 77 77 79 79 80 flocculation ...................................................................................... 66 65

Figure 5.6. Figure 5.7. Figure 5.8. Figure 5.9.

TIN removal as a function of the electrolysis time for different chloride concentrations in EC denitrification of landfill leachate ... TIN removal and chloride concentration as a function of contact time in the CFSTR for EC denitrification........................................ Current efficiency and TIN degradation as the function of reaction time in batch operation....................................................... Current efficiency and TIN degradation as the function of reaction time in continous................................................................

Figure 5.10. TIN removal as a function of power consumption .......................... Figure 5.11. Untreated leachate (left) and EC treated leachate(right).................. Figure 6.1. Figure 6.2. Figure 6.3. Figure 6.4. Figure 6.5. Figure 6.6. Figure 6.7. Figure 6.8. TOC and solution pH during EC phenol degradation...................... Phenol removal by the three types of anodes................................... TOC and solution pH vs. time during phenol degradation .............. Current efficiency on phenol removal ............................................. Current efficiency on TOC removal ................................................ HPLC analysis of intermediates of EC phenol degradation on the three anodes ..................................................................................... TOC components of intermediates of EC phenol degradation by the three anodes: Ti/SnO2-Sb-Al (left), Pt (middle), and Ti/RuO2 (right) ............................................................................................... Figure 6.9. Degradation of pure intermediate chemicals on the three anodes ... Figure 6.10. HPLC analysis of aromatic compounds........................................... Figure 6.11. HPLC analysis of aliphatic acids .....................................................

Phenol destruction on the three types of anodes.............................. 75

81 82 89 89

Figure 6.12. Phenol degradation pathway for electrode Ti/SnO2-Sb-Al.............. 90 xii

Figure 6.13. Phenol degradation pathway for electrode Pt .................................. Figure 6.14. Phenol degradation pathway for electrode Ti/RuO2 ........................ Figure 7.1. Figure 7.2. Figure 7.3. Figure 7.4. Figure 7.5. Figure 7.6. Figure 7.7. Figure 7.8. Figure 8.1. Figure 8.2. Figure 8.3. Figure 8.4. Figure 8.5. Figure 8.6. Figure 8.7. Phenol oxidation for electrode Ti/SnO2-Sb-Al to different concentrations .................................................................................. Phenol oxidation for electrode Ti/RuO2 to different

91 92

Phenol oxidation on the three types of anodes................................. 93 94

Phenol oxidation for electrode Pt to different concentrations ......... 96 concentrations .................................................................................. 96

Phenol oxidation for electrode Ti/SnO2-Sb-Al to different pH ....... 98 Phenol oxidation for electrode Pt to different pH............................ 99 Phenol oxidation for electrode Ti/RuO2 to different pH.................. EC phenol oxidation reactions ......................................................... TOC removal on the three types of anodes...................................... TOC degradation as a function of charge input at different current density .............................................................................................. MCP degradation as a function of the charge input to different anodes ............................................................................................. HPLC analysis of intermediates of EC MCP degradation on the three anodes (aromatic compounds) ................................................ HPLC analysis of intermediates of EC MCP degradation on the three anodes (aliphatic acids)........................................................... TOC components of intermediates of EC MCP degradation by the three anodes: Ti/SnO2-Sb-Al (left), Pt (middle), and Ti/RuO2 (right) ............................................................................................... 113 114 115 116 118 124 112 111 109 107 100 101 105

MCP destruction by the three types of anodes................................. 104

Figure 8.8. Figure 8.9.

MCP electrolysis pathways for the three anodes ............................. LSV of anode Ti/SnO2-Sb-Al in MCP solution...............................

Figure 8.10. LSV of anode Pt in MCP solution ................................................... Figure 8.12. Current density as a function of the scan rate square root ............... Figure 9.1. Figure 9.2. Figure 9.3. Titanium plate under microscope..................................................... A typical full-width XPS spectrum of the Ti/SnO2-Sb-Al electrode........................................................................................... xiii

Figure 8.11. LSV of anode Ti/RuO2 in MCP solution ......................................... 117

SEM images of the electrode in different service stage................... 125 126

Figure 9.4. Figure 9.5. Figure 9.6.

EDX analysis of electrode Ti/RuO2 ................................................. Contact angle measurement .............................................................

127

EDX analysis of electrode Ti/SnO2-Sb-Al ...................................... 129 130

Tables
Table 3.1. Table 3.2. Table 3.3. Coating solution for the Ti/RuO2 electrodes.................................... Coating solution for the Ti/IrO2 electrodes...................................... Coating solution (100 mL) for inner layer electro-plating of the Ti/SnO2-Sb-Al electrodes ................................................................ Table 3.4. Table 3.5. Electroplating conditions tested and selected .................................. Coating solution (100 mL) for the outer layer of the Ti/SnO2-SbAl electrodes .................................................................................... Table 3.6. Table 4.1. Table 4.2. Table 5.1. Table 5.2. Table 6.1. Table 9.1. Thermal deposition conditions tested and selected.......................... Basic water quality parameters of the sludge centrate samples....... Basic water quality parameters of the landfill leachate samples ..... 38 39 46 46 35 37 34 34

Effect of electro-flocculation on the sludge liquor treatment .......... 67 Effect of electro-flocculation on sludge liquor treatment ................ 55

Kinetic constants comparison for different species incineration ..... 85 Contact angle (advancing and receding contact 128 131

angle)measurement ......................................................................... Table 9.2. The quantities of organic adsorption on different electrodes...........

xiv

Abbreviations

AOP
o

advanced oxidation processes degree Celsius chemical oxygen demand cyclic voltammetry Dimensionally stable anodes Electrochemical Energy Dispersive X-ray granular activated carbon gas chromatography mass spectrum High performance liquid chromatography linear sweep voltammetry mono-chlorophenol ammonia nitrogen nitrite nitrogen nitrate nitrogen platinum electrode plate suspended solids scanning electron microscopy trihalomethane titanium electrode plate titanium electrode coated with iridium oxide film titanium electrode coated with ruthenium oxide film titanium electrode coated with tin oxide film dopped with antimony and aluminium

COD CV DSA EC EDX GAC GC-MS HPLC LSV MCP NH3-N NO2--N NO3--N Pt SS SEM THM Ti Ti/IrO2 Ti/RuO2 Ti/SnO2-Sb-Al TIN TOC UV XPS

total inorganic nitrogen total organic carbon ultraviolet X-ray Photoelectron Spectroscopy

xv

Chapter 1: Introduction

Chapter 1

Introduction
1.1 Problem statement 1.1.1 Advanced oxidation processes and electrochemical methods for wastewater treatment All communities and industrial activities produce liquid waste wastewater that contains various contaminants. Discharge of untreated wastewater into natural waters can cause serious water pollution problems. Decomposition of the wastewaterderived organic matter will consume dissolved oxygen in the receiving water and lead to nuisance conditions. Toxic compounds in the wastewater will increase the toxicity of the natural water and upset the ecological system. Nutrients such as nitrogen and phosphorus can stimulate rapid growth of algae, resulting in algal blooms and red tides. Therefore, wastewater treatment for effective removal of a wide variety of contaminants is essential to the protection of public health and the environment. Conventional wastewater treatment is a combination of physical, chemical and biological processes. The treatment system is usually designed for the removal of domestic organic matter, suspended solids and pathogenic microorganisms. Many conventional treatment processes are neither capable nor cost-effective for treating high-strength wastewater and various types of industrial effluents. For example, a high ammonium content in the wastewater influent has always been a challenge to the biological nitrogen removal system. Recalcitrant and toxic organic chemicals from industrial discharges are more difficult to deal with using conventional treatment methods. Therefore, wastewater treatment technologies require continuous advances to keep up with the development in industrial and commercial activity. In addition, new treatment methods have to be developed to allow compliance with more stringent

Chapter 1: Introduction wastewater discharge standards and to meet the requirements for many special applications. Advanced oxidation processes (AOP), such as ozonation, UV/H2O2 oxidation, Fentons reaction, photo-Fenton process, sonolysis and supercritical wet oxidation, have shown promise as alternative techniques for the detoxification of wastewaters containing toxic substances (Shen et al., 1995; Kajitvichyanukul et al., 2006). AOPs have been proven to be able to destroy numerous toxic organic and halogenated organic chemicals (Azbar et al., 2004; Yonar et al., 2006). AOPs have the potential to completely mineralize organic contaminants to inorganic salts, CO2 and H2O without generation of secondary pollution commonly associated with conventional treatment processes. The principal mechanism of AOPs involves the generation of highly reactive free radicals (Agustina et al., 2005). Many AOP technologies generate hydroxyl radicals that are a powerful, non-selective chemical oxidant capable of destructing a wide range of recalcitrant organic compounds. Electrochemical oxidation is another new AOP alternative. Recently, there has been an increasing research interest in the use of electrochemical (EC) methods for water purification and wastewater treatment (Vlyssides et al., 2002; Chen, 2004). Research has shown that toxic organics in various types of wastewater effluents can be effectively oxidized by electrochemical reactions (Vlyssides et al., 2002). The EC process is simple in configuration and robust in operation. With the unique features, the EC process has the potential to be developed as a cost-effective technology for treating toxic and refractory organic pollutants, particularly for low-volume applications (Li et al., 2005). The present study focuses on the development of the EC treatment technology for the degradation of aromatic and chlorinated aromatic organic pollutants in wastewater. In addition, the application of the EC process for nitrogen removal from landfill leachate and sewage sludge liquor is also investigated. 1.1.2 Electrochemical treatment of toxic organic pollutants Aromatic toxic organic compounds are common pollutants in the waste effluent from many industrial sectors, such as petroleum refineries, synthetic chemical plants, plastics, pulp and paper, textiles, detergent, pesticide and herbicider, and pharmaceutical plants. Among the harmful organic chemicals in industrial wastewaters, phenol and phenolic substances have attracted great attention in the last

Chapter 1: Introduction two decades owing to their toxicity and quantity in industrial discharges. Phenolic substances are persistent pollutants with high toxicity that are released in the wastewaters of a considerable number of industries (Morao et al., 2004). Moreover, phenol is often an intermediate product in the oxidation pathway of higher molecular weight aromatic hydrocarbons. Thus, phenol and chlorinated phenols have been commonly taken as model compounds for studying the effectiveness of various advanced wastewater treatment technologies. Treatment of industrial effluents containing toxic organics such as phenol and chlorinated phenols is a difficult task. In the case of phenol, the discharge limit is 20 g/L, yet it is difficult to extract phenol from wastewater when its concentration is lower than 4000 mg/L. This makes recycling/recovery impractical and hence the phenolic waste must be treated. Biological treatment, such as the activated sludge process, is normally the cheapest and simplest technology to remove waste organics. However, chlorinated phenols are often recalcitrant and must be treated differently. Two separate reports showed that wastewaters containing 200 mg/L or more of 2MCP could not be effectively treated at biological treatment facilities (Bonfatti et al., 1999; Kim et al., 1990). If a recalcitrant compound such as phenol and chlorophenol cannot be treated by a conventional biological approach, an alternate treatment technology will be directed towards modifying its chemical structure to make it more biodegradable, less toxic, or towards complete oxidization to CO2. Recent research has demonstrated that electrochemical technologies can be applied to the treatment of wastewater containing organic pollutants such as phenol and phenolics (Polcaro and Palmas, 1997; Bonfatti et al., 1999; Kim et al., 1990; Gu et al., 2006). However, the effectiveness of EC wastewater treatment depends on the types and properties of the anodes used in the application (Trasatti, 2000; Chen et al., 2002; Chen, 2004). Traditional electrodes, such as graphite and nickel, are not applicable for organic degradation (Rodgers et al., 1999). Titanium based dimensionally stable anodes (DSAs) are found to have a different degree of success. The commonly used RuO2 and IrO2 coating surface does not show a high reactivity for organic oxidation (Arikawa et al., 1998; Rossi and Boodts, 2002). Other DSA coating materials, such as PbO2 and SnO2 based coating, have been introduced to improve the treatment performance. These types of electrodes appear to be more reactive toward organic destruction (Trasatti, 2000). However, the stability, durability and reproducibility have been problems for these electrodes. The coating techniques 3

Chapter 1: Introduction still need to be modified to improve the quality of the coating surface for better wastewater treatment performance and more cost-effective application. The difference in the effectiveness of different anode materials also demonstrates the complexity of the EC oxidation mechanisms involved. Phenol oxidation could stop with the products such as maleic acid for the Pt anode (Comninellis and Pulgarin, 1991; Gattrell and Kirk, 1993). However, complete phenol degradation could be delivered by the PbO2 electrodes (Iniesta et al., 2001; Feng and Li, 2003). Therefore, the oxidation pathway of aromatic chemicals on different anodes remains a subject for investigation. The effect of the anode surface property on the EC reaction mechanism and reaction kinetics also needs to be characterized in greater details. 1.1.3 Electrochemical denitrification for high-strength wastewater Many types of wastewaters, such as the sewage sludge liquor and landfill leachate, contain a high ammonia concentration. Secondary biological wastewater treatment produces a large volume of sludge. The sludge is usually treated by anaerobic digestion followed by dewatering prior to its final disposal. The sludge liquor from the dewatering unit is then returned to the inlet of the treatment system. Although the flow rate of the sludge liquor is low, accounting for about 1% of the influent, it has a high nitrogen content with a NH4+-N concentration between 500 to 800 mg/L. The ammonia loading from sludge liquor contributes to 10-20% of the influent nitrogen loading. Removing this ammonia loading will either improve the performance or increase the capacity of the mainstream wastewater treatment process for nutrient removal (Jetten et al., 1997). In addition, separate treatment of the sludge liquor can take advantage of the high NH4+ concentration in the liquor, which eventually increases the overall cost-effectiveness of a municipal wastewater treatment plant (Arnold et al., 2000; Kolisch and Rolfs, 2000). Leachate from landfills is a special municipal waste liquid that probably is most difficult and expensive to treat. Landfill leachate usually has a NH4+-N concentration of more than 3000 mg/L together with about the same amount of refractory organic pollutants. Inflow of raw leachate into a municipal wastewater treatment plant will considerably increase the loading of the treatment facility. Pretreatment of landfill leachate is normally required before its flow into a municipal

Chapter 1: Introduction wastewater works (Lau, 2000). Current leachate treatment methods, including air stripping and biological processes, are either too expensive or less effective in actual applications. For example, the biodegradability of the leachate has an inverse trend with the landfill age, and only young landfill leachate is fit to the biological treatment process (Chiang et al., 1995). In using a biological nitrogen removal process, a large amount of readily biodegradable organic needs to be dosed into the system, increasing the treatment cost. In the following decades, landfill will continuously be employed as an option for solid waste disposal. Thus, there is a growing need for development of more cost-effective treatment technologies for the leachate from aged landfill facilities. The electrochemical (EC) method can be considered for nitrogen removal in treatment of high-strength wastewater such as sludge liquor and landfill leachate. An EC system is simple and robust in structure and operation. During EC treatment the waste liquid will be forced through a reactor of electrolysis, in which electrochlorination and the subsequent ammonia oxidization are expected. Due to the shortage of freshwater supply in recent years, seawater has been increasingly used for domestic purposes, such as toilet flushing, in Hong Kong and a few coastal cities in China. The inclusion of seawater in sewage increases its salinity content, which can largely improve the performance of the EC denitrification process (Li et al., 2002; Diao et al., 2003). As for the landfill leachate, its chloride concentration is usually high at a level of around 3000 mg/L. The EC method can be developed as an effective leachate treatment alternative for ammonia oxidization and nitrogen removal. Moreover, the effluent quality of EC denitrification is expected to be more consistent and reliable due to its simple working conditions compared to other nitrogen removal processes. 1.2 Objectives of this study In the present study, a coating technique was developed for making Ti-based electrodes. The Ti electrodes were coated with a layer of SnO2 doped with Sb and Al (Ti/SnO2-Sb-Al). Laboratory experiments were carried out to examine the reactivity of the Ti/SnO2-Sb-Al anodes toward destruction and oxidation of toxic organics. Phenol and chlorophenol were selected as the model aromatic compounds for EC treatment. Comparative experiments of organic degradation were conducted with the

Chapter 1: Introduction Pt electrode and standard DSA electrodes, Ti/RuO2 and Ti/IrO2. The effectiveness and performance of different anode types for toxic organic reduction were investigated. The intermediate products of the EC process with different anodes were measured by HPLC for determination of the phenol and chlorophenol degradation kinetics and pathways. The surface properties of different anodes were characterized and the related EC reaction mechanisms were proposed. In addition, laboratory research was conducted for ammonia removal by electrochemical oxidation on the Ti/RuO2 electrode. The capacity and efficiency of EC denitrification for treatment of landfill leachate and sewage sludge liquor were evaluated. The system configuration and operation parameters for better treatment process performance were determined. In summary, the objectives of the present study are:

To make Ti/SnO2-Sb-Al electrodes using a modified coating method to improve the reactivity of the electrodes toward organic oxidation and the stability and durability of the electrodes in EC wastewater treatment.

To examine the effectiveness of the Ti/SnO2-Sb-Al electrodes for the degradation of phenol and chlorophenol, in comparison with the treatment performance of the Pt anode and the conventional DSA Ti/RuO2 and Ti/IrO2 electrodes.

To determine the reaction kinetics, intermediate products and degradation pathways of electrolysis of phenol and chlorophenol on the three different anodes of Ti/SnO2-Sb-Al, Pt and Ti/RuO2.

To characterize the surface and electrochemical properties of the different electrodes in relation to their performance in organic electrolysis, and to investigate the underlining mechanisms of EC organic degradation by different types of electrods.

To develop the EC process for both ammonia removal and organic reduction from sewage sludge liquor and landfill leachate, to evaluate the treatment efficiency of EC denitrification and to optimize the process configuration and operation.

1.3 Scope of this study This research began with the preparation of effective catalytic active anodes in Ti/SnO2-Sb-Al for organic oxidation. The study focused on the reactivity of the 6

Chapter 1: Introduction anodes toward electrolysis of phenol and chlorophenol in comparison with other electrodes. In Chapter 2, previous studies on wastewater treatment using electrochemical methods were reviewed, including organic oxidation and related mechanisms. Chapter 3 described the methodologies used in preparation of electrodes Ti/SnO2-Sb-Al and Ti/RuO2 and characterization of the electrodes. The following Chapter 4 added the materials and methods of the wastewater treatment experiments, wastewater samples, electrolysis experiment procedures, operational conditions and analytical methods. In Chapter 5, experiment results of electrochemical denitrification and organic matters removal were reported. Chapters 6 and 7 presented the results of EC phenol treatment by different electrodes, including the reaction kinetics, intermediate products and possible degradation pathways. Chapter 8 summarized the findings of a similar electrolysis study for chlorophenol degradation. In Chapter 9, the effects of the surface properties of different anode types on the degradation of toxic organic matter and the underling EC treatment mechanisms were discussed. 1.3.1 Electrochemical treatment of saline sludge centrate and landfill leachate Laboratory experiments were carried out to investigate the effectiveness of EC treatment of saline sewage sludge liquor and landfill leachate for denitrification and organic removals. The influences of system and operational parameters, such as the current intensity, gap distance between the electrodes, chloride content, power input and contact time, on the performance of EC denitrification were evaluated. The formation of chlorination by-products during the EC treatment was also determined. 1.3.2 Preparation of catalytic active anodes for organic oxidation Titanium-based catalytic active metal oxide anodes in Ti/SnO2-Sb-Al were prepared using a modified coating approach. The electrode was coated on Ti substrate with double layers of Sn-Sb-Al oxides. The internal layer was deposited by electroplating and calcified at high temperature in an oven. The outer layer was prepared by sol-gel methods for metal element deposition followed by oxidation in a high temperature oven. Electrode Ti/RuO2 was also prepared by coating in accordance with the conventional thermal deposition method . 7

Chapter 1: Introduction 1.3.3 Degradation of phenol and chlorophenol by electrolysis at different electrodes Phenol and chlorophenol were selected as model toxic organic chemicals for testing the treatment capacity of the Ti/SnO2-Sb-Al electrode. The comparative degradation experiments were also conducted with other electrodes, including the Pt, Ti/RuO2 and Ti/IrO2 electrodes. Rates of aromatic destruction and organic reduction, current efficiency and degradation kinetics of the organic electrolysis on different anodes were determined. The intermediate products of the EC treatment process were measured by HPLC in great details, and the degradation pathways of phenol and chlorophenol by different types of anodes were formulated. EC oxidation tests were also conducted for individual pure chemicals that were intermediates of EC phenol degradation in order to identify the possible rate limiting steps in EC organic degradation. 1.3.4 Characterization of the electrochemical and surface properties of different electrodes and their effects on organic electrolysis The electrochemical properties of different types of anodes were characterized by voltammetry in a pure electrolyte solution and organic solutions of different concentrations. The responses of the current intensity to an increase in potential input in different solutions offered important information about the electrochemical reactions, including oxygen evolution, phenol and chlorophenol oxidation. In addition, the coating surface properties of the electrodes were analyzed for their morphology, hydrophobicity and element components. The findings suggest the influences of the anode surface on the performance of EC organic degradation and the underlining mechanisms.

Chapter 2: Literature review

Chapter 2

Literature review
2.1 Wastewater treatment for removal of toxic organics and nutrients 2.1.1 Toxic organic pollutants and nutrients in wastewater discharge Water is the most ubiquitous biological compound and is imperative to life. Runoff from urban settings has significant impact on the quality of the receiving waters (Amvrossios et al., 2006). In cities that have combined sewer systems, the runoff is treated with municipal wastewater in sewage treatment plants (Reemtsma et al., 2000). The combined sewer contains many contaminants, including potentially high concentrations of suspended solids, biochemical oxygen demand, oils and grease, toxics, nutrients, floatables, pathogenic microorganisms, and other pollutants (Amvrossios et al., 2006). Additional to municipal wastewater, wastewater treatment plants (WWTPs), especially those located in industrial areas, face discharges containing a complex mixture of various organic and inorganic substances (Farre et al., 2003; Karvelas et al., 2003; Katsoyiannis et al., 2004; Blanchard et al., 2004; Pham and Proulx, 1997). Persistent Organic Pollutants (POPs) are common pollutants in the waste effluent from many industrial sectors, such as petroleum refineries, synthetic chemical plants, plastics, pulp and paper, textiles, detergent, pesticide and herbicide, and pharmaceutical factories. POPs constitute a wide group of compounds, such as aromatic compounds (Breivik et al., 2004), chlorinated aromatic compounds, polychlorinated biphenyls (PCBs) and organochlorine pesticides (OCs). These compounds are characterized by pronounced persistence against chemical/biological degradation, high environmental mobility, and strong tendency for bioaccumulation in human and animal tissues, significant impacts on human health and the environment

Chapter 2: Literature review even at extremely low concentrations. Their low biodegradability makes them refractory to the biological wastewater treatment (Katsoyiannis et al., 2004, Katsoyiannis and Samara, 2006). Alternative technologies should be applied for toxic organic removal, which would modify the chemical structure of the organic to make it more biodegradable, less toxic, or more completely oxidize. Ammonium treatment is essential in wastewater treatment plants for preventing eutrophication of the receiving water. The traditional biological nitrogen removal can be achieved by a sequence of nitrification and denitrification processes (Kim et al., 2004). Among the nitrogenous compounds, sometimes the ammonium concentration can reach 10003000 mg/L, which is 50100 times higher than in municipal wastewater (Ra et al., 2000; Beline et al., 2001; Su et al., 2001; Cao et al., 2003; Carrera et al., 2003; Uygur and Kargi, 2004; Joo, 2006). Removing this ammonia loading will either improve the performance or increase the capacity of the mainstream wastewater treatment process for nutrient removal (Jetten et al., 1997). 2.1.2 Conventional wastewater treatment technologies 2.1.2.1 Wastewater treatment Wastewater treatment is a process in which the pollutants in wastewater are partially removed and partially changed by decomposition from highly complex, putrescible, organic solids to minerals or relatively stable chemicals. Conventional wastewater treatment consists of a combination of physical, chemical, and biological processes and operations to remove solids, organic matter and, sometimes, nutrients from wastewater. General terms used to describe different degrees of treatment, in order of increasing treatment level, are preliminary, primary, secondary, and tertiary and/or advanced wastewater treatment (Sonune and Ghate, 2004). 2.1.2.2 Biological treatment processes Microorganisms have been used for decades for the treatment of municipal wastewaters, and more recently for the treatment of industrial waste liquids. Its advantages include low cost and easy operation, although these are not always achievable in practice, and public acceptance. Biological wastewater treatment most 10

Chapter 2: Literature review often employs aerobic conditions, in order to mineralize organic compounds to CO2 and H2O. In general, aerobic bioreactors achieve satisfactory levels of organic removals, with problems of sludge formation and by-product off-odors (Sonune and Ghate, 2004). Biological denitrification is known to occur by the action of heterotrophic bacteria using available carbon sources (John and Robert, 1985; Lee et al., 1995 and Lee et al., 1997). Because the influent with a low C/N ratio is deficient in organic carbon and the low carbon source level can limit the overall biological denitrification results, sufficient organic source must be provided for proper denitrification (Itokawa et al., 2001; Kim et al., 2004). Thus, conventional nitrification can be implemented only after pretreating the wastewater to adjust the C/N ratio (Khin and Annachhatre, 2004) or diluting the wastewater (Rostron et al., 2001; Jung et al., 2004). Although the activated sludge process has been successfully used for treating various domestic and industrial wastewaters, the high concentration of toxic pollutants such as phenol, chlorophenol and other aromatic compounds severely inhibits biological activities of activated sludge (Amor et al., 2005; Liu et al., 2005; Young et al., 2006). Only granular activated carbon (GAC) and specially designed bio-treatment processes have been used commercially for treatment of toxic organic pollutants. There still remains the need for reliable, economic and safe treatment alternatives. Electrochemical oxidation is one of conceivable candidates for such a treatment technology. 2.1.2.3 Physical-chemical treatment methods Other methods for wastewater treatment that are being used or are in the development stage are physical and chemical processes, including flocculation, solvent extraction, precipitation, coagulation and sedimentation, sand filtration, activated carbon adsorption, membrane filtration, chemical or electrochemical oxidation (El-Geundi, 1991, Janos et al., 2003 and Meshko et al., 2001). For example, amongst chemical methods, Milstein et al. (1991) reported 75%, 59% and 80% removal of Adsorbable organic halides (AOX), COD and color, respectively, with the help of a mixture of polyethylene and modified starches (Savant, 2006). Using ultrafiltration, 99% reduction in AOX was achieved by Yao et al. (1994).

11

Chapter 2: Literature review Although these physical and chemical treatment processes might be effective for organic and color removal, they use more energy and chemicals than biological treatment processes, and chemical wastes are also generated from these processes (Sirianuntapiboon et al., 2006). In general, physical and chemical treatment technologies for removal of toxic compounds are uneconomical when applied alone in field scale operations. Additionally, physical and chemical treatment methods such as flocculation, adsorption and oxidation are not effective in removing the high concentration of toxic organics from the wastewater (Zhang et al., 2006). Chemical oxidants such as permanganate and hypochlorite may be applied when wastewater concentrations are too high for acclimatized biological treatment (Savant, 2006). The chlorine and hypochlorite generated by electrolysis of chlorides can be used to oxidize organic compounds in the wastewater. 2.1.3 Advanced oxidation processes (AOPs) 2.1.3.1 Ozonation The conventional municipal wastewater treatment does not fully eliminate toxic organic pollutants. It has been shown that this may be achieved upon treating the effluents with ozone (Huber et al., 2003, Huber et al., 2004, Huber et al., 2005 and Ternes et al., 2003). Ozonation has been shown to have a high potential for the oxidation of various types of toxic and refractory organics in wastewater discharge (Ternes et al., 2003; Huber et al., 2003, 2005). Ozone can generate OH radicals through the formation of an ozonide radical (O3) in the adsorption layer (Equation 2.1). The generated O3 species rapidly reacts with H+ in the solution to give HO3 radical, which evolves to give O2 and OH as shown in Equation 2.2 and 2.3 below (Agustina et al., 2005):
O3 + e O3

(2.1) (2.2) (2.3)

O3 + H + HO3

HO3 O2 + OH

In wastewater, O3 doses ranging from 5 to 15 mg/L led to a complete disappearance of most of the pharmaceutical organics (Huber et al., 2003; Huber et al., 2005). Ozonation is also considered as an effective process for the removal of 12

Chapter 2: Literature review ammonium nitrogen. The application of 15.4 mg/L ozone dose resulted in a nitrogen reduction of up to 35% (Petala et al., 2006). Pretreatment with ozone or peroxide hydrogen also enhances the biodegradability of wastewater such as kraft mill caustic extraction effluents (Savant et al., 2006). Hostachy et al. (1997) reported complete detoxification and removal of residual COD of bleached kraft mill effluent at low ozone doses of 0.51.0 mg/L. Mobius and Helble (2004) have described a combination of ozonation followed by biodegradation in a biofilm reactor to achieve a far reaching elimination of adsorbable organic halides (AOX), color and other toxic organic substances. A major limitation of the ozonation process is the relatively high cost of ozone generation process coupled with a very short half-life period of ozone. Moreover, the process efficiency is severely dependent on the efficient gas liquid mass transfer, which is quite difficult to achieve due to the low solubility of ozone in the aqueous solutions (Gogate and Pandit 2004). 2.1.3.2 Photolysis The photo-activated chemical reactions are characterized by a free radical mechanism initiated by the interaction of photons of a proper energy level with the molecules of chemical species present in the solution, with or without the presence of the catalyst (Bhatkhande et al., 2002; Agustina et al., 2005). A major advantage of the photo-catalytic oxidation based processes is the possibility to effectively use sunlight or near UV light for irradiation. This should result in considerable economic savings especially for large-scale operations (Bauer, 1994; Yawalkar et al., 2001; Agustina et al., 2005). Moiseev et al. (2004) have also shown that the use of photo-catalytic oxidation as a pretreatment step enhances the biodegradability of wastewater containing recalcitrant or inhibitory compounds and is an alternative to a long and energy intensive total pollutant mineralization. A major problem in the successful application of photo-catalytic oxidation is the non-uniform irradiation of the catalyst. This would result in a substantial loss of the incident energy and hence the overall economics may not be favorable. Groups of Ray Beenackers (Ray 1999; Mukherjee and Ray, 1999), Li and Yue (1998a,b, 1999) have tried to address this problem, but not much success has been reported yet. Still the efficiency of photo-catalytic oxidation cannot be underestimated as there is also 13

Chapter 2: Literature review large potential for the combination of photo-catalytic oxidation with other advanced oxidation processes (Agustina et al., 2005). 2.1.3.3 Hydrogen peroxide/Ultraviolet light It is well known for some time that Fentons reagent, a mixture of Fe2+ salts with hydrogen peroxide (H2O2), can easily oxidize organic compounds. However, Fentons reaction has been applied for wastewater treatment only in recent years (Bidga 1995; Chamarro et al., 2001; Lee and Hosomi 2001). This reagent is an attractive oxidative system, which produces in a very simple way OH radicals (Equation 2.4) for wastewater treatment. Iron is a very abundant and non-toxic element and hydrogen peroxide is easy to handle and is environmentally safe. Furthermore, it was found that the reaction can be enhanced by UV/VIS light (artificial or natural), which is named as the photo-Fenton reaction that produces additional OH radicals and leads to the regeneration of the catalyst (Equation 2.5) (Oliveros et al., 1997; Krutzler and Bauer 1999; Fallmann et al., 1999; Arana et al., 2001). These reactions are known to be the primary actions of the photochemical selfcleaning of atmospheric and aquatic environment (Faust and Zepp 1993; Kositzi 2004).
Fe 2+ + H 2 O2 Fe 3+ + OH + OH Fe 3+ + H 2 O + hv Fe 2+ + H + + OH

(2.4) (2.5)

2.1.3.4 Limitations of AOPs Advanced oxidation processes (AOPs) are frequently selected as a treatment option to oxidize toxic organic compounds present in contaminated waters. Such oxidation reactions are based on radical formation, e.g. OH (Valds 2006). OH is highly reactive and non-selective, with the reaction rate constants with organics generally in the order of 105109/Ms (Valds 2006). However, the presence of radical scavengers such as carbonates, bicarbonates and natural organic matter impairs OH formation. These scavengers terminate the radical-type chain reactions, inhibiting propagation reactions and consequently reduce the application of AOPs in industrial scales (Valds 2006).

14

Chapter 2: Literature review In general, a combination of several methods gives high treatment efficiency compared with individual treatment. For example, a certain organic compound can hardly be degraded by ozonation or photolysis alone and the treated wastewater may be more dangerous as a result of ozonation (Lambert and Graham 1995; Camel and Bermond 1998). However, a combination of several treatment methods, such as O3/H2O2/UV (ozone/hydrogen peroxide/ultraviolet), and UV/H2O2 (ultraviolet/hydrogen peroxide), can effectively improve the removal of organic pollutants from the wastewater (Hu et al., 1999; Plummer and Edzwald, 1998). Nonetheless, a majority of these AOP oxidation technologies are still too complex and expensive and less effective for actual application in wastewater treatment (Agustina et al., 2005).

2.2 Electrochemistry in wastewater treatment 2.2.1 Development of electrochemical wastewater treatment Dimensionally Stable Anodes (DSA) have been introduced for the electrochemical technology for several decades. DSA are prepared by the deposition of a thin layer of metal oxides on a base metal, usually titanium. Currently, RuO2 and IrO2 are widely used surface coating materials. One of the most stringent reasons for the success of DSA is the hydroxide formation. Transition metal oxides are among the most versatile electrocatalysts (as much as catalysts) ever known. These electrodes, made for chlorine evolution, have soon shown their activity for almost all other more common electrocatalytic reactions, such as O2 evolution, O2 reduction, even H2 evolution, as well as organic oxidation. The reason for activity has been identified in the surface redox reactions taking place at transition metal ions that act as active sites (Trasatti, 2000). These redox centers, coupled with the moderate interaction of intermediates with the oxide surface, are able to catalyze a variety of electrode reactions occurring in a potential range close to that of the surface redox couple. The same approach by Retschi and Delahay, based on metals, had only revealed a linear variation of overpotential with the enthalpy of hydroxide formation (Trasatti, 2000). Due to their high overpotential for water electrolysis large enough electrochemical windows are gained to produce hydroxyl radicals (Gerger and Haubner 2005). It was shown that organic carbon can be completely converted to 15

Chapter 2: Literature review carbon dioxide without the formation of by-products. Even with a long chain structure, organic molecules can be mineralized without the accumulation of intermediate products. These properties of new DSA electrodes come up to the requirements for an efficient electrochemical wastewater treatment (Gerger and Haubner 2005). 2.2.2 Application of electrochemical technologies in wastewater treatment 2.2.2.1 Electro-coagulation, electro-flocculation and electro-flotation Electro-coagulation, electro-flocculation and electro-flotation are the

electrochemical techniques which may be used separately in different reactors or combined in the same reactors. They have been studied for treatment of municipal and industry wastewater and recovery of chemicals (Buso et al., 1997; Jiang et al., 2002; Chen, 2004), such as removal of phosphate from aqueous solutions (Bekta et al., 2004), olive oil mill wastewater treatment (Inan et al., 2004), removal of sulfide, sulfate and sulfite ions (Murugananthan et al., 2004), and treatment of metal finishing effluents (Khelifa et al., 2005). However, the high cost normally involved in these processes limits their widespread application, especially in the case of low pollutant concentrations (Buso et al., 1997). 2.2.2.2 Electro-chlorination for wastewater treatment Dimensionally stable anodes have been used in chlorine production for about 30 years, and oxide electrodes for chlorine evolution in acidic chloride solutions have been well studied (Cornell et al., 2003). Many industrial wastewaters have been found to be of highly saline content (in terms of NaCl). With a high saline content, the wastewater has a rather high conductivity. To take advantage of this characteristic, a plausible alternative for dealing with saline wastewater is to use the electrochemical methods. The method of electro-chlorination has been successfully employed for treating various types of industrial wastewater (Lin et al., 1998). However, the main drawback of the indirect electrochemical organic degradation by chlorination is the formation of chlorinated compounds. Lin and Wang (2003) had found the formation of chlorinated organic compounds using chromatography while treating anionic surfactants. 16

Chapter 2: Literature review Indirect electro-oxidation, through hypochlorite formation, is a possible answer to an ammonianitrogen problem, through converting it into gaseous nitrogen (Vanlangendonck et al., 2005). The process is pH sensitive, and Lin and Wu (1996) repoeted that while the acidic environment favored the nitrite removal, the alkaline environment appeared to be beneficial to the ammonia removal. 2.2.2.3 Electro-chlorination for wastewater disinfection Electro-chlorination has been increasingly used for wastewater disinfection. In electro-chlorination, a stream of seawater is electrolyzed to produce a flow of chlorine solution that is injected into a wastewater flow. It also has been reported that direct electrochemical disinfection can destroy a wide variety of microorganisms from viruses to bacteria and larger species such as Euglena (Stoner and Cahen, 1982; Patermaraxis and Fountoukidis, 1990; Grahl and Markl, 1996; Park et al., 2003). During direct EC disinfection, all wastewater is forced through a disinfector that is equipped with electrodes on which current is applied. The EC process has the potential to be developed as a cost-effective and environmental friendly alternative of disinfection, particularly for saline sewage effluents encountered in Hong Kong (Li et al., 2002).

2.3 Electro-catalytic oxidation of toxic organics 2.3.1 Electrochemistry for toxic organic degradation Electrochemical treatment for the purpose of treating domestic wastewater was evaluated by Feng et al. (2003). The authors used a pilot reactor with a treatment capacity of 0.3 m3/h. A current density of 3 mA/cm2 was applied with 500 V pulse voltage at a frequency of 25 kHz. The results show that TN, NH4N, TP and COD could be removed from domestic wastewater effectively by the electrochemical treatment system. On the other hand, Vlyssides et al. (2002) applied an electrochemical treatment to domestic wastewater using 0.8% (w/v) sodium chloride as electrolyte and an electrolytic cell with Ti/Pt as the anode and stainless steel 304 as the cathode. Organic carbon and nitrogen content of the wastewater was wet oxidized to carbon dioxide and nitrogen gas. Organic phosphorus content of the wastewater 17

Chapter 2: Literature review was removed by precipitation. COD, volatile suspended solids (VSS), ammonia nitrogen and total phosphorus were reduced by 89%, 90%, 82% and 98%, respectively (Yonar et al., 2006). Many researches have demonstrated that electrochemical technologies can be efficiently applied to the treatment of wastewater containing organic pollutants (Polcaro and Palmas 1997; Bonfatti et al., 1999; Kim et al., 1990; Gu et al., 2006). Up to now, many works have been done to develop high performance anodes for a high catalytic activity and long service life. Examples of the anodes for the treatment purpose include glassy carbon electrode, carbon felt, Pt/Ti and graphite (Polcaro and Palmas 1997; Vlyssides et al., 1999; Greet et al., 2003; Gu et al., 2006). Complete oxidation of linear alkyl surfactants (LAS) and alkyl benzene surfactants (ABS) at bipolar dimensional stable anodes also has been studied (Panizza et al., 2005). The authors achieved a complete surfactant removal with an electrolyte addition of 0.05 M NaCl at a current density of 16.8 mA/cm2 in conjunction with chemical coagulation (Panizza et al., 2005). Bidga (1995) performed an electrochemical degradation test with carbon electrodes and achieved 82% surfactant removal on the total organic carbon (TOC) basis. 2.3.2 Dimensionally stable anodes (DSA) for organic oxidation Electrochemistry is an treatment alternative for organic degradation that has attracted considerable research attention (Comninellis, 1994; Rodgers et al., 1999; Brillas et al., 2000). Toxic organics can be effectively oxidized by electrochemical (EC) reaction (Brillas et al., 2000; Ktz et al., 1991; Chiang et al., 1995; Johnson et al., 1999; Feng and Li, 2003). With unique features such as simplicity and robustness in structure and operation, the EC process has the potential to be developed as a costeffective technology for the treatment of aromatic pollutants, particularly for lowvolume applications. The effectiveness of EC wastewater treatment depends on the nature of the anodes (Feng and Li, 2003; Stucki et al., 1991; Chen, 2004). Traditional electrodes, such as graphite and nickel, show a poor current efficiency in organic degradation (Rodgers et al., 1999). DSAs are found to have a varying degree of success. The commonly used RuO2 and IrO2-based coating surface does not appear to have a high reactivity for organic oxidation (Comninellis, 1994; Feng and Li, 2003; Simod and 18

Chapter 2: Literature review Comninellis, 1997). Other DSA coating materials, such as PbO2- and SnO2-based coatings, have been introduced to improve the treatment performance. PbO2 electrodes can be highly effective for complete organic destruction, e.g., EC incineration (Comninellis, 1994; Feng and Johnson, 1991; Houk et al., 1998; Schumann and Grundler, 1998). Hence, PbO2 and similar anodes are considered to be of sufficient electro-catalytic capacity for organic oxidation (Comninellis, 1994; Feng and Li, 2003; Feng et al., 1995; Tahar and Savall, 1999; Tanaka et al., 2002). However, concern over the possible toxicity of Pb leaching from the working anode would hinder the actual application of PbO2 electrodes. The SnO2-based coating has shown a similar reactivity as that of PbO2 for EC organic degradation (Rodgers et al., 1999; Ktz et al., 1991; Correa-Lozano et al., 1997; Polcaro et al., 1999). With further development and characterization, SnO2 anodes are expected to offer a better solution for the enhancement and application of the EC process to organic degradation in wastewater treatment. 2.3.3 Preparation of Ti-based catalytic active anodes 2.3.3.1 The substrate base material: Titanium Titanium has recently found increasing use in industrial application due to its high strength to weight ratio and excellent resistance in a variety of environment. Titanium is used in the chemical industry not only as a construction material but also for the preparation of dimensionally stable anodes for the production of chlorine, chlorate and other chemicals (Krysa and Mraz, 1994). The preparation of dimensionally stable anodes consists of a coating (painting) of a surface layer of metals on the titanium surface. To achieve good adhesion of the coating, the surface of the base metal must be pretreated to an appropriate degree of roughness. Etching in concentrated HCl is commonly used as a pretreatment of Ti. Kryasa and Mraz (1994) also confirmed that the real surface area after etching by HCl is larger than treated by other acids such as HF. 2.3.3.2 Antimony doped tin oxide Pure SnO2 is an n-type semiconductor with a band gap of about 3.5 eV. The valence band arises due to the overlap of filled oxygen 2p levels. The tin 5s states are 19

Chapter 2: Literature review at the bottom of the conduction band. This kind of oxide exhibits a very high resistance at room temperature and thus cannot be used as an electrode material directly. However, its conductivity can be improved significantly by doping Ar, B, Bi, F, P and Sb. In electrochemical application, Sb is the most common dopant of SnO2. Doped SnO2 films are usually used as transparent electrodes in high efficiency solar cells, gas detectors, far IR detectors and transparent heating elements (Chen, 2004). Kotz et al. (1991) first reported anodic oxidation of pollutants on Sb-doped SnO2-coated titanium electrodes (Ti/SnO2-Sb2O5). The current efficiency obtained on Ti/SnO2-Sb2O5 was about five times higher than that on Pt. Comninellis (1994) measured the current efficiency (CE) of SnO2-Sb2O5 to be 0.58 for 71% degradation of phenol from 10 mM, while the values for PbO2, IrO2, RuO2 and Pt were, respectively, 0.18, 0.17, 0.14 and 0.13 at I = 500 A/m2, pH = 12.5 and temperature = 70C. Grimm et al. (2000) investigated phenol oxidation on SnO2-Sb2O5 and PbO2 using a cycle voltammetric method and also found that the former was more active. Despite the high efficiency for pollutant oxidation, SnO2-Sb2O5 electrodes lack sufficient electrochemical stability just like PbO2 (Chen, 2004). Lipp and Pletcher (1997) conducted a long term test of SnO2Sb2O5 in 0.1 M H2SO4 solution at a constant potential of 2.44 V versus normal hydrogen electrode (NHE) and found that the current dropped from initial 0.2 to about 0.1 A within a few hours and to 0.06 A after 700 h. Correa-Lozano et al. (1997) also examined the stability of SnO2-Sb2O5 electrodes and found that the service life of Ti/ SnO2-Sb2O5 was only 12 h under an accelerated life test performed at a current density of 1000 A/m2 in 1 M H2SO4 solution. Addition of IrO2 into the SnO2Sb2O5 mixture increased the service life significantly. However, the resulting electrodes had a reduced overpotential of O2 evolution at 1.5 V versus NHE in 0.5 M H2SO4 electrolyte. 2.3.3.3 Sol-gel coating and thermo-decomposition Sb doped SnO2 can be effective electrode coating materials for the oxidation of hazardous organics in water (Comninellis, 1994). The conductive SnO2 films can be prepared by chemical vapor deposition (Kadam et al., 1990), reactive sputtering (Czapla et al. 1989), spray pyrolysis (Popova et al. 1990; Czapla et al. 1989), sol-gel, and brush-dry-bake techniques. The dip-coating sol-gel method has previously been used for the preparation of catalytically active membrane film (Maier, 1993). The 20

Chapter 2: Literature review method has major advantages (Lange et al., 2000): such as a high purity of the coating materials, low cost and easiness of coating large and complex-shaped substrates. Very high quality SnO2 layers with respect to structure and morphology can be obtained on the Ti surface by the sol-gel dipping technique. Chemical modifications for the enhancement of the electrode catalytic activity also can be achieved readily by the addition of doping elements. Moreover, the film provides a porous structure which enlarges the catalytic surface area of the film. In general, the preparation of electrodes uses a standard DSA procedure, which consists of the thermal decomposition of the suitable inorganic precursor salts onto a metallic Ti base. The precursors was prepared by a sol-gel solution and coated on the substrate. The coating film is then fixed on the surface by oxidation at high temperature in an oven. A direct relationship between the preparation method and the amount of SnO2 loaded on the electrode surface has been suggested (Comninellis and Battisti, 1996; Osimond et al., 1997).

2.4 Reaction pathways and mechanisms of electrochemical degradation of toxic organics 2.4.1 Degradation pathway of aromatic compounds It is generally agreed that the industrial organic pollutants initially present in the wastewaters are not directly oxidized to CO2, but to various organic intermediate products (Thornton et al., 1992; Ding et al., 1995; Duprez et al., 1996; Lin et al., 1996; Manzanares et al., 1996). In many cases, the end products of the EC oxidation are organic acids of short chains, such as acetic acid, formic acid, oxalic acids and others independently of what initial organic compounds were present in the wastewater. Total mineralization of the refractory chemical compounds can be often prohibitively expensive. Among the harmful organic compounds in industrial wastewaters, phenol and phenolic substances have attracted much attention in the last two decades. This is due to their toxicity and the frequency of industrial usage and discharge. Moreover, phenol is considered to be an intermediate product in the oxidation pathway of higher molecular weight aromatic hydrocarbons (Ding 1995). The oxidation route of phenol has been the subject of research work in the last two decades. Most of the papers try 21

Chapter 2: Literature review to identify the intermediates and to propose reaction schemes of different complexity (Ding et al., 1995; Duprez et al., 1996; Pintar et al., 1994; Levec et al., 1995; Maugans et al., 1997; Alejandre et al., 1998). The kinetics and mechanisms of phenol degradation by EC oxidation have been the subject of several investigations (Duprez et al., 1996; Rodgers et al., 1999; Feng and Li, 2003; Comninellis and Pulgarin, 1991; Gattrell and Kirk, 1993; Andreescu et al., 2003; Lund and Baizer, 1991; Iniesta et al., 2001). 2.4.1.1 EC degradation of phenol The routes of phenol oxidation usually consider the hydroxylation of phenol to hydroquinone and catechol as the first step with a further oxidation to benzoquinones. Eisenhauer (1964) has proposed the oxidation mechanism for phenol as the hydroxylation on the ortho-position (Chu, 1995). The hydroxyl radical attacks the ortho position to form the catechol, and further oxidized into 1,2-benzoquinoe. The further consideration is that phenol oxidation begins with an electron transfer that leads to phenoxy radical reactions (Lund and Baizer, 1991). Possible reactions related

Figure 2.1. Summary of the reaction mechanisms of electrochemical phenol degradation. (Santos et al., 2002) to phenoxy radicals include radicalradical coupling, radical disproportionation,

22

Chapter 2: Literature review radical elimination, and radical oxidation of cations. Radical reactions result in the formation of benzoquinone, which is believed to be an important intermediate of phenol degradation (Comninellis and Pulgarin, 1991; Iniesta et al., 2001). In the subsequent stage, benzoquinone can be degraded with ring breakage to various carboxylic acids. Benzoquinone, a diol-compound, is directly oxidized to muconic acid, which realized the ring cleavage (Chu, 1995). Several mechanisms have been proposed for benzoquinone degradation. If benzoquinone is adsorbed onto the anode surface and gives up an electron, then the carbon that is double-bonded with the oxygen will be attacked by a neighboring hydroxyl radical generated from water electrolysis (Li et al., 2005). When this process is repeated at the para position, the ring could be opened and the benzoquinone would be broken down into small organic molecules such as carboxylic acids (Houk et al., 1998; Lund and Baizer, 1991). A summary of these electrochemical phenol oxidation schemes can be found in Figure 2.1 (Santos et al., 2002). 2.4.1.2 EC degradation of chlorophenol EC oxidation of chlorophenol is also a radical involving repulsive process, and the hydroxyl radicals attack 2-chlorophenol to form chlorodihydroxy-cyclohexadienyl as confirmed by Metelitsa in 1971. The degradation of chlorophenol could be split into three routes while the hydroxyl radical generation is a continuous actin (Rodgers, 2000). Firstly, it is the dimerization and disproportionation. In an oxygen saturated system, dimerization and polymerization are not the major reactions comparing with ring oxidation (Chu, 1995). Hydroxylation on the ortho position is the second rout. Chlorophenol degradation then follows the same pathway as phenol oxidation mentioned above (Eisenhauer, 1964; Chu, 1995). The chlorocatechol, product of hydroxylation will degrade quickly to chloromuconic acid via chlorohydroquinone (Sedlak and Andren, 1991b, Chu, 1995). Thirdly, the hydroxyl radical attack on the para position is another kind of mechanism. The product is chlorohydroquinone which will rearrange to the ortho form, 3-chlorocatechol and proceed to the mineralization (Chu, 1995). At this step, no differences between hydroquinone and catechol are made and it is assumed that both dihydroxylbenzenes follow the same oxidation route. However, the confirmation of the scheme by stoichiometry and the influence of the variables 23

Chapter 2: Literature review such as temperature, oxygen pressure and catalyst concentration in the formation and disappearance of the different intermediates are still an open question. It has been proposed in some cases (Ding, 1995; Levic et al., 1995) that oligomerization or polymerization reactions occur in the liquid phase yielding non-oxidizable compounds. Once the ring cleavage takes place, the subsequent degradation will convert the aliphatic compound through ethers and simple alcohols, and eventually mineralize to acetic acid, formic acid and carbon dioxide (Devlin and Harris, 1984; Houk et al., 1998). The reactions that follow are oxidation of the unsaturated bonds in aliphatic acids such as muconic acid and fractionation into maleic acid anhydride which later changes into maleic acid. The oxidation proceeds further to break down the C = C bond and decarboxylation to form smaller acids. The continuous oxidation effects by hydroxyl radicals will finalize the carbon dioxide transformation. The more recent research work further confirmed the existence of those complex intermediates. The intermediate maleic acid has been experimentally proven (Johnson et al., 1999; Houk et al., 1998; Comninellis and Pulgarin, 1991). With continuous electrolysis, maleic acid would first be reduced to succinic acid at the cathode, followed by oxidation at the anode to malonic acid and then to acetic acid, and finally to CO2 (Johnson et al., 1999; Feng and Li, 2003; Houk et al., 1998; Iniesta et al., 2001). All these reaction details mainly suggest a general result of hydroxyl radical oxidation in a particular reaction system. However, different mechanisms and different reaction pathways may lead organic oxidation carried out by different active and non-active electrodes. The effects of the anode surface properties on the pathway and mechanism of organic electrolysis have not been well addressed. Further research work should be conducted on the comparison of degradation routes and underlying reaction mechanisms for different types of electrodes. As for the EC organic degradation pathways for different anodes, the formation and destruction of many intermediates are still unclear. Many possible intermediates have not been identified and a more elaborate description has to be drawn. 2.4.2 Mechanisms of electro-catalytic organic oxidation According to the final products after EC treatment, two different pathways for the anodic oxidation of undesirable organics are described in the literature by 24

Chapter 2: Literature review Comninellis and co-workers (Comninellis, 1994; Simninellis and Comninellis, 1997; Grimm, 2000): electrochemical conversion and electrochemical combustion. Electrochemical conversion transforms only the toxic, non-biodegradable pollutants into biocompatible organics, and biological treatment is still required after the electrochemical oxidation. This method has been utilized industrial wastewater treatment as a pretreatment unit, such as tannery waste water and dye wastewater (Polcaro and Palmas 1997; Bonfatti et al., 1999; Kim et al., 1990; Gu et al., 2006). In comparison, electrochemical combustion or incineration yields carbon dioxide and water and no further purifications are necessary. 2.4.2.1 OH generation Under most situations, electrochemical oxidation of organic matters is a free radical involving processes. Electrochemical reaction is based on the decomposition of water, accompanied with the oxidation of organic matters on the anode is the oxygen evolution reaction. According to the two equations below, OH can be effectively generated if the electrode materials present an over-potential that is low for water decomposition and high for oxygen evolution (Polcaro s et al, 2000; Tanka et al. 2002). The hydroxyl radical formation potential from water electrolysis is around 2.80 V (Chen, 2004).
2 H 2 O 2 OH + 2 H + + 2e 2 OH 2 H + + O2 + 2e

(2.6) (2.7)

The generation of hydroxyl radical (OH) can be realized from the electrolytic oxidation of H2O/OH- (Chen et al., 1999). This means that the production of OH should be pH dependent at pH > 7.0 due to oxidation of OH-. For most EC treatment conditions, however, organic oxidation is an acid generation process and the pH value usually is lower than 7.0. The majority of OH is produced by the decomposition of water on the anode surface. 2.4.2.2 Electron transfer It is generally believed that organic compounds in aqueous solution can be oxidized on an anode by direct electron transfer or by oxygen atom transfer (Rodgers

25

Chapter 2: Literature review et al., 1999; Chiang et al., 1995; Polcaro et al., 1999; Iniesta et al., 2001; Kirk et al., 1985; Li et al. 2005). In the electron transfer process, organics are adsorbed on the anode surface and give up electrons to the anode. The current input causes the electron transfer from the organic molecules to the anode surface, resulting in direct oxidation of double-bonds and hydrogenation, and conversion or oxidation of organics. The organic oxidation on electrode Pt follows this pattern. However, the strong adsorption of organic compounds to anode Pt may lead to rapid anode fouling, making it unsuitable to wastewater treatment (Gomes, 2004). With the oxygen atom transfer, it is generally considered that oxygen radicals, especially the hydroxyl radicals generated from water electrolysis, play a critical role in the EC oxidation of organic substances (Comninellis, 1994; Li et al., 2002; Iniesta et al., 2001; Simod et al., 1997). Pollutants can also be oxidized by the electrochemically generated hydrogen peroxide (Brillas et al., 1995; Brillas et al., 1996; 1997; 1998; Chen, 2004). In this system, Fe2+ can be added as catalytic reagent for the initiation of the electro-Fenton reaction. Sometimes, UV exposition and oxygen injection could be utilized to assist. The electrically generated ozone is also reported for wastewater treatment (Stucki et al., 1987; Wang et al., 2006). 2.4.2.3 Direct oxidation There is a widely accepted concept of direct organic oxidation on the anode (Comninellis, 1994). Electro-oxidation of pollutants can occur directly on anodes by generating physically adsorbed "active oxygen" (adsorbed hydroxyl radicals, OH) or chemisorbed "active oxygen" (oxygen in the oxide lattice, MOx+1). The physically adsorbed "active oxygen" causes the complete combustion of organic compounds (R) and the chemisorbed "active oxygen"(MOx+1) participates in the formation of selective oxidation products. Figure 2.2 describes the reaction procedures (Simod et al., 1997, Carmem et al., 2003). In general, OH is more effective for pollutant oxidation than O in MOx+1 (Chan, 2004). MOx is the special metal oxide anode materials. The EC organic degradation begins from the reaction (2.8) to form hydroxyl radicals; reaction (2.11) is the formation of the higher metal oxide; reaction (2.12) is the partial (selective) oxidation of the organic compounds, R, via the higher metal oxide; reaction (2.13) is for oxygen evolution by chemical decomposition of the 26

Chapter 2: Literature review higher metal oxide; reaction (2.9) is the combustion of the organic compound via hydroxyl radicals; and reaction (2.10) is for oxygen evolution by electrochemical oxidation of hydroxyl radicals. The difference between physical adsorption and chemical adsorption is defined by the characteristics of MOx, which defines active anodes and non-active anodes. The non-active anodes have been defined as the electrodes that do not provide any catalytic active site for adsorption and oxidation of reactants and/or products in aqueous media. At non-active electrodes the possible anode reactions are, in principle, outer sphere reactions (when the reactant and product do not interact strongly with the electrode surface) and water discharge (since the electrode is considered to be covered by at least one adsorbed layer of water molecules). Physically adsorption coating electrode is usually a kind of non-active anode, such as Sn-Sb, and through the steps of (2.8), (2.9) and (2.10) to accomplish pollutants oxidation, while this process was also called as direct oxidation (Chu, 1995).
MOx + H 2 O MOx (OH ) + H + + e R + MOx (OH ) n CO2 + nH + + ne + MOx MOx (OH ) MOx + 1 / 2O2 + H + + e

(2.8) (2.9) (2.10)

For the oxidant as OH, a kind of non-selective powerful reagent (Chen, 1999), it leads a complete electrochemical combustion of organic matters. Chemical adsorption is the property of active anodes, such as RuO2 and IrO2. Through the step of (2.8), (2.11), (2.12) and (2.13) the oxidation or conversion of organic pollutants is achieved, which also is called as indirect oxidation (Chu, 1995).
MOx (OH ) MOx +1 + H + + e

(2.11) (2.12) (2.13)

MOx+1 MOx + 1 / 2 H 2 R + MOx +1 MO x + RO

In order to provide more evidence on the description of physical adsorption and chemical adsorption of hydroxyl radicals, Comninellis (1994) measured the concentration of OH accumulated on different kinds of anode materials and found that the Ti/SnO2 prefer to the accumulation of OH on the anode surface, while the concentration was zero on Pt and Ti/IrO2. Tanaka (1998) also found hydroxyl radicals formed and accumulated on the diamond anodes by using p-nitrosodimethylaniline as a selective scavenger. Chemical adsorption causes the oxygen atom transfer to the

27

Chapter 2: Literature review internal of oxide lattice for the formation of the oxidant intermediate MOx+1, which is selective to the objective organic matters. Several years later, the organic pollutant incineration mechanisms mentioned above had been further modified by Simond (1997), who provided a theoretical model on organic oxidation on metal oxide anodes. This model has been confirmed by his subsequent tests on aliphatic acids oxidation on electrode Ti/IrO2 (Simond et al., 1998).

Figure 2.2. Reaction mechanism of anodic organic oxidation.

Although the oxidation mechanism mentioned above has been examined by the detection of surface hydroxyl radicals, the different effects of various anode coating elements on EC reactions have not been well addressed. The nature of many coating metal oxides as a kind of semiconductor materials needs to be taken into account in analyzing the electrochemical behaviors of the anodes. The strong organic adsorption on electrode Pt has been confirmed; however, the adsorption capacities of other metal oxide electrodes remain to be determined. The electron and mass transfers in relation to organic electrolysis at the anode surface also should be considered for investigation of the EC reaction kinetics. Characterization of the electrochemical properties of different types of anode materials and their different performance in organic electrolysis is of great importance to the development of more effective EC

28

Chapter 2: Literature review techniques for toxic organic wastewater treatment as well as to the elucidation of the underlining mechanisms involved in EC organic degradation.

29

Chapter 3: Preparation of Ti-based electrodes

Chapter 3

Preparation of Ti-based electrodes


3.1 Metal substrate and coating materials 3.1.1 Metal substrate: Titanium Titanium was selected as the base substrate for making the electrodes in the present study. Titanium is a metal with unique features such as high strength and durability, anticorrosion, high conductivity and easiness for treatment. It is the most reliable and widely used support metal of electrodes for semi-conductive metal oxide coating. Ti-based electrodes are commonly employed in electrochemical applications, including water and wastewater treatment (Pintar et al., 2001; Habazaki et al., 2002). In this work, Ti plate was used for preparation of different types of anodes coated with RuO2 and SnO2, respectively. The titanium (purity > 99.6%, Goodfellow Cambridge) was supplied in a large plate 1.0 mm in thickness. The Ti plate was cut into small squares with a dimension of 2 2 cm2 and a 10 cm handrail for handling and electric connection. Small Ti plates were polished thoroughly before surface coating. 3.1.2 Dimensionally stable anodes (DSA) Titanium as an anode can be easily oxidized electrochemically to TiO2 in an electrolyte solution. The TiO2 film formed on titanium surface is nonconductive, which makes the Ti electrode no longer useful for any electrochemical purposes. Proper surface coating can prevent Ti electrodes from rapid oxidation and hence largely extend their service life from a few minutes to thousands of hours. The Ti anodes with different types of conductive metal oxide coating layers are not only more stable in chemical application but also bring in different surface properties for electrochemical reactions. The well known DSA (Dimensional Stable Anode), which was first developed 40 years ago, has changed the practice of the chlorine-alkali

30

Chapter 3: Preparation of Ti-based electrodes industry (Trasatti, 2000; Polcaro et al., 2000). In DSA titanium is coated with a layer of RuO2 or IrO2 with or without other dopants. The RuO2 or IrO2 coating layer on Ti electrodes has been found to greatly decrease the overpotential of chlorine evolution in seawater and chloride salt solutions (Lin et al., 1998; Lin and Wang, 2003). 3.1.3 Coating materials 3.1.3.1 RuO2 coating film RuO2 is the most common DSA coating layer used in the alkaline industry. The RuO2 coating has a proven high stability in electrolysis of chloride ions. As a typical coating material, RuO2 was chosen in the present study for making one type of the Ti electrodes. The oxide coating composition strongly affects the electrochemical behavior of the electrode. The oxidation reactions in the anodic zone in an electrolyte containing organics include two main competitive reactions, oxygen evolution and organic oxidation. The traditional RuO2 electrodes are low in chlorine and oxygen evolution overpotentials. For the purpose of organic oxidation in wastewater treatment, the low oxygen evolution potential has a negative effect on the formation of OH free radicals during water electrolysis (Polcaro et al, 2000; Tanka et al. 2002; Chen, 2004). The RuO2 coating film is therefore more favorable to chlorine and oxygen evolution and less favorable to other electrochemical reactions such as organic oxidation. For direct organic electrolysis, the Ti electrode coated with RuO2 may not be the best choice due to its low catalytic activity towards free radical generation and organic oxidation (Vaidya and Mahajani, 2002). 3.1.3.2 SnO2-based coating film A few other metal oxides also may be used for making DSA electrodes. SnO2based coating film has been reported to have a high electrochemical catalytic capability towards OH free radical generation (Comninellis and Pulgarin 1993; Comninellis , 1994). Pure SnO2 itself is nearly nonconductive. However, doping with Sb in the coating solution can dramatically increase the conductivity of the SnO2 film by the formation of a large number of electron holes (Rodgers, 2000). The SnO2-Sb

31

Chapter 3: Preparation of Ti-based electrodes coating layer presents a high oxygen evolution overpotential, which is more suitable than RuO2 for use in electrochemical treatment of organic pollutants. The major drawback of the SnO2-Sb electrode has been its low stability and short service life under anodic polarization. The metal crystal of titanium is difficult to match the SnO2-Sb crystal in size, which causes an increase of the stress at the interface between the two crystals and hence weakens the attachment of the coating layer to titanium (Ulrike, 2003; Krsi et al., 2005). The porous structure and the cracks of the coating film would allow the penetration of active oxygen atoms through the coating layer to reach the Ti surface. The reaction of oxygen with substrate titanium forms TiO2 on the metal surface, resulting in passivation of the electrodes. Thus, deactivation of the SnO2-Sb coating layer is mainly caused by the peeling of the coating materials and the formation of nonconductive TiO2 interlayer (Pilla et al., 1997; Arikawa et al., 1998; Alves et al., 1998). In this experimental study, a modified method was developed to coat SnO2-Sb on Ti plates. The new SnO2-Sb coating layer had an improved stability and performed well in organic wastewater treatment.

3.2 Pretreatment of titanium plates Pretreatment is an important step before surface coating. Pretreatment helps to construct a fundamental Ti base for better coating quality. The quality of pretreatment affects the coating results and the electrochemical performance of the electrode. Generally speaking, pretreatment included three steps (Pilla et al., 1997; Arikawa et al., 1998; Alves et al., 1998; Vicent et al., 1998; Zanta et al., 2003; Chen et al., 2004 etc.): Step 1 was to remove the surface TiO2 film by physical polishing. Pure titanium can be easily oxidized in air, and the surface dioxide is inert for the protection of internal metal. TiO2 crystallites in a rutile type structure are quite inert towards chemical etching with common acid or alkalis. Commercial titanium plates usually carry this dioxide surface film. In the pretreatment, a sandblasting roller and grand paper (Silicon carbird, P-220, Lynx Abrasives; Silicon carbird, 400CW, 800CW, Flying Parrot) were used to remove the entire surface layer from the Ti plates. The polished plates were then immersed in ethanol solution and cleaned in an ultrasonic bath for 20 minutes to remove all of the metal particles from the Ti plates.

32

Chapter 3: Preparation of Ti-based electrodes Step 2 was to remove oil and grease from the Ti plates. Shipping, polishing and even contacting with hands could contaminate the plate surface. The Ti plates were degreased completely in 40% NaOH at 80oC for 2 hrs. The plates were then retrieved from the alkaline solution and washed thoroughly with DI water. These procedures should be carefully handled using a forceps to avoid recontamination. Step 3 was acid etching of the Ti plates. There were two purposes for this surface etching step. One was to compensate the physical polishing step for complete removal of the TiO2 layer from the surface. The other was to modify the surface morphology of the Ti plates. A rough and porous surface structure after acid etching could effectively enhance the attachment of a coating layer on the plate surface and to ensure the conductivity of the finished electrode. The Ti plates after previous treatment were etched in 15% oxalic acid boiling at 98oC for 2 hrs. The plates were then washed thoroughly with DI water. The treated Ti plates became gray with a rough surface under a microscope (Fig 3.1) and lost their metallic sheen. The well treated Ti plates were immersed in ethanol solution before surface coating.

Figure 3.1 Surface morphology of Ti plate: (A) under microscope. (B) SEM photo.
It should be noted that an improper degree of acid etching could cause various adverse effects on the quality of subsequent Ti surface coating (Krysa and Mraz, 1994). Strong acid etching may cause a high level of passivation of the coated Ti electrodes. A shallow etching, however, would lead to a loose attachment of the coating solution on the Ti plates and possibly incomplete coverage of the plate surface by the coating materials. The naked and uncovered points of the Ti surface would then allow rapid titanium oxidation and TiO2 formation in the rest coating procedures. In general, weak acid itching at room temperature may not be able to achieve the

33

Chapter 3: Preparation of Ti-based electrodes desired degree of surface roughness. Boiling oxalic acid is therefore often used for acid etching of Ti plates (Arikawa et al., 1998; Alves et al., 1998; Montilla et al., 2004).

3.3 Preparation of Ti/RuO2 and Ti/IrO2 electrodes The Ti/RuO2 electrodes were prepared by the thermal decomposition of RuO2 on the Ti plates. A fully polished Ti plate was dipped into a coating solution (Table 3.1) that consisted of 0.5 M RuCl3 (Aldrich) in n-butanol (99+%, Aldrich) with 10% by volume of concentrated HCl (37%). The wet coating surface was dried in an oven at 60 C for 30 min. After repeating the dipping and drying ten times, the Ti plate was heated in airflow at an annealing temperature of 450 C in a muffle oven for 1 hr. The above procedure was repeated three times to produce the Ti/RuO2 electrode. Table 3.1 Coating solution for the Ti/RuO2 electrodes. Chemical 1 2 3 RuCl3 HCl (37%) n-butanol 10% Solvent (no water) Concentration 0.5 M Comment Precursor salt

Table 3.2 Coating solution for the Ti/IrO2 electrodes. Chemical 1 2 3 IrCl3 HCl (37%) n-butanol 10% Concentration 0.1 M 48 hrs Solvent (no water) Comment Precursor salt

A similar thermal decomposition procedure was followed to make Ti/IrO2 electrodes. Similar to Ti/RuO2 electrodes, Ti/IrO2 is another popular type of DSA used in industry. The IrO2 coating solution contained 0.1 M IrCl3 (Aldrich) in nbutanol (99+%, Aldrich) with 10% by volume of concentrated HCl (37%). A Ti plate was dipped into the solution and then dried in an oven at 60 C for 30 min. After repeating the dipping and drying 5 times, the Ti plate was heated in airflow at 450 C

34

Chapter 3: Preparation of Ti-based electrodes in a muffle oven for 1 hr. The above procedure was repeated 3 times to produce the Ti/IrO2 electrode.

3.4 Electroplating for the inner layer of Ti/SnO2-Sb-Al electrodes Coating of SnO2-Sb oxides on Ti plates consisted of two major phases. The first was the inner layer coating by electroplating, followed by annealing at a high temperature. The inner layer electroplating on the Ti plates is essential for the improvement of the durability and stability of the SnO2 coating layer on the Ti anodes (Li et al., 2005). The second phase was the conventional thermal decomposition after dipping and brushing the coating materials onto the Ti plates for the outer layer. 3.4.1 Electroplating solution The sol-gel method was used to prepare the coating solutions by dissolving the metal chemicals into an organic solvent. For electro-deposition, the coating solution in 100 ml of ethanol contained 17.5 g SnCl45H2O (Aldrich), 0.73 g Sb2O3 (Aldrich) and 2 ml concentrated HCl (37%) (Table 3.3). SnCl4 and Sb2O3 had a molar ratio of 10:1. Sb2O3 could not be dissolved directly into ethanol. HCl (37%) was added in proportion to help dissolve Sb2O3. Ti(BuO)4 was added at a concentration of 3% by volume to increase the viscosity of the coating solution during electroplating. The final solution appeared a color of lemon yellow. Table 3.3 Coating solution (100 mL) for inner layer electro-plating of the Ti/SnO2Sb-Al electrodes. Chemical 1 2 3 4 3 SnCl45H2O Sb2O3 HCl Ethanol Ti(BuO)4 3% Concentration 17.5 g 0.73 g 2 ml Solvent Comments Sn = 0.25 M Sn:Sb = 10:1

3.4.2 Cathodic electro-plating

35

Chapter 3: Preparation of Ti-based electrodes

A well treated Ti plate was immersed in 100 ml of the coating solution placed in a 150-ml beaker for electroplating (Figure 3.2). The Ti plate was a cathode which faced two anodes on both sides of the Ti cathode. Ti/RuO2 electrodes were used as the counter anodes for their high stability and low possibility of contamination by the organic in the coating solution. A constant current density 40 mA/cm2 was charged and sustained for 20 min for electro-deposition on Sn and Sb on the cathode. In general, a higher current density would result in a larger size of metal oxide crystals to be formed on the substrate surface. The electroplating process was carried out under a static condition without any disturbing and mixing. If the solution was stirred during electroplating, the coating surface would be smoother with a lower roughness. This was not preferred for the inner layer coating of the Ti/SnO2-Sb-Al electrodes.

Figure 3.2 Electro-plating set-up. 3.4.3 Annealing at a high temperature After electroplating, the dark electrode was placed into a muffle oven for hydration and oxidation. Sn and Sb atoms had been deposited on the Ti surface. However, pure metal atoms were not stable and had little catalytic capability in electrochemical reactions. Annealing at a high temperature would transform the metals to catalytic metal oxides of SnO2 and Sb2O5. Crystals were generated from the crystal nucleus and clusters to the crystal island formation, which expanded to a

36

Chapter 3: Preparation of Ti-based electrodes complete crystal layer on the Ti surface. A proper heating temperature was the key for the formation of desired metal oxide crystals. A complete oxidation could not be achieved under a low temperature, whilst a too high temperature would result in overoxidation, instead of formation of metal oxides at the preferred oxidation values, i.e., Sn (IV) and Sb (V). Table 3.4 Electroplating conditions tested and selected. Parameters of electroplating 1 2 3 4 5 6 7 8 Components of plating solution Concentration of plating solution 0.5; 0.25; 0.125; 0.075 Input current 0.02; 0.04; 0.08; 0.16 Electroplating time 5; 10; 20; 120 Heating temperature 200; 400; 500; 600 Heating time Repeat 1; 3 Coating density Comment Sn : Sb = 10 : 1 0.25 mol/L 0.16 A 20 min 500C 120 mins 1 15-20 mg/cm2

3.5 Thermal decomposition of the outer layer of the Ti/SnO2-Sb-Al electrodes 3.5.1 Coating solution A traditional thermal decomposition method was employed to coat the outer layer on Ti electrodes. For the thermal decomposition, the coating solution in 100 ml n-butanol consisted of 30 g SnCl45H2O, 0.8 g Sb2O3, 0.1 g AlCl3 and 2.5 ml concentrated HCl (37%) (Table 3.5). The concentration of Sn in the coating solution was 0.5 M, and the solution contained Sn and Sb with a molar ratio of 15:1. To dissolve Sb2O3 into butanol, 5 ml hydrochloride acid was added. In addition, aluminum was dosed in Al2O3 at a rate of 1% of Sn by weight into the coating solution. It was found that doping Al into the coating solution could improve the

37

Chapter 3: Preparation of Ti-based electrodes stability and reproducibility of the Ti/SnO2-Sb-Al electrodes. The final solution for dipping Ti plates was a clean and viscous liquid without color. Table 3.5 Coating solution (100 mL) for the outer layer of the Ti/SnO2-Sb-Al electrodes. Chemical 1 2 3 4 5 SnCl45H2O Sb2O3 Al2O3 HCl Butanol Amount / Concentration Comments 30 g 0.8 g 0.1 g 37% Sn = 0.5 M Sn:Sb = 15:1 Sn-Sb-Al = 100:7:1 Con. as solvent Solvent (no water)

3.5.2 Dipping and drying procedures Under an ambient condition, a Ti plate with the inner coating layer was dipped into the coating solution and drawn out. The plate was heated over an electric oven (500 W) to evaporate the solvent and dry the surface. This dipping and drying procedure was repeated for 5 times to ensure sufficient decomposition of the coating materials on the Ti surface. 3.5.3 Annealing The Ti plate was then placed in a muffle oven for complete hydration and the formation of metal oxide crystals. A temperature program below was designed for the annealing process, including a heat preservation period of 30 min at 100 C to accomplish hydration before the oxidation reactions. Heating at 550 C allows complete metal oxidation and formation of SnO2 and Sb2O5 crystals on the surface of Ti electrodes (Figure 3.2). 25C (ambient ) 8C / 100C (30 min) 8C / 550C (20 min) min min The hydration and oxidation reactions may be written as:
SnCl 4 + H 2 O SnO + 2 HCl + Cl 2 SnCl 2 + H 2 O SnO + 2 HCl 2 SnO + O2 2 SnO2

38

Chapter 3: Preparation of Ti-based electrodes

Sb2O3 430~ Sb2O4 900C


430 900C Sb2O3 ~ Sb2O5

Table 3.6 Thermal decomposition conditions tested and selected. Parameters of thermal plating 1 2 3 4 5 6 7 Components of plating solution Sn : Al = 100 : 0/1/20/50 Concentration of plating solution, Sn Heating temperature I Heating time Heating temperature II Heating time Heating temperature III Heating time Repeating Coating density 0.5 M 450 C 30mins 450 C 30 min 550 C 120 min 3 15-25 mg/cm2 Comments Sn-Sb-Al = 100:7:1

3.5.4 Procedure of the Ti/SnO2-Sb-Al electrode preparation (Figure 3.4).

3.6 Other electrodes used for the experimental research 3.6.1 Platinum electrodes Platinum is a precious metal commonly used for electrochemical studies. In the present research, Pt electrodes were also tested for their effectiveness and

39

Chapter 3: Preparation of Ti-based electrodes performance in electrochemical degradation of organic pollutants, and the results were compared with those of Ti-based electrodes. The Pt plates (purity > 99.95%, Goodfellow Cambridge) 0.1 mm in thickness were cut into squares with a size of 2 2 cm2, which was the same as the Ti-based electrodes. 3.6.2 Ti/RuO2 mesh electrodes used for electrochemical denitrification Traditional DSA electrodes used in the chlorine alkali industry were selected for the experiments of electrochemical denitrification. The electrodes were made of Ti mesh plates rectangular in shape with a dimension of 54 cm2 (Figure 4.3-C). The mesh electrodes were purchased from Beijing Chemical Equipment Company, Beijing, China. According to the manufacturer, the electrode surface was coated with RuO2, TiO2 and ZrO2 by thermal decomposition. Such a treatment on titanium surface has been commercially used to make DSA electrodes with much improved resistance. The Ti/RuO2-ZrO2 mesh electrodes were found to be of a high stability during the experimental studies.

3.7 Characterization of the electrodes 3.7.1 Linear sweep voltammetry (LSV) The electrochemical properties of all three different electrodes (2 2 cm2), Ti/SnO2-Sb-Al, Ti/RuO2 and Pt, were characterized by cyclic voltammetry (CV) and linear sweep voltammetry (LSV). The LSV tests were conducted with a computercontrolled potentiostat (Radiometer Copenhagen/Dynamic-Eis Voltalab PGZ301). The tests were performed at room temperature in a three-electrode cell with stainless steel as the counter electrode and a saturated Ag/AgCl wire as the reference electrode. The electrolyte was 0.25 M Na2SO4 that was the same as used in the organic electrolysis experiments. In addition to the voltammetrical measurements in the electrolyte without organic substances, the LSV of the anodes were also determined with the addition of phenol and chlorophenol at various concentrations such as 1000, 5000 and 10000 mg/L.

40

Chapter 3: Preparation of Ti-based electrodes Most LSV experiments were performed at a scan rate of 50 mV/s over a potential range 0.2 to 2.1 V. During a LSV test, the change in current flowing through the electrodes in the cell was recorded as the potential on the working electrode increased linearly with time. A high enough potential would be reached to reduce/oxidize the substrate in the electrolyte, which could result in an obvious rise in current. The total current still increased with the potential charged until the mass transfer of ions or molecule from the solution to the electrode became rate limiting in comparison to the charge transfer at the surface of the electrode. The maximum current is known as the limiting current. 3.7.2 Surface contact angle measurement The surface property of the electrodes for chemical adsorption was analyzed through the contact angle measurement. The contact angles of a drop of water on the electrode surface, including the wetting, advancing and receding contact angles, were measured following the sessile drop method (Musselman and Chander, 2002). The shape of the water drop was recorded by a CCD camera (PCO1200, PCO imaging) connected through IEEE1394 to a PC and the digital image of the drop was analyzed. The wetting angle of a water drop on a flat surface, the advancing angle at the lower edge and the receding angle at the upper edge of a water drop on an inclined surface were measured. The difference in contact angles between different electrodes suggests their different hydrophobic/hydrophilic surface features and chemical adsorption properties. 3.7.3 Surface morphology 3.7.3.1 Microscopic observations The electrodes at different stages of polishing and coating treatment and after various usage were regularly examined under an optical microscope (BX60, Olympus) equipped with a digital camera (DP10, Olympus) that was connected to a PC. The electrodes were observed directly without any special treatment, and photos were taken for comparison of the surface morphology.

41

Chapter 3: Preparation of Ti-based electrodes 3.7.3.2 Scanning electron microscopy (SEM) Scanning electron microscopy (SEM) (Leica stereoscan 440) was used to examine the coating film on an electrode surface in a greater detail. No special pretreatment was required for the conductive metal samples. A high energy (typically 10 keV) electron beam was employed to scan across the surface for SEM imaging. The Ti electrodes after different durations of electrochemical usage were inspected by the SEM to examine the deterioration of the coating surface. 3.7.3.3 X-ray Photoelectron Spectroscope (XPS) X-ray Photoelectron Spectroscopy (XPS) was performed with an X-ray photoelectron spectroscope (Shimadzu ESCA-3400) using an Al K X-ray source (1486.6 eV of photons). The full scan image provided the element information of the coating materials. 3.7.3.4 Energy Dispersive X-ray (EDX) Analysis EDX was performed with a scanning electronic microscope (JEOL, JSM5600V, Japan). EDX - Energy Dispersive X-ray analysis further provided the proportions of coating materials on titanium plate.

42

Chapter 3: Preparation of Ti-based electrodes

Electrode

Physical polishment HCL, 90C, 2hr Chemical etching Ethanol, 30min Ultrasonic bath

SnCl4, Sb2O3 Ethanol as solvent 0.5mol/L NaF ; Ti(OBu)4 Cathode deposition

Ru as counter electrode Parallar placed for e-plating 20 minutes of e-plating 0.16 A of current input

550C, 120min Annealing

SnCl4, Sb2O3 Sn:Sb:Al=100:7:1 butanol as solvent 0.5mol/L Thermal decomposition

Dipping for solvent attachment Drying for solvent evaporation Baking at high temperature Repeating for 3 times 450C, 120min Annealing

Electrode Figure 3.4 Flow chart of the Ti/SnO2-Sb-Al electrode preparation. 43

Chapter 4: Experimental materials and methods

Chapter 4

Experimental materials and methods


4.1 Electrochemical denitrification 4.1.1 Wastewater samples 4.1.1.1 Saline sludge centrate Sludge liquor was collected from a local municipal sewage treatment works, Shatin Sewage Treatment Works (STW) in Hong Kong, in which the conventional activated sludge process is used. Sludge from both primary and secondary sedimentation tanks are mixed and treated by anaerobic digestion of around 10 days (Figure 4.1). The digested sludge is then dewatered by centrifugation to a solid content of around 30%, while the centrate is returned to the main stream of sewage flow for treatment. The raw sludge centrate collected from the dewatering unit was dark-yellow in color with a NH4+-N content of around 500 mg/L (Table 4.1). Due to the usage of seawater for toilet flushing in Hong Kong, seawater accounts for around 20% of the sewage influent to the Shatin STW. As a result, the sludge liquor samples had a high salinity with a Cl- concentration of more than 5000 mg/L.

4.1.1.2 Landfill leachate Raw leachate samples were collected from the West New Territories (WENT) Landfill Facility, the largest solid waste disposal facility in Hong Kong (Figure 4.2). The landfill has a total design capacity of 60 million m3 solid wastes. It is expected to last for 25 years from the day of commencement in 1999. As it is a major facility for disposal of slaughterhouse wastes, the decay of animal bodies and residues 44

Chapter 4: Experimental materials and methods

45

Chapter 4: Experimental materials and methods gives the leachate high ammonia and COD contents (Table 4.2). The leachate has a chloride concentration of around 3000 mg/L. Currently, hot steam air striping is used to remove ammonia-nitrogen from the leachate, followed by biological wastewater treatment with extended aeration of activated sludge for COD degradation. Table 4.1 Basic water quality parameters of the sludge centrate samples. Parameters NH4 -N BOD COD TOC ClConductivity pH
+

Unit mg/L mg/L mg/L mg/L mg/L ms

Value 518137 16040 22484 6211 5278422 13.052.75 7.700.21

Table 4.2 Basic water quality parameters of the landfill leachate samples. Parameters NH4+-N COD TOC ClConductivity pH 4.1.2 Electrochemical denitrification set-up The EC experiment system for denitrification from the sewage sludge liquor included an EC reactor on a magnetic stirrer, two potentiostats as the DC power suppliers, a pair of titanium electrode plates for electro-denitrification, and sometimes a pair of needle iron electrodes for additional electro-flocculation (Figure 4.3). The EC reactor was a 1-L beaker filled with the waste liquid. The mesh Ti/RuO2-ZrO2 electrode plates described previously with a dimension of 54 cm2 were immersed in the waste liquid. The gap distance between the two electrode plates could be varied Unit mg/L mg/L mg/L mg/L ms Value 3000-5000 13200-21280 2213.5-3714 2800-3500 29.9-31.3 8.1-8.5

46

Chapter 4: Experimental materials and methods

Power supplier

Needle iron electrode Electrode mesh Stirrer Figure 4.3 (A) Electrochemical denitrification system illustration

Figure 4.3 (B) Electrochemical denitrification set-up

47

Chapter 4: Experimental materials and methods

Figure 4.3 (C) Ti/RuO2-ZrO2 electrode mesh plates for different testing conditions. The two power sources, when both were used, could charge the pair of Ti plates and the pair of iron needles separately. The DC input from any of the power suppliers was adjustable by varying both the current (I), and the voltage (V). Because of different treatment purposes served, the power input to the Ti plates for denitrification was much greater than that to the Fe needle electrodes for dissolving iron and electro-flocculation. 4.1.3 Experimental conditions 4.1.3.1 Batch denitrification tests for the sludge liquor The EC treatment experiments were carried out mainly in batch mode. Three gap distances between the Ti plates, 3, 8 and 15 mm, were tested. For each electrode setting, electrolysis experiments were conducted at different current input, including 0.5, 1.0, 1.5, 2.0 and 2.5 A, resulting in different current intensities of 25, 50, 75, 100, 125 mA/cm2 for EC denitrification. For each test run which lasted from 1 hr to more than 2 hrs depending on the current intensity, samples were withdrawn from the

48

Chapter 4: Experimental materials and methods reactor at various time intervals for evaluation of the treatment result, particularly on the removal of total inorganic nitrogen (TIN). 4.1.3.2 Denitrification plus electro-flocculation for organic removal In combination with EC denitrification, electro-flocculation was tested for additional organic removal from the sludge centrate. The pair of iron needles placed 65 mm apart were charged at a low current. With the DC charged, iron dissolved into the solution at a controlled rate by adjusting the current input. Fe2+ was expected to function as a flocculant for particulate organics in the sludge liquor. By the end of EC treatment, the sludge solution was flocculated for 30 min followed by sedimentation of 120 min. In addition to the nitrogen content, the supernatant was analyzed for organic and SS reductions, and the sludge accumulated in sediment was measured. 4.1.3.3 Continuous flow stirred tank reactor (CFSTR) for denitrification of landfill leachate EC denitrification tests also were conducted for landfill leachate using a CFSTR with continuous inflow and outflow. The EC reactor was a 500-mL beaker filled with leachate 300 mL in working volume. Three mesh electrode plates were immersed in the liquid with one anode facing two cathodes. The electrodes had a gap of 5 mm between them, and a high current intensity was charged at 125 mA/cm2. To enhance the EC process, NaCl was added into the raw leachate influent to increase the chloride content from about 3000 to over 4000 mg/L. Leachate was pumped into the EC reactor at four different flow rates, which resulted in 4 different treatment times of 3, 5, 7, and 9 hrs. Samples were withdrawn and analyzed for the results of EC denitrification from the landfill leachate. 4.1.4 Electro-denitrification and electro-flocculation reactions. 4.1.4.1 Electro-chlorination reactions Electro-chlorination is believed to be essential to EC denitrification of the waste liquor. In the presence of Cl-, NH4+ and other impurities, electro-chlorination

49

Chapter 4: Experimental materials and methods for nitrogen removal is a complex process. The following electrolysis reactions are important (Leonard, 2003; Roger, 1959):

2Cl 2e Cl 2 E 0 = 1.36 2 H 2 O 4e O2 +4 H + E 0 = 1.229 2 H + + 2e H 2 E 0 = 0 Cl 2 + H 2 O HOCl + H + + Cl Bess, 2003) as

(4.1)

(4.2) (4.3)

The combined reaction of Cl- electrolysis can be written (Parsons, 1959; Casson and

Cl

+ H

+ H 2 O HOCl + H 2

(4.4)

The principal chlorination reactions in the presence of NH4+ and the resulting denitrification reaction (Metcalf and Eddy 2003) are 2 NH 4 + 3HOCl N 2 +3H 2 O + 5H + + 3Cl Thus, the overall reaction of EC denitrification can be written as
Power 2 NH 4 N 2 +3H 2 +2 H + +
+

(4.5)

(4.6)

4.1.4.2 Electro-flocculation reactions

50

Chapter 4: Experimental materials and methods The pair of iron needles placed 65 mm apart were charged at a low current. With the DC charged, iron dissolved into the solution at a controlled rate by adjusting the current input. Fe2+ was expected to function as a flocculant for particulate organics in the sludge liquor, according to the following reactions.

Fe + 2 H + Fe 2+ + H 2 ,
4 Fe 2+ + 8OH + O2 + 2 H 2O 4 Fe(OH ) 3

(4.7) (4.8)

4.1.5 Analytical methods 4.1.5.1 Chlorine compounds and chloride Chloramines NH2Cl, NHCl2 and NCl3 were determined for their concentrations by the titration method in accordance with the Standard Methods (1998) Section 4500-Cl F, DPD ferrous titrimetric method. Chlorine concentration was determined by the portable datalogging spectrophotometer (DR/2010, HACH) at 530 nm following Section 4500-Cl G: DPD Colorimetric Method in the Standard Methods (1998). Chloride concentration was measured with the same equipment at 550 nm following the methods defined by the manufacturer. 4.1.5.2 Nitrogen (NH3-N, NO2--N, NO3--N) concentrations Total inorganic nitrogen (TIN) in different forms was analyzed. The analytical procedures of NH3-N, NO2--N, and NO3--N followed Section 4500-NH3 D: AmmoniaSelective Electrode Method, Section 4500-NO2- B: Colorimetric Method, and Section 4500-NO3- B: Ultraviolet Spectrophotometric Screening Method in the Standard Methods (1998), respectively. Ammonium nitrogen was analyzed by the electricalchemical method using an ammonia-selective electrode (95-12, Orion) with an electrometer (920A, Orion). An UV-visible spectrometer (Lambda 12, Perkin Elmer, 1cm path length) was used for NO3- and NO2- measurements following the Standard Methods (1998). 4.1.5.3 Trihalomathane (THM) formation potential measurement

51

Chapter 4: Experimental materials and methods

Figure 4.5 Portable datalogging spectrophotometer (DR/2010, HACH)

Figure 4.6 UV/Visible spectrometer (Lambda 12, Perkin Elmer)

52

Chapter 4: Experimental materials and methods Formation of trihalomathanes (THM) after the EC treatment in a sample was assessed following the extraction with pentane. A Hewlett Packard (HP) 6890 series gas chromatograph (GC) equipped with an Agilent J&W DB-5 capillary column (30.0 m 320 m 1.0 mm nominal), an auto-sampler, and a linearized Electron Capture Detector (ECD) was employed for the THM analysis (Figure 4.10). The flow rate was adjusted to 1.1 mL/min with ultra-high pure (UHP) nitrogen as the carrier gas. The initial oven temperature was set at 35 oC and held for 5 min. The temperature was then gradually increased to 70 oC at a rate of 10 oC/min and further increased to 200
o

C at a rate of 20 oC/min. The total running time for the temperature program was 15 C (Li and Chu, 2003).

min. The equipment was ready for the next run when the column cooled down to 35
o

4.1.5.4 Other water quality parameters The input of DC current and voltage could be read directly from the power suppliers. All analyses of conventional water quality parameters, such as the biochemical oxygen demand (BOD), chemical oxygen demand (COD) and the suspended solids (SS), were performed following the Standard Methods (1998). COD measurement followed the Standard Methods (1998), Closed Reflux Method in Section 5310-B, high temperature combustion method. SS was determined according to Section 2540-D, total suspended solids dried at 103-105C. BOD was measured by the OxiTopBOD measuring devices. Total organic carbon (TOC) was measured by a TOC analyzer (TOC-5000A, Shimadzu) using the combustion-infrared method, which followed Section 5310 B: High Temperature Combustion Method in the Standard Methods (1998). The solution pH was measured by a pH meter (420A, Orion), and the turbidity was monitored by a Turbidimeter (2100N, HACH) according to Section 2130 B, Nephelometric Method in the Standard Methods (1998). In addition, A jartest device (PB700, Phipps & Bird) with a flat paddle mixer (3.81.25 cm2) was employed for the flocculation test. 4.1.5.5 Current efficiency of EC denitrification

53

Chapter 4: Experimental materials and methods

Figure 4.7 TOC analyzer (TOC-5000A, Shimadzu).

Figure 4.8. Turbidimeter (2100N, HACH)

54

Chapter 4: Experimental materials and methods The current efficiency of electrolysis for denitrification can be estimated (Prentice, 1991) by
3 F N 14 I t

(4.9)

where: N is the amount of TIN removed from 1 L of the sludge liquor within a treatment time of t at the current I, and F is the Faraday constant (96,500 coulombs).

4.2 Electrochemical degradation of aromatic organic pollutants

4.2.1 Wastewater samples 4.2.1.1 Phenol solution Chemical reagent was used to make artificial wastewater for the EC treatment experiments. The stock phenol solution was prepared by dissolving 0.5 g of the phenol (Aldrich) into 250 ml DI water. The solution was placed in an ultrasonic bath for 30 min for complete phenol dissolution. The stock phenol solution had a concentration of 500 mg/L, which had a TOC concentration of about 383 mg/L. Additionally, a stock phenol solution at a higher concentration of 2000 mg/L was also prepared occasionally. 4.2.1.2 Chlorophenol solution Chlorophenol wastewater was prepared in the same way as described above for the phenol wastewater. However, a longer ultrasonication time of 60 min was needed for complete chlorophenol dissolution. The stock chlorophenol solution had a concentration of 500 mg/L, and the total organic carbon (TOC) concentration was about 280.16 mg/L. 4.2.1.3 Solutions of the intermediate organics of EC phenol degradation Solutions of a number of pure aromatic components and aliphatic acids were prepared for the EC oxidation experiments. These organic chemicals were expected to be the intermediate products of phenol and chlorophenol electrolysis. The aromatic

55

Chapter 4: Experimental materials and methods chemical solutions included hydroquinone (Aldrich), benzoquinone (Aldrich), catechol (Aldrich). The organic acids included muconic acid (IL), oxoglutaric acid (Aldrich), fumaric acid (Aldrich), maleic acid (Aldrich), succinic acid (Aldrich), molonic acid (BDH), oxalic acid (BDH), axetic acid (Aldrich) and formic acid (Aldrich). The stock solutions of these pure chemicals had a concentration of 200 mg/L. All above chemical solutions, if not used immediately, were stored in refrigerator at 4 C before the EC degradation experiments. 4.2.2 Experimental set-up and operational conditions 4.2.2.1 Experimental set-up

Phenol and chlorophenol degradation experiments were conducted in batch electrolysis cells using 100-ml glass beakers. Each had a cover cap with two open slits for assembling the electrodes. As described previously, three different types of the anodes, Ti/SnO2-Sb-Al, Ti/RuO2 and Pt, were investigated through the experimental tests. Together with the working anodes, stainless steel plates of the same size (2 2 cm2) were used as the counter cathodes. The two electrodes in each EC cell were

56

Chapter 4: Experimental materials and methods placed facing each other 8 mm apart. A DC potentiostat was used as the power supply with a voltage output up to 35 V.

Figure 4.10 Gas chromatograph mass spectrum (GC-MS) for THM analysis (HP 6890 )

4.2.2.2 Operation and sampling For a test, 60 ml of a chemical solution at a predetermined concentration was placed in the EC cell. For example, phenol had an initial concentration of 100 mg/L. The solution contained 0.25 M Na2SO4 as the supporting electrolyte. A consistent DC current, e.g., 0.08 A, was charged to the electrode pair, resulting in a certain current intensity, e.g., 20 mA/cm2, between the electrodes. The EC cells were placed on a magnetic stirrer for continuous mixing, and the temperature stayed at around 25oC without special control. Through the electrolysis process, the cell voltage and solution pH were recorded. Samples were withdrawn from the solutions at various intervals for chemical analysis of the phenol or chlorophenol, the total organic carbon (TOC), and intermediate products in the electrolyzed liquids.

4.3 Analytical methods

57

Chapter 4: Experimental materials and methods 4.3.1 High performance liquid chromatography (HPLC) for organic analysis High performance liquid chromatography (HPLC) was used to determine the concentrations of phenol, chlorophenol and the intermediate organics of the EC degradation process. The intermediates included aromatic compounds of benzoquinone, hydroquinone and catechol and aliphatic acids of oxoglutaria acid, maleic acid, fumaric acid, succinic acid, malonic acid, oxalic acid and acetic acid. The HPLC (Waters 2695) was equipped with a UV detector (Waters 2996) used at working wavelengths from 190 to 400 nm. For analysis of phenol and other aromatic compounds, a capillary column in Symmetry C18 5 m (3.9 150 mm, Waters) was used. Methanol/water = 1/1 (V/V) was used as the mobile phase at a flow

Figure 4.11 High performance liquid chromatography (HPLC) (Waters 2695 2996)

rate of 0.5 ml/min under room temperature. For analysis of aliphatic acids the column Atlantis dC18 5 m (3.9 150 mm, Waters) was chosen. NaH2PO4 (0.25 mM) solution with a pH = 2.75 adjusted by H3PO4 was used as the mobile phase at a flow rate of 0.5 ml/min. Before each analysis, the sample was filtered through a 0.45-m

58

Chapter 4: Experimental materials and methods membrane filter (Millex-LCR) in the ambient environment. The sample injection volume was 10 l. Two examples of HPLC analysis of the aromatic compounds and aliphatic acids are given in Figures 4.12and 4.13.

Figure 4.12 HPLC analysis of the four 4 types of aromatic compounds

formic acid 3.467


0.016

malonic acid 4.158 8.166 maleic acid 6.956 oxoglutaric acid 4.661 acetic acid 5.583

fumaric acid

0.014

oxalic acid 2.721


0.012

0.010 AU

0.008

succinic acid 9.409

0.006

0.004

0.002

0.000 1.00 2.00 3.00 4.00 5.00 6.00 7.00 Minutes 8.00 9.00 10.00 11.00 12.00 13.00 14.00 15. 00

Figure 4.13 HPLC analysis of aliphatic acids

59

Chapter 4: Experimental materials and methods 4.3.2 Current efficiency and TOC removal efficiency Current efficiencies were calculated for phenol and chlorophenol destruction and TOC removal by the EC process. The efficiency was used to evaluate the performance and energy effectiveness of different electrodes for electrolysis of toxic organics. The current efficiency for chemical destruction can be obtained from C1 C 2 C1 C 2 = 100 I t Q

(4.10)

where C1- C2 refers to the concentration change of phenol or chlorophenol after a

TOC Liquor GC-MS HPLC

Current C-V SEM Electrode XPS/EXD Contact angle

Figure 4.14 Characterization of different anodes

treatment time of t, I is the current input and Q hence is the total charge input within time t. Similarly, the current efficiency for TOC removal can be calculated by TOC1 TOC 2 TOC1 TOC 2 = 100 I t Q

(4.11)

where TOC1- TOC2 is the TOC change during a time period of t at a current of I.

60

Chapter 4: Experimental materials and methods 4.3.3 Rate constants for EC oxidation of organic chemicals Organic decreased in concentration during the electrolysis process. The destruction of a pure organic by the EC treatment with different anodes was found to generally follow a first order kinetics. Hence, the reduction in the chemical concentration can be described by

dC = kC dt

(4.12)

where C is the chemical concentration after an electrolysis duration of t, and k is the rate constant of the first order oxidation process. The kinetic equation may be transformed to
C ln 0 C t = kt

(4.13)

The chemical usually had an initial concentration of 100 mg/L. Samples were withdrawn from the electrolyzed solution at various time intervals to obtain a series of the chemical concentrations. From a linear regression of ln(

C0

Ct

) vs. t, the rate

constant for EC destruction of a specific organic, phenol, chlorophenol, and their degradation intermediates, can be estimated.

61

Chapter 5: Nitrogen removal from wastewater by electrochemical denitrification

Chapter 5

Nitrogen removal from wastewater by electrochemical denitrification


5.1 General characteristics of the waste liquor samples and EC treatment

In wastewater treatment plants, the sludge liquor from the dewatering unit is returned to the inlet of the treatment system. Although the flow rate of the sludge liquor is low, accounting for about 1% of the influent, it contains a high nitrogen content of more than 500 mg/L NH4+-N. The ammonia loading from the sludge liquor contributes to 10-20% of the influent nitrogen loading (Janus et al., 1997). Leachate from landfills is a complex and high strength wastewater, which is rather difficult and expensive to treat by conventional means. Landfill leachate (in Hong Kong) usually has a NH4+-N concentration of more than 3000 mg/L together with about the same amount of refractory organic pollutants. Inflow of raw leachate into a municipal wastewater treatment plant would considerably increase the loading of the treatment facility. Pre-treatment of landfill leachate is normally required before its flow into a municipal wastewater works. Removing the ammonia input from either the sludge liquor or leachate by a separate treatment will improve the performance and increase the capacity of the mainstream wastewater treatment process for nutrient removal (Jetten et al., 1997). Electrochemical (EC) treatment can be considered for effective treatment of the sludge liquor and landfill leachate. An EC system is simple and robust in structure and operation. During EC treatment the sample water will be forced through a reactor of electrolysis, in which electro-chlorination and the subsequent ammonia oxidization are expected. The high salinity content in the sludge liquor and landfill leachate of Hong Kong will improve the performance and efficiency of the EC process for nitrogen removal. In the present study, laboratory experiments were carried out to investigate the effectiveness of EC treatment for denitrification and organic removals.

62

Chapter 5: Nitrogen removal from wastewater by electrochemical denitrification


5.2 Electrochemical denitrification of sludge centrate liquor

5.2.1 Electrochemical denitrification The EC treatment was demonstrated to be highly effective for denitrification of the saline sludge liquor. Nitrogen measured as total inorganic nitrogen (TIN) was removed continuously from the sludge centrate samples by electrolysis. With a sufficient duration of the EC treatment, complete denitrification could be achieved (Figure. 5.1). An increase in the current intensity resulted in a faster rate of TIN reduction. For example, at a current intensity of 50 mA/cm2 for sludge liquor with an initial NH4+-N of more than 350 mg/L, denitrification could be completed within 150 min. This treatment time was shortened to around 60 min at an elevated current intensity of 125 mA/cm2. Thus, higher current input favoured the kinetics of EC denitrification. As anticipated, electrolysis of the ammonia solution formed intermediate products of chloramines prior to the completion of denitrification. However, there did not appear to be significant accumulation of chloramines, including NH2Cl, NHCl2 and NCl3, during the EC treatment (Figure 5.2). NH2Cl-N could reach a concentration of around 20 mg/L, which disappeared from the solution by the end of EC denitrification. Formation of NO3- and NO2- were hardly detectable during the EC process. Thus, NH4+ was the predominant TIN species throughout the EC treatment.
400 350 50 mA/cm2 75 mA/cm2 100 mA/cm2 125 mA/cm2

TIN concentration (mg/L).

300 250 200 150 100 50 0 0 40 80 Time (min)

120

160

Fig. 5.1 TIN removal as a function of the contact time and current intensity with a gap distance of 8 mm between the electrodes.

63

Chapter 5: Nitrogen removal from wastewater by electrochemical denitrification Chloride in the sludge liquor was electrolysed to chlorine, which was believed to be essential to EC denitrification. While both free chlorine and combined chlorine were present in the solution, free chlorine accounted for the major fraction of total chlorine in the reactor (Figure 5.2). The combined chlorine of NH2Cl and NHCl2 was found in a low and rather stable concentration level. Upon complete NH4+-N removal, the combined chlorine was gradually eliminated, while free chlorine was found to increase with time by continuous electrolysis. The importance of chloride to EC denitrification was also demonstrated with a model wastewater which was a (NH4)2SO4 solution containing 500 NH4+-N mg/L. Without salt (Cl-) addition, electrolysis of the model solution did not result in any effective nitrogen removal. While Cl- was dosed into the solution at a level of 4000
400 350 300 250
N (mg/L)
NH4-N NHCl2 NO3pH NH2Cl NCl3 NO2-

8 7.5 7 6.5 6 5.5 5 0 15 30 45 T ime (min) 60 75


pH

200 150 100 50 0

400 350 300


Cl (mg/L)

Free-Cl2 NH2Cl NHCl2 NCl3 T otal Cl2

250 200 150 100 50 0 0 15 30 45 T ime (min) 60 75

Figure 5.2 (a) Nitrogen removal as a function of treatment time and (b) chlorine compounds during the EC process at a current intensity of 125 mA/cm2 with an electrode gap of 3 mm.

64

Chapter 5: Nitrogen removal from wastewater by electrochemical denitrification mg/L, the same denitrification result that was obtained for the actual sludge centrate was reproduced. Thus, electro-chlorination can be considered as the major mechanism responsible for EC denitrification of the saline sludge liquor (Stoner and Cahen, 1982; Casson and Bess, 2003). 5.2.2 Electro-chlorination for denitrification of the sludge centrate Electro-chlorination is believed to be essential to EC denitrification of the saline sludge liquor. In simplified terms, the overall reaction of EC denitrification is:
Power 2 NH 4 N 2 +3H 2 +2 H +

(5.1)

Electrolysis of NH4+ would lead to an accumulation of H+ ions and thus a continuous decrease of the solution pH, which was observed during the EC process (Figure 5.2a). After the completion of nitrogen removal, with continuous chlorine production, the following electrolysis reaction would become dominant: 2 Cl + 2 H 2 O Cl 2 + 2 OH

+ H2

(5.2)

Thus, a turn point of the pH value was expected, after which the pH would increase with time. This change in solution pH was also demonstrated by the test results (Figure 5.2a). One of the major concerns in use of EC treatment methods is the formation of chlorination by-products (CBPs). As an indicator, trihalomethanes (THMs) were used for assessing the potential CBPs problem in EC denitrification of the sludge liquor. The total THMs in the treated sludge liquor was found to range only from 200 to 300

g/L. The low THMs formation during the electro-chlorination was likely due to the
short contact time of 1 hr or so used. Since the sludge liquor will be returned to the main stream of sewage flow, the low level of THMs in the denitrified liquor will be degraded by the downstream biological treatment. Hence, it is predictable that the accumulation of THMs and other CBPs during EC denitrification will be insignificant, which is not supposed to cause any deterioration in the effluent quality of wastewater treatment. 5.2.3 Power consumption and current efficiency

65

Chapter 5: Nitrogen removal from wastewater by electrochemical denitrification For a given EC system, parameters such as the current intensity, the gap distance between the electrodes and treatment time regulate the effectiveness and performance of EC denitrification. As shown in Figure 5.3, an increase in the current input would accelerate the nitrogen removal process by electrolysis. However, for the same gap distance between the electrodes, a higher current intensity might not improve the efficiency of the current used for denitrification. Instead, a too high current input would decrease the current efficiency (Figure 5.3). In the EC treatment, in addition to NH4+ electrolysis to nitrogen gas, current input was also consumed for electrolysis reactions of H2O, Cl- and other compounds, as well as for heat generation. It is apparent that the fraction of current used for denitrification did not increase with the current intensity, although a higher current favoured the nitrogen removal kinetics. The gap distance of the electrodes had more profound effect on the current efficiency for EC denitrification. Among the 3 distances examined, a gap of 8 mm gave the best current efficiency as high as 65% or more. Increasing the gap to 15 mm resulted in a large reduction of the current efficiency to around 55%. More electricity was likely wasted for heat generation as the gap distance increased, as evidenced by the temperature rise observed. However, when the gap was reduced to only 3 mm, the current efficiency dropped more dramatically to around 40%. Bubbles trapped between the narrow gap could have a negative effect on the EC nitrogen removal

100

Current efficiency (%)

80 60 40 20 0 40 60 80 100 120 140

3mm 8mm 15mm

Current intensity (mA/cm2)


Figure 5.3 Current efficiency as a function of current input for different electrode gaps in EC denitrification.

66

Chapter 5: Nitrogen removal from wastewater by electrochemical denitrification process. Possible short circuit between the electrodes would considerably lower the current efficiency for NH4+ electrolysis. Thus, an appropriate gap distance between the electrodes should be adopted in the EC system to ensure a high current efficiency for denitrification. For a batch operation of NH4+ electrolysis, the amount of nitrogen removal from the sludge liquor depended directly on the power input into the EC treatment system. The power efficiency for denitrification was also affected by the current density and the gap distance of electrodes (Figure 5.4). It could be estimated that the power consumption rate for nitrogen removal from the sludge liquor was about 23 kWh/kg N under the optimum operation conditions. This value is rather comparable with other denitrification processes (William, 1989). In addition, EC denitrification is simple in structure and easy in operation. It has the potential to be developed as a robust and cost-effectiveness technology for separate treatment of saline sludge liquor. 5.2.4 Electro-flocculation for organic removal
400

TIN concentration (mg/L).

350 300 250 200 150 100 50 0 0 5 10

D=3 mm D=8 mm D=15 mm

15

20

Power consumption (kWh/L) Figure 5.4 TIN removed as a function of power consumption for different electrode gaps at the same current intensity of 75 mA/cm2.

The EC process did not appear to be highly effective for the removal of organic materials from the sludge liquor, although a certain extent of reduction in the organic content was observed (Figure 5.5). Some organic molecules could likely be oxidised by EC reactions or the oxidizing products of the EC process. However, EC oxidisation was not expected to be highly efficient for large organic molecules and

67

Chapter 5: Nitrogen removal from wastewater by electrochemical denitrification particulate organics. With the electro-flocculation examined in this study, organic removal of the sludge liquor was improved considerably (Table 5.1, Figure 5.5). Electrolysis with the iron electrodes would dissolve Fe2+ into the sludge solution. Fe2+ and its hydrolysis products are excellent flocculants. They are capable to destabilise the colloidal solution of large organic molecules, particulate organics and other impurities. In the present tests, Fe2+ compounds produced by the EC process functioned as a flocculant to promote flocculation of organic materials. Large flocs formed by electro-flocculation were effectively removed in the subsequent sedimentation. With the EC system, the rate of Fe2+ dissolving into the sludge liquor from the iron anode could be well controlled by regulating the current input. Electroflocculation did not have any adversary effect on EC denitrification. Instead, the combined action of electro-denitrification with electro-flocculation had better treatment results in organic and SS removals (Figure 5.5). The electrolysed effluent after flocculation and sedimentation was transparent and colourless with a rather low turbidity (Table 5.1). It should be noted, however, over-dosage of Fe2+ by the EC reaction did not result in more organic and SS reductions from the sludge liquor, while sludge production was largely increased. In comparison, electro-flocculation without EC denitrification only removed a part of organic and SS and had no function of nitrogen removal (Figure 5.5). Chemical flocculation using FeCl3 in combination with EC denitrification had a similar treatment result as the coupled process of EC denitrification and electroflocculation. However, chemical flocculation lowered the solution pH to around 3 and produced more sludge sediment (Figure 5.5). In actually application, chemical flocculation would require more facilities and operation attention. Therefore, electrodenitrification together with electro-flocculation appears to be the best treatment approach for nitrogen removal and organic reduction of sludge liquor.
Table 5.1 Effect of electro-flocculation on the sludge liquor treatment.

Electroflocculation No Yes

NH4+-N mg/L 0.18 0.12

TOC mg/L 44.4 17.2

COD mg /L 100 44

SS mg/L 6.74 2.89

Sludge mg/L 0.22 441

Turbidity NTU 17.8 0.57

68

Chapter 5: Nitrogen removal from wastewater by electrochemical denitrification

700

TIN concentration (mg/L)

600 500 400 300 200 100 0 20

A C

B D

60 0 50

40 60 T ime (min)

80

100

TOC (mg/L)

40 30 20 10 0
100 0 80

A B C D 20 40 60 80
A B C D

100

T ime (min)

SS (mg/L)

60 40 20 0

3000 0 2500

20

40 T ime (min)

60

80

100

Sludge (mg/L)

2000 1500 1000 500 0 0 20 40 T ime (min) 60 80 100 A B C D

Figure 5.5 Comparison in removals of TOC, COD and SS, sludge production and TIN degradation for conditions of (A) EC-denitrification + EC-flocculation, (B) ECdenitrification only, (C) EC-flocculation only and (D) EC-denitrification + chemical flocculation.

5.3 Electrochemical denitrification of landfill leachate

69

Chapter 5: Nitrogen removal from wastewater by electrochemical denitrification 5.3.1 Batch EC leachate treatment tests The electrochemical treatment was also demonstrated to be highly effective for denitrification and organic removal of the landfill leachate. In batch tests, with a sufficient duration of the EC treatment, complete denitrification could be achieved (Figure. 5.6). Nitrogen measured as TIN was removed continuously from an initial concentration of more than 1600 mg/L by 3 hrs of electrolysis. Adding NaCl into the solution could increase the rate of denitrification. As the initial Cl- content was increased from around 2200 to about 2700 mg/L, the time required for complete EC denitrification was shortened from 3 hrs or longer to 2 hrs. About 30% of Cl- was consumed during the electrolysis process for nitrogen removal from the leachate.

Figure 5.1-A Untreated leachate (left) and EC treated leachate (right).

3000

3000

N
2500

2000

ClN (NaCl) Cl- (NaCl)

2500

2000

1500

1500

1000

1000

500

500

0 0 1 2 3 4

Time (h)

Figure 5.6 TIN removal as a function of the electrolysis time for different chloride concentrations in EC denitrification of the landfill leachate. 70

Cl- (mg/L)

N(mg/L)

Chapter 5: Nitrogen removal from wastewater by electrochemical denitrification The landfill leachate was different from the sludge liquor. The ammonia concentration in the leachate was 6-10 times higher than that in the sludge centrate, while both wastewater samples had a similar level of chloride ions. To achieve 100% EC nitrogen removal, a much longer time was needed for the leachate. As a large amount of chloride ions were consumed by electrolysis, the chloride concentration could drop to a rather low level. This was shown to be rather unfavourable to the reaction kinetics of EC denitrification. When chloride concentration in solution was reduced to 300 mg/L or below, the electro-chlorination and ammonia oxidation reactions were about to stop. Therefore, salt addition was necessary for EC nitrogen removal from the landfill leachate. 5.3.2 EC denitrification during the continuous flow process CFSTR was also tested for EC nitrogen removal from landfill leachate. As illustrated in figure 5.7, the influent had a nitrogen concentration of 3500 mg/L. The chloride concentration was about 3500 mg/L (fresh leachate sample while different from batch tests). With the pretreatment retention time of 2 hrs, the nitrogen concentration decreased to 1800 mg/L. A longer duration of treatment in hydraulic retention time (HRT) would achieve a higher level of nitrogen removal. Five hours of

4000

4000

3000

3000

2000

2000

Influent NH4+-N Effluent NH4+-N


1000

Influent ClEffluent Cl-

1000

0 1 2 3 4 5 6

Retention time (h)

Figure 5.7 TIN removal and chloride concentration as a function of the contact time (HRT) in the CFSTR for EC denitrification.

71

Cl- (mg/L)

N (mg/L)

Chapter 5: Nitrogen removal from wastewater by electrochemical denitrification HRT by electrolysis in the CFSTR achieved complete denitrification, while a chloride concentration at about 2500 mg/L could be maintained in the EC reactor. Both batch and continuous CFSTR operations could be used for EC denitrification of landfill leachate. As anticipated, the TIN removal percentage increased consistently with the batch treatment time or HRT of electrolysis in the CFSTR (Figure 5.8, 5.9). However, the current efficiency for nitrogen removal showed a trend of decrease with the time. In comparison, the CFSTR mode appeared to perform better than the batch treatment. For the CFSTR, the optimal HRT could be defined as 4 hrs. With this HRT, the TIN removal could reach 95%, while the current efficiency was around 50% (Figure 5.9). For the batch reactor, however, it was difficult to decide the optimum treatment duration. A treatment of 4 hrs had only 80% nitrogen removal, while the current efficiency was already lower than 40% (Figure 5.8). In addition to ammonia electrolysis, the EC process was highly effective for discoloring of the leachate. The original dark brown liquor was changed gradually into a light yellow water by electrolysis (Figure 5.11). This was likely brought about by electro-chlorination and related organic oxidation reactions. EC reaction would cause destruction of large and colored organic molecules. The decoloring function should be favorable to the subsequent biological treatment of landfill leachate.

100 90 80

100 90 80 70 60 50 40 30 20

Current efficenency (%)

70 60 50 40 30 20 10 0 1 2 3 4 5 6

TIN degradation (%)

Current efficency TIN degradation

10 0

Time (hr)

Figure 5.8 Current efficiency and TIN degradation as the functions of reaction time in batch operation.

72

Chapter 5: Nitrogen removal from wastewater by electrochemical denitrification 5.3.3 Power consumption and current efficiency The amount of nitrogen removal from the landfill leachate depended on the power input into the EC treatment system. The current density and the gap distance of the electrodes also affected the power efficiency for denitrification. The power consumption rate was rather high when the bulk solution had a high nitrogen concentration. The power consumption for nitrogen removal increased nearly linearly with an ever decreasing nitrogen concentration during the EC process (Figure 5.10). The power consumption rate for nitrogen removal from the landfill leachate was apparently higher than that from the sludge centrate liquor (about 23 kWh/kg N). However, for both batch operation and CFSTR, the power consumption rates were similar during EC denitrification.

5.4 Summary

The laboratory experimental results indicate that the EC process is highly effective for denitrification of the saline sludge liquor with a NH4+-N of around 500 mg/L. Complete nitrogen removal could be achieved within a contact time of 1 hr or so. The best current efficiency for nitrogen removal was obtained with a gap distance between the electrodes at 8 mm, followed by the gap of 15 mm and then 3 mm. Electro-chlorination was considered to be the major mechanism for EC denitrification. Nonetheless, there was not significant accumulation of chloramines during the EC reactions. The power consumption rate for nitrogen removal of the sludge liquor was around 23 kWh/kg N. Additional electro-flocculation with a pair of iron needle electrodes could enhance the flocculation and subsequent sedimentation of colloidal organics in the sludge liquor, which increased the organic removal from less than 30% to more than 70%. The EC system also performed well in denitrification of landfill leachate. With a HRT of 5 hrs in CFSTR, complete TIN removal of more 3000 mg/L of ammonia nitrogen could be achieved. The EC process was highly effective in decoloring the landfill leachate, which would lead to easy downstream treatment. In general, the EC process has the potential to be developed as a cost-effective alternative for nitrogen removal and organic degradation in high strength wastewater treatment, such as for separate treatment of saline sludge liquor and landfill leachate. 73

Chapter 5: Nitrogen removal from wastewater by electrochemical denitrification

100 90 80

100 90 80 70 60 50 40 30 20

Current efficency (%)

70 60 50 40 30 20 10 0 1 2 3 4 5 6

TIN degradation (%)

Current efficiency TIN degradation

10 0

Time (hr)

Figure 5.9 Current efficiency and TIN degradation as the functions of reaction time in continuous.

4000

3000 Continnous Semi-batch

N (mg/L)

2000

1000

0 0 20 40 60 80

Power consumption (kWh/kg-N)

Figure 5.11 TIN as a function of the power consumption.

74

Chapter 6: Electrochemical degradation kinetics of phenol and its intermediate products

Chapter 6

Electrochemical degradation kinetics of phenol and its intermediate products


6.1 Performance of selected anodes in EC phenol degradation

6.1.1 Phenol degradation and TOC removal The four different anodes, Ti/SnO2-Sb-Al, Ti/RuO2, Ti/IrO2 and Pt, showed rather different effectiveness for EC phenol degradation (Figures 6.1 and 6.3). The best treatment result was obtained with the Ti/SnO2-Sb-Al anode, followed by the noble Pt anode and then the conventional DSA of Ti/RuO2 and Ti/IrO2 anodes. Phenol was destructed nearly completely from the initial 100 mg/L after 3 hrs by the Ti/SnO2-Sb-Al anode and 5 hrs by the Pt anode (Figures 6.1 and 6.3). At the same current input rate to the Ti/RuO2 anode, only 42% phenol removal was achieved after 6 hrs of electrolysis. From an initial concentration of 500 mg/L, phenol was degraded almost completely after 10 hrs by the Ti/SnO2-Sb-Al electrode, in comparison to less than 50% by the Ti/IrO2 electrode (Figure 6.1). The four electrodes had an even greater difference in organic mineralization measured by the TOC reduction (Figures 6.2 and 6.4). Complete organic removal was achieved after 5 hrs on the Ti/SnO2-Sb-Al anode. However, under the same electrolysis conditions at a current density of 20 mA/cm2, the Pt and Ti/RuO2 anodes reached just 36% and 12% TOC removals, respectively. For the phenol solution with an initial concentration of 500 mg/L, almost 95 % of TOC was removed after 12 hrs with the Ti/SnO2-Sb-Al anode. However, the TOC removals were only 28-30% for the Ti/IrO2 and Ti/RuO2 electrodes under the same electrolysis conditions at a current density of 20 mA/cm2.

75

Chapter 6: Electrochemical degradation kinetics of phenol and its intermediate products

500 Ti/SnO2-Sb-Al Ti/IrO2 400 Ti/RuO Figure 6.1 500mg/L phenol incineration on 3 kinds of2 anode

12

10

Phenol concentration (mg/L)

300

200

100

0 0 2 4 6 8 10 12

Reaction time (hr)

Figure 6.1 Phenol destruction on the three types of anodes

350 300 250

14

12

10 Sn Ir Ru

TOC (mg/L)

200 150 100

4 50 2 0 0 2 4 6 8 10 12

Reaction time (hr)

Figure 6.2 TOC and solution pH during EC phenol degradation

76

pH

Voltage (V)

Chapter 6: Electrochemical degradation kinetics of phenol and its intermediate products The Ti/SnO2-Sb-Al cell had a voltage input of about 4.2 V, which was somewhat higher than those of the Pt and Ti/RuO2 cells (Figure 6.1). Pure SnO2 is a kind of typical n-type semiconductor. Although its conductivity could be greatly increased by doping of Sb, the conductivity of the coating film was apparently incomparable to pure metals such as Ti and Pt. However, because of a greater efficiency phenol and TOC reduction on the Ti/SnO2-Sb-Al anode than that on the other three anodes, the Ti/SnO2-Sb-Al anode required a much less charge consumption to achieve complete phenol and TOC removals. Compared to these anodes, the Ti/SnO2-Sb-Al anode is considered to have a catalytic capacity toward the destruction of aromatic compounds and degradation of organic pollutants (Kotz et al., 1991; Comninellis 1994; Grimm et al., 2000). 6.1.2 The solution pH during the EC process The residual TOC in the solutions after EC phenol destruction suggests the formation and accumulation of organic intermediate products. TOC reduction was much slower than phenol destruction particularly on the Pt and Ti/RuO2 anodes. The solution pH changed probably with the intermediate production during the electrolysis process (Figure 6.2). In the Ti/SnO2-Sb-Al system, the pH dropped initially from 5.5 to around 4.0 while the phenol was being oxidized. Thereafter, the pH increased gradually to 5.4 as TOC mineralization approached its completion. There was a pronounced drop of pH in the first 2 hours of electrolysis, while about 55 % of phenol disappeared (Figures 6.1, 6.2). This reached the lowest pH point which followed by an increase. A slightly increase trend appeared during next 10 hrs. On the contrary, in the Pt and Ti/RuO2 systems during 6 hrs of electrolysis (figure 6.4), the pH decreased continuously to around 4.0 with no sign of increase, while a great TOC content remained in the solution. The drop in pH was apparently caused by the formation of acidic substances from the phenol destruction. Aliphatic acids were believed to be one of the main intermediates produced by phenol electrolysis (Comninellis and Pulgarin, 1991; Comninellis, 1994; Houk et al., 1998; Feng and Li, 2003; Li et al., 2005). Continuous pH decrease in the solutions electrolyzed by the electrodes Ti/IrO2 and Ti/RuO2 (figure 6.2, 6.4) indicated the slow degradation and hence accumulation of intermediate organic acids in the systems.

77

Chapter 6: Electrochemical degradation kinetics of phenol and its intermediate products


14 100 Sn Pt Ru

12

Phenol concentration (mg/L)

80 10 60 8 40 6 20

0 0 1 2 3 4 5 6

Reaction time (hr)

Figure 6.3 Phenol removals by the three types of anodes

14

80

12

TOC (mg/L)

60

10 Sn Pt Ru

40 6 20

0 0 1 2 3 4 5 6

Reaction time (hr)

Figure 6.4 TOC and solution pH vs. time during phenol degradation

78

pH

Voltage (V)

Chapter 6: Electrochemical degradation kinetics of phenol and its intermediate products 6.1.3 Current efficiency and TOC removal efficiency Electrode Ti/SnO2-Sb-Al had an average current efficiency of around 50% for phenol oxidation from 500 mg/L at a current input of 20 mA/cm2. The value of the current efficiency for phenol destruction decreased as the phenol concentration became lower (Figure 6.5). On the electrodes of Pt and Ti/RuO2, much lower current efficiencies of about 7% were recorded. With an initial phenol concentration of 100 mg/L, the current efficiency of electrode Ti/SnO2-Sb-Al anode was significantly greater than the Pt and Ti/RuO2 anodes for TOC removal (Figure 6.6). Electrode Ti/SnO2-Sb-Al was two to three times greater than the other electrodes. Current efficiency on TOC destruction during the final 3 hrs of electrolysis faster than the first 3 hrs on electrode Ti/SnO2-Sb-Al, which means the major intermediate compounds in the solution has replaced phenol and the degradation rate was accelerated by the catalytic effect of this anode.

6.2 Intermediate products and reaction routes for different electrodes

6.2.1 Direct products of phenol oxidation The phenol solutions exhibited color changes during the electrolysis processes. In the anode Ti/SnO2-Sb-Al cells, the liquid turned to light yellow first, which began to fade after 3 hrs when the phenol concentration was reduced to zero. For the Pt and Ti/RuO2 cells, however, the solution turned to yellow and became darker with the EC process. The color did not fade even after 6 hrs of electrolysis. It is generally believed that the light yellow color is caused by benzoquinone, which is produced by phenol oxidation (Comninellis and Pulgarin, 1991). Benzoquinone and hydroquinone are known as an active redox couple in equilibrium in an aqueous solution (Rieger, 1987). The HPLC results indicate the transformation of the phenol by electrolysis to benzoquinone, hydroquinone and catechol (Figure 6.7). These aromatic compounds had no detectable accumulation in the solution electrolyzed by the Ti/SnO2-Sb-Al anode for 6 hrs. However, they accumulated to a higher concentration level and for a much longer time in the solutions treated by the Ti/RuO2 and Pt anodes. The total detectable amount of aromatic compounds was about 5 mg/L in the cell of Pt and 20 mg/L in the cell of Ti/RuO2 based on the carbon content. Benzoquinone was found in 79

Chapter 6: Electrochemical degradation kinetics of phenol and its intermediate products


25

Current efficiency on phenol degradation (mg/L C)

20

Ti/SnO2-Sb-Al Pt Ti/RuO2

15

10

0 1 2

Reaction time (hr)

Figure 6.5 Current efficiency on phenol removal

Current efficiency on TOC removal (mg/LC)

Ti/SnO2-Sb-Al Pt Ti/RuO2

0 1 2 3 4 5 6

Reaction time (hr) Figure 6.6 Current efficiency on TOC removal

80

Chapter 6: Electrochemical degradation kinetics of phenol and its intermediate products

50

Phenol incineration on Ti/SnO2-Sb-Al anode

Phenol oncineration on Pt anode

Phenol incineration on Ti/RuO2 anode

40

Concentration-C (mg/L)

30

Hydroquinoe 1,2-benzoquinone 1,4-benzoquinone Catechol

20

10

0 Muconic acid Oxoglutaric acid Maleic acid Fumaric acid Succinic acid

Concentrration-C (mg/L)

0 Pyruvic acid Malonic acid Oxalic acid Acetic acid Formic acid

10

Concentrration-C (mg/L)

0 0 60 120 180 240 300 360 0 60 120 180 240 300 360 0 60 120 180 240 300 360

Retention time (min)

Retention time (min)

Retention time (min)

Figure 6.7 HPLC analysis of the intermediates of EC phenol degradation on the three anodes

81

Chapter 6: Electrochemical degradation kinetics of phenol and its intermediate products

100

80

Phenol Aromatic intermediates 6, 5, 4 - C aliphatic acids 3, 2, 1 - C aliphatic acids

Concentration - C (mg/L)

60

40

20

0 0 60 120 180 240 300 360

Time (min)
Figure 6.8 TOC components of the intermediates of EC phenol degradation by the three anodes Ti/SnO2-Sb-Al (left), Pt (middle), and Ti/RuO2 (right)
B B B B

82

Chapter 6: Electrochemical degradation kinetics of phenol and its intermediate products

Hydroquinone
60

Catechol

Oxoglutaric acid

Concentration (mg/L)

40

Ti/SnO2-Sb-Al 20 Pt Ti/RuO2

50 0

Fumaric acid

Maleic acid

Succinic acid

40

Concentration (mg/L)

30

20

10

50 0

malonic acid

Acetic acid

Oxalic acid

40

Concentration (mg/L)

30

20

10

0 0 60 120 180 240 300 360 0 60 120 180 240 300 360 0 60 120 180 240 300 360

Time (min)

Time (min)

Time (min)

Figure 6.9 Degradation of pure intermediate chemicals on the three anodes

83

Chapter 6: Electrochemical degradation kinetics of phenol and its intermediate products the Pt cells and both hydroquinone and benzoquinone lefted in the Ti/RuO2 cells (Figure 6.7). 6.2.2 Aliphatic acids from phenol oxidation A number of carboxylic acids with 2 to 6 carbon atoms were identified by HPLC as the intermediates of phenol degradation in the electrolyzed solutions. The acids were the intermediate products after aromatic ring cleavage. However, there were considerable differences in the formation and degradation of the intermediate organic acids between the three EC systems of different anodes. For the Ti/SnO2-SbAl cells, oxoglutaric acid and oxalic acid were the most dominant organic acids. All acids were reduced to zero in concentration after 6 hrs, which was in agreement with the TOC measurement. For the Pt cells, oxoglutaric acid, maleic acid and acetic acid accumulated in the solution as aromatic compounds were degraded by electrolysis. For the Ti/RuO2 system, maleic acid, oxalic acid and formic acid gradually increased with the EC process. The total amount of aliphatic acids was 6 mg/L based on the sum of carbon, which was less the concentration of 12 mg/L in the cell of Pt (Figure 6.8). This was in line with a higher concentration of phenol and other aromatic intermediates totally 75 mg/L remaining in the solution electrolyzed by the Ti/RuO2 anode (Figure 6.8). Generally speaking, as indicated by the HPLC analysis of individual intermediates, the overall degradation rate of the intermediate organics by the Pt and Ti/RuO2 anodes was much slower than that with the Ti/SnO2-Sb-Al anode under the same experimental conditions. 6.2.3 Polymeric compounds produced during the EC process As described above, a yellow to brown color built up in the solutions that were treated by the Ti/RuO2 and Pt anodes. In the Ti/SnO2-Sb-Al system only a light yellow color was observed for a period of time during the EC process. The color disappeared well before the TOC had been completely removed. HPLC results did not suggest the substances that would cause the color appearance. The color was probably caused by the polymeric compounds produced by the EC oxidation of phenol (Gattrell and Kirk, 1993; Rajeshwar et al., 1994, Li et al., 2005). The products of polymerization consist of hydroquinone-benzoquinone monomeric units that 84

Chapter 6: Electrochemical degradation kinetics of phenol and its intermediate products generally exhibit a dark brown color (Feng et al., 1995). These amorphous polymers are formed by a radical reaction mechanism by phenate anions (Mengoli and Musiani, 1986; Rodgers et al., 1999). The color compounds could not be detected by the present HPLC analysis, however, they were apparently included in the TOC measurement (Figure 6.2, 6.4). 6.2.4 Grouping of intermediate organic compounds The organic components analyzed by HPLC were combined into a few groups for comparison (Figure 6.8). TOC1 referred to the sum of four aromatic compounds of hydroquinone, 1,2-benzoquinone, 1,4-benzoquinone and catechol. TOC2 included five long carbon chain acids in muconic acid, oxoglutaric acid, fumaric acid, maleic acid and succinic acid. TOC3 was the sum of five short carbon chain acids of malonic acid, oxalic acid, acetic acid and formaic acid. With the decrease of phenol, there were accumulations of benzoic compounds, organic acids and unknown organics that could be polymeric materials. It was a gradual substitute process that simple organics replaced complex organics, or might be reversed. These intermediates substituted the organic carbons of the parent substrate. The change in the concentrations of different TOC groups showed the organic degradation steps from larger molecules to smaller molecules. Electrolysis of 6 hrs was enough for all intermediate species to be oxidized by the Ti/SnO2-Sb-Al anode. However, the same treatment duration time only could transfer phenol into carboxylic acids on the Pt anode, and even less than 10% of TOC transformation on the Ti/RuO2 anode. This comparison suggested that the acid oxidation would be the rate-limiting step for complete TOC removal by electrolysis with the Pt anode, while both ring cleavage and acid oxidation were difficult for the Ti/RuO2 anode. 6.2.5 EC destruction of the intermediate products of phenol electrolysis The oxidation kinetics of the above intermediates of phenol electrolysis were examined individually with pure chemicals as the parent substrates. Similar to phenol oxidation, the EC destruction process of a pure intermediate was traced by HPLC for solution analysis (Figure 6.9). According to the curve of concentration reduction, the

85

Chapter 6: Electrochemical degradation kinetics of phenol and its intermediate products


Table 6.1 Kinetic rate constants for EC oxidation of different intermediates by the

three types of anodes (min-1) Ti/SnO2-Sb-Al Phenol Hydroquinone Benzoquinone Catechol Oxoglutaric acid Maleic acid Fumaric acid Succinic acid Malonic acid Oxalic acid Acetic acid 0.028 0.0099 0.14 0.002 0.026 0.025 0.033 0.032 0.032 0.026 0.008 0.019 0.028 0.024 0.0023 0.0067 0.0015 0.0004 0.0002 0.000009 0.0049 0.00002 Pt Ti/RuO2 0.0015 0.0046 0.0055 0.0041 0.011 0.001 0.0006 0.000003 0.00004 0.0019 0.000007

first-order rate constant could be estimated for the EC destruction reaction of a specific intermediate compound (Table 6.1). Nearly all of the chemicals were tested and the Ti/SnO2-Sb-Al anode delivered the fastest EC destruction rate, followed by the Pt anode and then the Ti/RuO2. For the EC destruction of aromatic compounds, the Ti/SnO2-Sb-Al anode was about a half to one order of magnitude faster than the Pt anode and around one to two orders of magnitude faster than the Ti/RuO2 anode. As to the carboxylic acids, the destruction rate constants on the Ti/SnO2-Sb-Al anode were about one to two orders of magnitude higher that those on the Pt anode and around two to four orders of magnitude higher than those on the Ti/RuO2 anode. These results further confirmed the generally high effectiveness of the Ti/SnO2-Sb-Al anode to oxidation of any type of organic compounds. It is apparent that there was no pronounced rate-limiting step in complete phenol degradation by the Ti/SnO2-Sb-Al anode. However, oxidation of the aliphatic acids, rather than destruction of the aromatic compounds or ring cleavage apparently was the rate-limiting step in EC phenol degradation on the Pt anode. Both the destruction of aromatic rings and oxidation of carboxylic acids were rate limiting processes to phenol mineralization on the Ti/RuO2 anode. In addition, on any specific anode, the organic destruction by the EC oxidation was faster for large molecules than for smaller ones (Table 6.1). EC treatment was confirmed to be more effective for

86

Chapter 6: Electrochemical degradation kinetics of phenol and its intermediate products treatment of large organic compounds than smaller organic molecules. Furthermore, in the phenol degradation tests by the Pt and Ti/RuO2 anodes, the slow degradation rate of intermediate acids led to an accumulation of H+ and continuous drop of the solution pH.

6.3 EC phenol degradation pathways on different types of electrodes

6.3.1 Major phenol electrolysis reactions The pathway of phenol degradation by EC oxidation has been the topic of many investigations (Comninellis and Pulgarin, 1991; Gattrell and Kirk, 1993; Rodgers et al., 1999; Feng and Li, 2003, Li et al., 2005). It is generally considered that phenol oxidation begins with an electron transfer that leads to phenoxy radical reactions (Lund and Baizer, 1991). Radical reactions result in the formation of benzoquinone and hydroquinone, which can be degraded with ring breakage to various carboxylic acids. (Comninellis and Pulgarin, 1991; Tahar and Savall, 1999). Once the ring cleavage takes place, the subsequent degradation will quickly convert the aliphatic compounds through ethers and simple alcohols. The oxidation proceeds further to break down the C = C bond and decarboxylation to form smaller acids, and eventually mineralize to acetic acid, formic acid and carbon dioxide (Devlin and Harris, 1984; Houk et al., 1998). Through a series of reactions an organic acid will be oxidized completely with continuous electrolysis. For instance, maleic acid would first be oxidized to oxalic acid, or be reduced to succinic acid at the cathode followed by oxidation at the anode to malonic acid and then to acetic acid, and finally to CO2 (Houk et al., 1998; Johnson et al., 1999; Feng and Li, 2003; Li et al., 2005). 6.3.2 Aromatic ring cleavage vs. polymerization The present experimental findings in the removal of phenol and TOC, and in the formation of intermediates, are in general agreement with the reaction pathway of EC phenol degradation outlined above. In addition, the pathway should include the polymer formation from aromatic compounds via phenoxy radicals that attack either benzoquinone or hydroquinone. Polymerization is hardly a reversible reaction, and the polymers are more recalcitrant to EC degradation than phenol. For benzoquinone and 87

Chapter 6: Electrochemical degradation kinetics of phenol and its intermediate products hydroquinone, ring cleavage and polymerization are two parallel processes that occur simultaneously. If the aromatic ring cleavage takes place rapidly, then there is little accumulation of benzoquinone and insignificant formation of polymeric compounds, which is the case with the Ti/SnO2-Sb-Al anode. In contrast, if the ring breaks rather slowly, then polymerization becomes more significant, which is in fact the case with the Ti/RuO2 and Pt anodes. Comninellis (1994) has detected the accumulation of OH radicals on the surface of anode SnO2 and no OH appearance on anode RuO2. The existence of OH radicals definitely strengthens the oxidation capability of the Ti/SnO2-Sb-Al anode for ring cleavage and organic degradation. In such a system, dimerization and polymerization are not the major reactions in comparison to ring oxidation (Chu , 1995). For electrodes Ti/RuO2 and Ti/SnO2-Sb-Al, the difference in oxidation strength appears to be the key for the different levels of the formation and accumulation of polymer products in the electrolyzed solutions. Hence, compared to the Ti/RuO2 and Pt anodes, the Ti/SnO2-Sb-Al anode imposed a catalytic capability for organic oxidation. 6.3.3 Pathways of EC phenol degradation The present HPLC analysis revealed more complete intermediate products and their abundances during the EC phenol degradation process. The results suggest that phenol oxidation might not follow a simple series of reactions. In particular, the ring cleavage could form aliphatic acids of different molecular weights with one to six carbons. Small organic acids were already detected in the early phase of electrolysis. Hence, small acid could be produced directly from either ring cleavage of aromatic compounds or from destruction of larger organic acids. According to the electrolysis performance and intermediate abundances during the process, the EC phenol degradation pathways were proposed for the three electrodes (Figures 6.10, 6.11 and 6.12). The electrodes shared largely a similar reaction pathway. However, there might be some variations in the routes of the pathways. Such as the degradation of oxoglutaric acid, oxalic acid was the intermediate on Ti/SnO2-Sb-Al anode, succinic acid and oxalic acid composed the intermediates in the cell of Pt, while succinic acid and formic acid formed the intermediates of electrode Ti/RuO2. Different electrodes with different surface properties performed rather differently in the electrolysis of phenol and its intermediates. The surface characteristics of an anode could to a certain 88

Chapter 6: Electrochemical degradation kinetics of phenol and its intermediate products extent regulate the predominant routes of phenol degradation as well as to determine the reaction kinetics. Nonetheless, it is difficult to identify a single chemical reaction as a rate-limiting step for EC phenol degradation at any electrodes.

6.4 Summary

Phenol could be readily mineralized by electrolysis at the Ti/SnO2-Sb-Al anode. However, phenol degradation was considerably slower at the Ti/RuO2, Ti/IrO2 and Pt anodes. The intermediate products of EC phenol degradation, including benzoquinone and organic acids, were subsequently oxidized rapidly by the Ti/SnO2Sb-Al anode, but accumulated in the cells of Ti/RuO2, Ti/IrO2 and Pt. The Ti/SnO2Sb-Al anode had a high current efficiency for phenol destruction and TOC removal, which was 2-3 times higher than the current efficiencies of Ti/RuO2, Ti/IrO2 and Pt anodes. The intermediates according to HPLC analysis included aromatic compounds of hydroquinone, 1,2-benzoquinone, 1,4-benzoquinone and catechol and aliphatic acids of muconic acid, oxoglutaric acid, fumaric acid, maleic acid, succinic acid, malonic acid, pyruvic acid, oxalic acid, actic acid and formic acid. In addition, there were apparently dark-colored polymeric products formed at the Ti/RuO2, Ti/IrO2 and Pt anodes, which decreased the EC organic degradation effectiveness. EC tests with pure intermediate chemicals also demonstrated that different electrodes performed rather differently in organic destruction. Based on the electrochemical performance and the abundances of the intermediate chemicals during the EC process, the modified reaction pathways of phenol degradation on the different electrodes were proposed. It is suggested that ring cleavage was critical to phenol oxidation. Ring breakage resulted in the formation of intermediate organic acids with different molecular weights. There could be some variations in the routes of the pathways of phenol degradation by different anodes. The surface characteristic of the anode could to a certain extent regulate the predominant routes of phenol oxidation as well as to determine the reaction kinetics.

89

Chapter 6: Electrochemical degradation kinetics of phenol and its intermediate products


OH [O] OH [O] OH O [O] O

Phenol Hydro quinone


OH OH O [O] [O] OH OH O

Benzoquinone

Polymeric products

Catechol I
COOH COOH

Benzoquinone

Muconic acid II
COOH COOH O 2 COOH COOH CO 2

Oxalic acid Oxoglutaric acid


III COOH COOH

COOH

HOOC

Maleic acid

Fumaric acid
COOH 2 COOH CO 2 H 2O

Oxalic acid
COOH COOH C O2 COOH COOH

Succinic acid

Malonic acid
COOH CO 2 COOH CO 2 H 2O

Figure 6.12 Proposed phenol degradation pathway at anode Ti/SnO2-Sb-Al.

90

Chapter 6: Electrochemical degradation kinetics of phenol and its intermediate products


OH [O] OH [O] OH O [O] O

Phenol Hydroquinone
OH OH O [O] [O] OH OH O

Benzoquinone

Polymeric products

Catechol I
COOH COOH

Benzoquinone

Muconic acid II
COOH COOH O COOH HCOOH COOH

Succinic acid
2

COOH CO2 COOH H2 O

Oxoglutaric acid

COOH 2 COOH

Oxalic acid
HCOOH CO2 H2O

Oxalic acid
COOH III COOH COOH COOH 2 COOH COOH CO2 H2O

Maleic acid

Succinic acid
COOH

Oxalic acid
H2O

2HCOOH COOH

CO2

Oxalic acid
COOH COOH 2HCOOH HOOC COOH CO2 H2 O

Fumaric acid
COOH

Oxalic acid
COOH 2 CO2 COOH H2 O

COOH

Succinic acid

Oxalic acid

Figure 6.13 Proposed phenol degradation pathway at anode Pt.

91

Chapter 6: Electrochemical degradation kinetics of phenol and its intermediate products

OH [O]

OH [O]

OH

O [O]

Phenol Hydroquinone
OH OH O [O] [O] OH OH O

Benzoquinone

Polymeric products

Catechol I
COOH COOH

Benzoquinone

Muconic acid II
COOH COOH O COOH HCOOH COOH

Oxoglutaric acid
III COOH

Succinic acid
COOH COOH

COOH

HOOC

COOH

Maleic acid

Fumaric acid

Succinic acid

COOH 2HCOOH COOH CO 2 H 2O

Figure 6.14 Proposed phenol degradation at anode Ti/RnO2.

92

Chapter 7: Surface and electrochemical properties of the electrodes and EC phenol degradation mechanisms

Chapter 7

Surface and electrochemical properties of the electrodes and EC phenol degradation mechanisms
7.1 Characterization of the electrodes by linear sweep voltammetry (LSV)

7.1.1 LSV and oxidation potentials The electrochemical properties of the four kinds of electrodes were analyzed by the voltammetric technique. The measurements were conducted in 0.25 M Na2SO4 with and without phenol addition at different concentrations. The LSV was performed

Figure 7.1 Phenol oxidation on the three types of anodes.

93

Chapter 7: Surface and electrochemical properties of the electrodes and EC phenol degradation mechanisms at a scan rate of 50 mv/s and the potential ranged from 0.2 to 2.1 V. The LSV results would suggest the potentials of phenol oxidation at the surface of different anodes. 7.1.1.1 Two electrochemical reactions: phenol oxidation and oxygen evolution There were two electrochemical reactions expected at an anode during the LSV scanning process, phenol oxidation and oxygen evolution from water electrolysis. Occurrence of a reaction at a certain voltage would result in a sudden current density increase as a response which was caused by the electron transfer. For example, as observed on electrode Ti/SnO2-Sb-Al, the sudden LSV increases in the current corresponded to the two electrochemical reactions at the voltage of 0.9 V and 1.8 V respectively (Figure 7.1). The peak of phenol oxidation potential was less positive than the oxygen evolution potential of water electrolysis. Similar phenomena could be observed at the Pt electrode, on which the current density 1.1 mA/cm2. However, only water oxidation peaks at 1.1 V potential values could be found on electrode Ti/RuO2. Oxygen evolution potential was advanced than the phenol oxidation potential, which would hinder the phenol oxidation peak.

Figure 7.2 Phenol oxidation at electrode Ti/SnO2-Sb-Al with different phenol concentrations.

94

Chapter 7: Surface and electrochemical properties of the electrodes and EC phenol degradation mechanisms 7.1.1.2 LSV of the electrodes in the electrolyte without phenol The LSV curves in the pure electrolyte show that the Pt and Ti/RuO2 anodes had a lower oxygen evolution potential, with respective values of around 1.3 and 1.1 V vs. the reference of the saturated Ag/AgCl electrode (Figures 7.3 and 7.4). In comparison, Ti/SnO2-Sb-Al anode had a considerably higher oxygen evolution potential of 1.8 V vs. the reference (Figure 7.2). It is believed that oxygen molecules are formed by radical reactions involving the OH free radical generated electrochemically (Polcaro et al, 2000; Tanaka et al. 2002). The high overpotential of oxygen at the Ti/SnO2-Sb-Al anode suggests that the reaction forming molecular oxygen was probably restrained. This would facilitate the generation and accumulation of OH radical at the electrode surface. In contrast, oxygen evolution took place rapidly at a lower potential on the Pt and Ti/RuO2 anodes, which was unfavorable to the accumulation of OH radicals. 7.1.2 LSV curves in the solutions with phenol addition 7.1.2.1 Anode Ti/SnO2-Sb-Al Addition of phenol into the electrolyte resulted in new shoulders or peaks along the LSV curves of the Ti/SnO2-Sb-Al anode (Figure 7.2). The peak current increased significantly from 0.2 mA/cm2 to 3.1 mA/cm2 with the increase in phenol concentration from 1000 to 10,000 mg/L. The peaks were located at about 1.1- 1.5 V for different concentrations. While the impute potential ranged below 1.6 V for oxygen evolution was not expected at the Ti/SnO2-Sb-Al anode. Hence these peaks suggest direct anodic oxidation of phenol during the LSV tests. The Ti/SnO2-Sb-Al anode apparently was able to oxidize phenol directly on the electrode surface. A phenol concentration of 10,000 mg/L accelerated the oxidation rate comparing with the concentration of 1000 mg/L. 7.1.2.2 Anode Pt

95

Chapter 7: Surface and electrochemical properties of the electrodes and EC phenol degradation mechanisms

Figure 7.3 Phenol oxidation at electrode Pt for different concentrations.

Figure 7.4 Phenol oxidation at electrode Ti/RuO2 for different concentrations.

96

Chapter 7: Surface and electrochemical properties of the electrodes and EC phenol degradation mechanisms For the Pt anode, current peaks were observed in the LSV curves with phenol addition and increased with phenol concentration. Phenol also could be oxidized directly at the Pt surface, however, the oxidation rate was not as faster as that at the Ti/SnO2-Sb-Al anode. The corresponding phenol oxidation potential at Pt was somewhat lower than the potential on the Ti/SnO2-Sb-Al anode. The response of the limiting current might be related to the mass transfer of the organic from bulk solution to the anode surface. Increasing the phenol concentration to 10,000 mg/L led to a raise of the current density up to 1.2 mA/cm2. In comparison, an increase in phenol concentration introduced more electrons into EC reaction on the Ti/SnO2-Sb-Al anode surface which reached a current density of 3.2 mA/cm2. Hence, Ti/SnO2-Sb-Al anode had a greater reactivity towards direct phenol oxidation on its surface. 7.1.2.3 Anode Ti/RuO2 With the Ti/RuO2 anode, however, phenol addition into the electrolyte did not cause significant changes in the LSV measurement. There were no peaks or shoulders found in the LSV curves with the phenol concentration increasing. The result suggests that the possible phenol oxidation potential at the Ti/RuO2 anode was higher than the potential of water oxidation at 1.1 V (Arikawa et al., 1998; Simond and Comninellis, 1997). Hence, water discharge, oxygen evolution was the dominant EC reaction and only an insignificant fraction of charge impute would be utilized for phenol oxidation, which hindered direct phenol oxidation on the Ti/RuO2 anode surface. In other words, direct phenol oxidation at the Ti/RuO2 anode was much more difficult than that on the Ti/SnO2-Sb-Al anode and the Pt anode as well. On the other hand, the starting background current density in the LSV curve at the Ti/RuO2 anode was about 0.1 V higher than the other two anode types. This shows a better conductivity to the Ti/RuO2 anode. Ti/RuO2 anode apparently have more active sites on the surface which favors current transfer, though its reactivity for direct phenol oxidation was not as good as its conductivity. 7.1.2.4 Effect of phenol concentration on the LSV curves In general, concentration affects the mass transfer process from the bulk solution to the anode surface, and additionally, organic concentration also could shift 97

Chapter 7: Surface and electrochemical properties of the electrodes and EC phenol degradation mechanisms

the oxidation potential of the substance compound. Hence in the voltammetric sweep, an increase in phenol concentration may advance the onset phenol oxidation potential. The comparison among voltammetric sweep figures of electrodes Ti/SnO2-Sb-Al, Pt and Ti/RuO2 illustrated their different electrochemical properties, and also exhibited their different catalytic capability toward phenol oxidation (Figures 7.2, 7.3, and 7.4), while the concentration effects during electrolysis should not be overlooked. To the different phases of electrochemical phenol oxidation, electron transfer and mass transfer refered to the two alternatives limiting steps. According to energy requirement, the oxidation reaction carried out by electrodes Ti/SnO2-Sb-Al and Pt were easier than that by Ti/RuO2. Due to the catalytic effect of anode Ti/SnO2-Sb-Al, the energy hump (G) for phenol oxidation reaction was lowered. Voltage impute at 0.8 V may cause phenol oxidation, which 1.8 V is enough to led water discharge and oxygen evolution. 7.1.3 Effect of the solution pH on the phenol oxidation over-potential

98

Chapter 7: Surface and electrochemical properties of the electrodes and EC phenol degradation mechanisms

The dependence of oxidation potential on pH was demonstrated by the cyclic voltammograms at electrodes Ti/SnO2-Sb-Al, Pt and Ti/RuO2 (Figure 7.5, 7.6, and 7.7). With decreasing of pH, there was a slight shift towards the lower potential of the onset of oxidation reactions. To the electrode Ti/SnO2-Sb-Al, the potential got 0.1 V increasing with the solution pH dropping. The oxidation of phenol was accompanied with the loss of protons. A pH change could to a certain extent influence the thermodynamics of phenol oxidation reactions, as suggested by the LSV results, and have maximum 0.1 ma/cm2 increasing of limiting current.

7.2 Electrochemical phenol oxidation mechanisms

The EC organic degradation process includes mass transfer steps and electrochemical reaction steps (Za, 2002), and the process may be simplified to three steps (Figure 7.8): (1) Mass transfer, from bulk to electrode surface, which includes two reactions, and . (2) Electrochemical reaction, . (3) Mass transfer steps shown as and .

99

Chapter 7: Surface and electrochemical properties of the electrodes and EC phenol degradation mechanisms 7.2.1 EC phenol oxidation process and mass transfer Reaction is the mass transfer to the surface of an electrode, the adsorption of the organic matter onto the electrode surface. This is a reaction accompanied by the increase in chemical concentration at the anode surface and the decrease of its concentration in the bulk solution, which also is affected by temperature, the chemical concentration in the solution and the surface hydrophilicity of the anode. The adsorption behavior of a chemical solute could be related to the adsorption free energy of water molecules on the surface with no respect to specific chemicals (Za, 2002). A good hydrophilic surface favors chemical adsorption and increase the amount of substances involved in EC reaction. This is expected to be the case for phenol oxidation on the Ti/SnO2-Sb-Al anode. Reaction is the generation and accumulation of hydroxyl radicals at the electrode surface, and the two EC reactions for OH radical generation (Polcaro et al., 2000; Tanka et al., 2002; Chen, 2004). The rate of OH radical generation and the amount of OH accumulation are determined by the surface property of the electrode.

100

Chapter 7: Surface and electrochemical properties of the electrodes and EC phenol degradation mechanisms The electrode with a higher overpotential of oxygen evolution would lead to more OH radical production from water electrolysis. Meanwhile, a hydrophilic polar surface would allow continuously accumulation of OH radical production on the surface.

R Ti MOx O2 RO
Figure 7.8 EC phenol oxidation process.

H2O / OH-

Ti Titanium plate MOx Metal oxide coating film R Organic matters RO Oxidation product of organic R

7.2.2 EC phenol degradation steps Reaction is the electrochemical oxidation of the organic on the anode surface. Free radical chain reactions would result in a series of organic degradation reactions as demonstrated in Chapter 6. The anode surface property also affects the route of reactions and the reaction kinetics as illustrated in Chapter 6 for different anodes. On the electrode Ti/SnO2-Sb-Al, during reaction , hydroxyl radicals are adsorbed by the metal oxides on the electrode surface (Simond et al., 1997; CorreaLozano et al., 1997; Tanaka et al., 2002). In doped coating, partially substitutes the place of Sn4+ with Sb5+, and the conductivity of SnO2 coating increased due to the presence of oxygen vacancies and free electrons. Additionally, to the cations Sn4+ and Sb5+, the d10 electronic configuration plays a prominent role in allowing a wide band gap to be stabilized. This also favors a more covalent character of the metal-oxygen bond than for other dn cations (Rodrigues and Olivi, 2003; Choisenet et al., 2004). On the other hand, this decreases the need of oxygen atom in the metal oxide crystal lattice, the EC oxidation reactions are simple and direct as follows:
MOx + H 2 O MOx (OH ) + H + + e MOx (OH ) + R MOx + RO + 1 / 2O2 + H + + e

(7.1) (7.2)

where MOx refers to metal oxide, R is the organic matter, and RO is the oxide of the organic matter. 101

Chapter 7: Surface and electrochemical properties of the electrodes and EC phenol degradation mechanisms Oxidation reactions on electrode Ti/RuO2, during reaction are illustrated in following equations (7.3, 7.4 and 7.5). The selectivity of dn electronic configuration of metal oxide made phenol oxidation on Ti/RuO2 anode more complex (Rodrigues and Olivi, 2003; Choisenet et al., 2004).
MOx + H 2 O MOx (OH ) + H + + e MOx (OH ) MOx +1 + H + + e MO x +1 + R MO x + RO + 1 / 2O2

(7.3) (7.4) (7.5)

where MOx+1 refers to a higher valent metal oxide. 7.2.3 Oxygen evolution Reaction and are another mass transfer step, the desorption and releasing of electrolyzed products, including the intermediates and end products. Organic oxidation products that are difficult to be desorpted cause the contamination of Pt. Additional step is the oxygen evolution, as well as the generation of other byproducts. There were considerable differences between anodes Ti/Sn-Sb-Al, Pt and Ti/RuO2 at each step during the phenol degradation process.

7.3 Summary

Electrochemical oxidation with the catalytic electrode Ti/SnO2-Sb-Al was demonstrated as an effective alternative to degradation of toxic and refractory organic pollutants. According to the analysis of electrode electrochemical behaviors, anode Ti/SnO2-Sb-Al had an obvious catalytic capability toward phenol oxidation. The three types of electrodes performed rather differently during LSV tests. The surface properties of the electrodes appeared to affect the reactions involving in EC organic degradation. In comparison to the conventional Pt and Ti/RuO2 anodes, there could be two main reasons for the high effectiveness of the Ti/SnO2-Sb-Al anode in phenol. First, its high overpotential of oxygen evolution would allow more production and accumulation of hydroxyl radicals near the anode surface under a higher potential condition. Organic matter could be readily destructed by the high-energy, nonselective oxidant of free radicals. In contrast, oxygen evolution took place at a rapid rate at a lower potential on the Ti/RuO2 anodes, which was unfavorable to the 102

Chapter 7: Surface and electrochemical properties of the electrodes and EC phenol degradation mechanisms generation of hydroxyl radicals. Second, the peaks in the LSV curves of the Ti/SnO2Sb-Al anode in phenol solution indicates direct organic oxidation at the anode surface. Thus, power could be utilized for direct oxidation of chemical compounds on the anode, which was much difficult to be achieved with the Pt and Ti/RuO2 anodes.

103

Chapter 8: Electrochemical degradation of chlorophenol: Kinetics and pathways

Chapter 8

Electrochemical degradation of chlorophenol: Kinetics and pathways


8.1 EC chlorophenol oxidation

Mono-chlorophenol-2 (MCP) was selected as another model organic compound for treatment by the electrochemical method. Chlorophenol is much more toxic than phenol. The experimental procedures were similar to the phenol degradation study. Three types of electrodes, Ti/SnO2-Sb-Al, Ti/RuO2 and Pt, were examined for their performance and effectiveness in chlorophenol electrolysis. The degradation pathways were proposed based on the HPLC analysis of the intermediate products for different electrodes. The electrochemical properties of the electrodes were

12 100 10

MCP concentration (mg/L)

80

60

Pt Ti/RuO2 6

40

20

0 0 1 2 3 4 5 6

Reaction tiem (hr)

Figure 8.1 MCP destruction by the three types of anodes.

104

Voltage (V)

Ti/SnO2-Sb-Al

Chapter 8: Electrochemical degradation of chlorophenol: Kinetics and pathways characterized, and the results were compared with what obtained from the phenol electrolysis study. 8.1.1 Chlorophenol oxidation and TOC removal by electrolysis The three different anodes, Ti/SnO2-Sb-Al, Ti/RuO2 and Pt, showed rather different effectiveness for EC chlorophenol degradation (Figure 8.1). The best treatment result was obtained with the Ti/SnO2-Sb-Al anode, followed by the noble Pt anode and then the conventional Ti/RuO2 anode. Chlorophenol was destructed nearly completely from an initial concentration of 100 mg/L after 4 hrs by the Ti/SnO2-SbAl anode. In comparison under the same operation conditions, 55% chlorophenol was oxidized by the Pt anode and only 10% chlorophenol was destructed by the Ti/RuO2 anode (Figure 8.1). To the Ti/SnO2-Sb-Al anode, the rate of chlorophenol destruction was comparable to that of phenol destruction. The reduction rate in chlorophenol concentration was particularly fast in the first half hour. However, on the Pt and Ti/RuO2 anodes, chlorophenol destruction took place at an even slower rate than that observed previously for phenol destruction. The three electrodes had an greater difference in organic mineralization as
50 14

40

12

10

TOC (mg/L)

30 Pt Ti/RuO2 8

20

6 10

0 0 1 2 3 4 5 6

Reaction time (hr)

Figure 8.2 TOC removal at the three types of anodes.

105

pH

Ti/SnO2-Sb-Al

Chapter 8: Electrochemical degradation of chlorophenol: Kinetics and pathways indicated by the TOC removal (Figure 8.2). Complete organic oxidation was achieved after 5 hrs at the Ti/SnO2-Sb-Al anode. However, under the same electrolysis conditions at a current density of 20 mA/cm2, just 30% and 10% TOC removals were deliverded by the Pt and Ti/RuO2 anodes, respectively. Electrode Ti/SnO2-Sb-Al appeared to be highly effective for complete degradation of chlorophenol. On the other hand, electrodes Pt and Ti/RuO2 appeared to be worsein performance for treating chlorophenol than phenol degradation tests. 8.1.2 The solution pH during electrolysis The residual TOC in the solutions after EC chlorophenol destruction suggests the formation and accumulation of organic intermediate products. TOC reduction was worse than chlorophenol destruction, particularly at the Pt and Ti/RuO2 anodes. The solution pH changed correspondingly with the components of intermediates at the electrolysis process (Figure 8.2). In the Ti/SnO2-Sb-Al system, the pH dropped initially from 5.5 to around 3.6 while the chlorophenol was being oxidized. Thereafter, the pH increased gradually to 5.3 as TOC mineralization approached completion. On the contrary, in the Pt and Ti/RuO2 cells during 6 hrs of electrolysis, the pH decreased continuously to around 4.2 with no sign of increase, while a high TOC content remained in the solutions. The drop in pH was apparently caused by the formation of aliphatic acidic substances from the chlorophenol destruction. Similar to phenol electrolysis, aliphatic acids were believed to be the main intermediates produced by EC degradation of chlorophenol (Comninellis and Pulgarin, 1991; Comninellis, 1994; Houk et al., 1998; Feng and Li, 2003; Li et al., 2005). 8.1.3 Current efficiency The current efficiency indicates both the performance of the EC process and the specificity of the surface reactions involved. For a current input at 20 mA/cm2, the Ti/SnO2-Sb-Al anode had a current efficiency of more than 50% for TOC destruction. The efficiency decreased to around 10% by the end of the EC process with an everdecreasing chlorophenol concentration in the electrolyzed solution. A much lower current efficiency of chlorophenol treatment was recorded for both Pt and Ti/RuO2 electrodes under the same EC conditions (Figure 8.4). The Pt anode had a current 106

Chapter 8: Electrochemical degradation of chlorophenol: Kinetics and pathways efficiency of around 8% throughout the test period, while the Ti/RuO2 had an efficiency of only about 5%. For the Ti/SnO2-Sb-Al anode, the current efficiency based on TOC for EC chlorophenol treatment decreased slightly as the current density increased (Figure 8.3). An increase in current density was able to accelerate the reaction rates. However, a higher current input might result in more power contribution to heat generation and oxygen evolution. Different from phenol electrolysis, dechlorination was supposed to one of the important steps in chlorophenol oxidation. As exhibited in Figure 8.3, continuous chloride release of from the chemical substance was identified by the solution analysis in the cell of electrode Ti/SnO2-Sb-Al. After 2 hrs of the electrolysis, when 95% chlorophenol was destructured, the chloride concentration reached the maximum point. Chloride ions in the solution would in turn lead to chloride electrolysis and chlorine evolution, as the overpotential of chlorine evolution is slightly lower than oxygen evolution. On the other hand, chlorine is an effective oxidant. Chlorine production would work on the oxidation of chlorophenol and its intermediates. Therefore, as described previously, chlorophenol degradation on the Ti/SnO2-Sb-Al anode had double engine forces, and definetly faster than phenol degradation which
40 20 mA/cm2 40 mA/cm2 60 mA/cm2 Cl- 20 mA/cm2 Cl- 40 mA/cm2 Cl- 60 mA/cm2 30

25

30

Current efficiency (%)

20 Cl (mg/L)

20

15

10 10 5

0 0 1 2 3 4 5 6

Reaction time (hr)

Figure 8.3 TOC degradation as a function of charge input at different current density

107

Chapter 8: Electrochemical degradation of chlorophenol: Kinetics and pathways was pushed by one engine. On the contrary, chloride release was detected at a much slower rate in the chlorophenol solutions electrolyzed by the Pt and Ti/RuO2 anode. It was apparent that dechlorination was a important limiting step in the electrolysis of chlorophenol by electrodes Pt and Ti/RuO2, which led to a considerable slow rate of chlorophenol destruction and organic degradation.

8.2 Intermediates and pathways of EC chlorophenol degradation

8.2.1 HPLC analysis of intermediate compounds The intermediates of chlorophenol electrolysis were analyzed by HPLC by the solutions treated by different electrodes (Figures 8.5 and 8.6). Eight kinds of aromatic compounds and chlorophenol were identified, and these substances in number was more than what encountered during the phenol degradation tests. The aromatic intermediates included chlorohydroquinone, 4-chlorocatechol, 3-chlorocatechol, phenol, hydroquinone, benzoquinone, 1, 2-benzoquinone, 1, 4-benzoquinone. These eight aromatic compounds summed to 2 mg/L (based on C) in the Ti/SnO2-Sb-Al cells (Figure 8.5) after 6 hrs treatment. However, they accumulated at a higher concentration level and for a much longer time in the solutions treated by the Ti/RuO2 and Pt anodes. Also, about 4 mg/L substances left in the cell of Ti/RuO2 while 3 mg/L in the Pt cell. There was more chlorohydroquinone found in the Pt cells and both hydroquinone and benzoquinone were found in the Ti/RuO2 systems (Figure 8.5). During the EC treatment process with different electrodes, the chlorophenol solutions did not exhibit obvious color changes as observed for phenol electrolysis. This was probably due to the decoloring effect of hypochlorite that was formed by chloride electrolysis following EC dechlorination and chlorophenol oxidation. It is generally considered that phenol is the major intermediate product of chlorophenol oxidation (Comninellis and Pulgarin, 1991). However, the concentrations of chlorohydriquinone, 4-chlorocatechol, 3-chlorocatechol were not less than phenol in considerable quantity in the electrolyzed solutions. Chlrohydroquinone was apparently the main product in the Pt and Ti/RuO2 cells. Chlorophenol degradation also produced the same aromatic intermediates as phenol electrolysis (Figure 8.5). Most of the aromatic compounds were completed removed by the end of electrolysis

108

Chapter 8: Electrochemical degradation of chlorophenol: Kinetics and pathways


60

50

Ti/SnO2-Sb-Al Pt Ti/RuO2

Current efficiency (%)

40

30

20

10

0 0 1 2 3 4 5 6

Reaction time (hr)

Figure 8.4 MCP degradation as a function of the reaction time for different anodes.

in the Ti/SnO2-Sb-Al system, while a large amount of them were left in the other two cells. Similar to EC phenol degradation, a number of aliphatic acids were found during chlorophenol electrolysis with the three different electrodes (Figure 8.6). The intermediate carboxylic acids included muconic acid, oxoglutaric acid, fumaric acid, maleic acid, succinic acid, malonic acid, acetic acid and formic acid. Muconic acid was the major intermediate in any of the electrolysis cell. Only in the Ti/SnO2-Sb-Al cell muconic acid was completely oxidized. The result of chlorophenol degradation further validated the higher catalytic activity of anode Ti/SnO2-Sb-Al toward organic mineralization. All types of aliphatic acids could be fully oxidized by the Ti/SnO2-SbAl anode, which resulted in complete TOC removal. Also for the other two electrodes, both destruction of chlorophenol and degradation of its intermediates were slower by an order of magnitude or more comparing to the tests with the Ti/SnO2-Sb-Al anode. 8.2.2 Grouping of the intermediate organic compounds

109

Chapter 8: Electrochemical degradation of chlorophenol: Kinetics and pathways The overall organic compounds were grouped into three parts, the parent substrate chlorophenol, the sum of 8 kinds of aromatic compounds (chlorohydriquinone, 4-chlorocatechol, 3-chlorocatechol, phenol, hydroquinone, benzoquinone, 1, 2-benzoquinone, 1, 4-benzoquinone), and the sum of 8 specis aliphatic acids (muconic acid, oxoglutaric acid, fumaric acid, maleic acid, succinic acid, malonic acid, acetic acid and formic acid). Y-axis refers to the sum of organic carbon concentrations that were calculated from the results of HPLC analysis (Figure 8.7). With the destruction of chlorophenol during the EC process, there were accumulations in benzoic compounds, organic acids and polymeric products. The summation of organic carbons in these chemicals demonstrates the change of total organic content from the parent substrate. Within three hours of electrolysis, 99% chlorophenol was destructed and 80% TOC was removed at the Ti/SnO2-Sb-Al anode. However, only 25% organic removal was achieved on the Pt anode. The situation in the Ti/RuO2 cell was even worse with a 28% chlorophenol destruction and a 5% overall TOC removal (Figure 8.7). Till the end of the 6-hr electrolysis tests, almost all organic species were mineralized on the Ti/SnO2-Sb-Al anode. In comparison, more than 50% TOC was still left in the Pt cells and about 80% TOC was left in the Ti/RuO2 cell. There was no obvious rate-limiting step in chlorophenol degradation at the Ti/SnO2-Sb-Al anode. However, oxidation of the organic acids appeared to be the limiting step for the Pt anode. On anode Ti/RuO2, both ring cleavage and organic acid oxidation were apparently the rate limiting steps for complete chlorophenol degradation. 8.2.3 Transformation and degradation of chlorophenol during electrolysis As suggested previously, dechlorination is a crucial step in EC chlorophenol oxidation. The dechlorination reaction could take place in two ways according to the direction of electron transfer, e.g, reductive dechlorination and oxidative dechlorination. If chlorine is the electron donor, the reaction is reductive dechlorination, contrarily the process is oxidative dechlorination. The oxidative reaction of a chlorinated aromatic compound has two effects: the first is the oxidation of the aromatic ring, and the second is the releasing of chlorine. As to the reductive

110

Chapter 8: Chlorophenol incineration: kinetics and pathway

20 Hydroquinone 1,2-benzoquinone 1,4-benzoquinone Catechol Phenol Chlorohydroquinone 3-chlorocatechol 4-chlorocatechol

Ti/SnO2-Sb-Al

Pt

Ti/RuO2

Concentration - C (mg/L)

15

10

20 0

Concentration - C (mg/L)

15

Chlorohydroquinone 3-chlorocatechol 4-chlorocatechol

10

0 3

Concentration - C (mg/L)

Hydroquinone 1,2-benzoquinone 1,4-benzoquinone Catechol Phenol

0 0 60 120 180 240 300 360 0 60 120 180 240 300 360 0 60 120 180 240 300 360

Reaction tieme (min)

Reaction time (min)

Reaction time (min)

Figure 8.5 HPLC analysis of intermediates of EC MCP degradation on the three anodes (aromatic compounds).

111

Chapter 8: Chlorophenol incineration: kinetics and pathway

30

Ti/SnO2-Sb-Al

25

Concentration - C (mg/L)

20

Muconic acid Oxoglutaric acid Fumaric acid Maleic acid Succinic acid Malonic acid Acetic acid Formic acid

Pt

Ti/RuO2

15

10

0 12

10

Concentation - C (mg/L)

Muconic acid Oxoglutaric acid Fumaric acid Maleic acid Succinic acid

0 20 Malonic acid Acetic acid Formic acid 15

Concentration - C (mg/L)

10

0 0 60 120 180 240 300 360 0 60 120 180 240 300 360 0 60 120 180 240 300 360

Reaction time (min)

Reaction time (min)

Reaction time (min)

Figure 8.6 HPLC analysis of intermediates of EC MCP degradation on the three anodes (aliphatic acids).

112

Chapter 8: Chlorophenol incineration: kinetics and pathway

60 Chlorophenol-C Quinones-C Acids-C

50

Concentration -C (mg/L)

40

30

20

10

0 0 60 120 180 240 300 360

Reaction time (min)


Figure 8.7 TOC components of intermediates of EC MCP degradation by the three anodes: Ti/SnO2-Sb-Al (left), Pt (middle), and Ti/RuO2 (right).

113

Chapter 8: Chlorophenol incineration: kinetics and pathway

OH [O] OH Cl [O] OH

OH

O [O]

OH

Hydroquinone Chlorophenol Phenol


Table 1: Phenol percentage
OH OH

OH

Benzoquinone
O

[O]

[O]

Time Sn Pt Ru

30 0.5% 0.5% 0.7%

360 0 0.1% 0.7%


OH Cl OH OH Cl OH Cl

Polymers

Catechol

Benzoquinone
B OH OH

Acids

Table 2: Chlorophenol percentage

Time Sn Pt Ru

30 1.5% 16% 0.8%

360 55% 1.3% 0.9%

Chlorohydroquinone 3-Chlorocatechol 4-Chlorocatechol

Figure 8.8 Proposed MCP electrolysis pathways for the three anodes. 114

Chapter 8: Electrochemical degradation of chlorophenol: Kinetics and pathways dechlorination, the chloride ions are released from the chemical. The degradation of chlorophenol may consist of two different phases. The first one was dechlorination for a chlorinated chemical compound. Dechlorination could take place during chlorophenol destruction. However, it might also take place for different chlorinated intermediates, sometimes even after ring cleavage. The other phase of chlorophenol degradation would be the further oxidation of dechlorinated organic compounds, which was comparable to the process of EC phenol degradation. The reactions included the oxidation of hydroxyl groups and double bonds in the phenolic ring, the subsequent ring cleavage, and the oxidation and mineralization of the aliphatic acids. It appeared that most dechlorination took place before the ring breakage. There were little chlorinated aliphatic acids detected in the electrolyzed solutions. 8.2.4 Basic pathways of EC chlorophenol degradation

Figure 8.9 LSV of anode Ti/SnO2-Sb-Al in MCP solution.

115

Chapter 8: Electrochemical degradation of chlorophenol: Kinetics and pathways According to the above analysis and the results of EC phenol degradation study, the pathways for chlorophenol mineralization by electrolysis were proposed for the different electrodes (Figures 8.8). As to the part of acid destruction, the pathways were similar to phenol listed in Chapter 6. The majority of dechlorination apparently was advanced than ring cleavage, which transformed a large portion of chlorophenol into phenol accompanied by releasing a chlorine form the ring. Both oxidative dechlorination and reductive dechlorination could take place during the EC process. The following oxidation pathways of the organic intermediates after dechlorination should be similar to those proposed for EC phenol oxidation. In a simplified expression, the degradation of chlorophenol by electrolysis could be summarized as:
[O ] MCP [O ] [O ] Phenol aromatic compounds aliphatic acids [ O ] CO2 + H 2 O [O ] chloride aromatic compounds aliphatic acids

8.3 Voltammetric study of the anodes in monochlorophenol solutions

8.3.1 LSV of the electrodes in chlorophenol solutions

Figure 8.10 LSV of anode Pt in the MCP solution.

116

Chapter 8: Electrochemical degradation of chlorophenol: Kinetics and pathways Measurements of the electrodes began from the pure electrolyte, 0.25 M Na2SO4 solution, without chlorophenol produced a rather smooth LSV curve (Figures 8.9). Similar to the phenol study, addition of chlorophenol into the electrolyte resulted in new shoulders or peaks in the LSV curve of the Ti/SnO2-Sb-Al anode (Figure 8.9). The peak current increased significantly from 1.2 to 6 mA/cm2 as the chlorophenol concentration increased from 1,000 to 10,000 mg/L. Respecting to the peak current, the voltage was located within 1.1- 1.4 V for different chlorophenol concentrations. In the potential range below 1.6 V, significant oxygen evolution was not expected from the Ti/SnO2-Sb-Al anode. Hence these peaks suggest direct anodic oxidation reaction during the LSV tests. The Ti/SnO2-Sb-Al anode apparently was able to oxidize chlorophenol directly on the electrode surface. A higher chlorophenol concentration accelerated the oxidation reaction rate at the anode surface. On the Pt anode, current peaks were observed in the LSV curves with MCP addition, and increased with the chlorophenol concentration (Figure 8.10). Chlorophenol also could be oxidized directly at the Pt surface. However, the oxidation

Figure 8.11 LSV of anode Ti/RuO2 in the MCP solution.

117

Chapter 8: Electrochemical degradation of chlorophenol: Kinetics and pathways rate was not as faster as that at the Ti/SnO2-Sb-Al anode. The corresponding chlorophenol oxidation potential with 10,000 mg/L at Pt was 1.0 V, lower than the potential at the Ti/SnO2-Sb-Al anode. Increasing the chlorophenol concentration to 10,000 mg/L led to a raise of the current density up to 1.8 mA/cm2. In comparison, an increase in chlorophenol concentration brought more electron transfer into EC reaction at the Ti/SnO2-Sb-Al anode surface. Hence, Ti/SnO2-Sb-Al anode had a greater reactivity towards direct chlorophenol oxidation on its surface, which has been demonstrated by the LSV tests for phenol solutions. With the Ti/RuO2 anode, however, chlorophenol addition into the electrolyte caused little change in the LSV curves. There were no peaks and shoulders found in the LSV curves as the chlorophenol concentration increased (Figure 8.11). The result suggests that the potential of possible chlorophenol oxidation at the Ti/RuO2 anode was higher than the potential of water electrolysis (Arikawa, 1998; Simond, 1997). Only an insignificant fraction of charge input could be used for chlorophenol oxidation. In other words, direct chlorophenol oxidation on anode Ti/RuO2 was much more difficult than that on the Ti/SnO2-Sb-Al anode. Oxygen evolution was the

25

20

Current dencity (mA/cm )

1000 mg/L MCP 5000 mg/L MCP 10000 mg/L MCP

15

10

0 6 8 10 12 14 16 18 20 22 24

Scan rate square root

Figure 8.12 Current density as a function of the scan rate square root

118

Chapter 8: Electrochemical degradation of chlorophenol: Kinetics and pathways dominant electrochemical reaction that allowed little increase in potential and hence hindered the direct chlorophenol oxidation on the Ti/RuO2 anode surface. 8.3.2 Peak current densities versus CV scan rate The mass transfer behavior of chlorophenol onto the electrode surface also could be investigated by means of cyclic voltammetry (CV). Figure 8.12 illustrates the peak current density as a function square root of the scan rate for the oxidation of chlorophenol at the anode Ti/SnO2-Sb-Al. Three different chlorophenol concentrations were tested for the correlation. In general, an increase in the scan rate square root would cause a linear increase as a response in the oxidation peak current density. A faster LSV scan rate could result in more current flow for chlorophenol oxidation. However, with an increase in the chemical concentration, the current density increased at a much faster rate with the LSV scan rate. A higher chlorophenol concentration resulted in a faster oxidation reaction on the anode surface. This suggests that the EC decomposition of chlorophenol was probably controlled by the mass transfer of the chemical from the aqueous phase to the anode surface. A higher concentration would allow a faster mass transfer rate to the anode interface, which would facilitate the oxidation reaction. Chemical mass diffusion, rather than electron transfer at the anode, appeared to be the rate-limiting step for chlorophenol electrolysis during the LSV process. In other words, the rate of EC reaction catalyzed by the active Ti/SnO2-Sb-Al surface was faster than the rate of substrate transfer to the anode surface.

8.4 EC Chlorophenol degradation pathways

8.4.1 Electrolysis mechanisms Electrode Ti/SnO2-Sb-Al had a much better performance in electrochemical chlorophenol degradation likely due to two main reasons. The first was the direct chemical oxidation on the anode surface. This was indicated by the peaks of current density in the LSV curve for the addition of chlorophenol in the electrolyte. Electrolysis of chlorophenol allowed a new current peak during the LSV test. Thus, power could be utilized for direct degradation of chlorophenol on the anode. Another 119

Chapter 8: Electrochemical degradation of chlorophenol: Kinetics and pathways reason could be the generation of hydroxyl radicals from water electrolysis during the EC process. According to the LSV experiments, the Ti/SnO2-Sb-Al had a higher overpotential of oxygen evolution than the other two anodes. Chlorophenol and other organics could be easily destructed by the high-energy free radicals in the vicinity of the anode. The LSV curves in the pure electrolyte showed that the Pt and Ti/RuO2 anodes had a lower oxygen evolution potential, with respective values of around 1.4 and 1.1 V vs. the reference of the saturated Ag/AgCl electrode. In comparison, Ti/SnO2-Sb-Al anode had a considerably higher oxygen evolution potential of 1.8 V vs. the reference. It is believed that oxygen molecules are formed by radical reactions involving the OH radical generated electrochemically. The high overpotential of the Ti/SnO2-Sb-Al anode suggests that the radical reaction forming molecular oxygen was probably restrained. This would facilitate the generation and accumulation of OH radical at the electrode surface under a higher potential condition. In contrast, oxygen evolution took place at a rapid rate at a lower potential on the Pt and Ti/RuO2 anodes, which was unfavorable to the production and accumulation of OH radicals. Chemical concentration did show a great effect on the effectiveness of the EC treatment process. A higher chlorophenol concentration caused the increase of the peak current density on the Ti/SnO2-Sb-Al anode, suggesting a faster chlorophenol oxidation on the anode. EC chlorophenol degradation would take place at a faster rate for a higher substrate concentration. This is why a faster reaction kinetics and a greater current efficiency were obtained in the early phase of EC chlorophenol treatment than the later phase when the organic concentration became lower. 8.4.2 Analysis of chlorophenol degradation pathways Among the intermediates, the concentration of chlorohydroquinone was much higher than others in the solution electrolyzed by anode Pt. Different from the situation with the Ti/SnO2-Sb-Al anode for which dechlorination advanced than ring cleavage, the majority of dechlorination apparently took place after ring cleavage on the Pt anode. The accumulation of chlorohydroquinone in the cell of Ti/RuO2 was not as higher as that in the Pt cell. About half of dechlorination took place before ring cleavage and half after ring cleavage (Figure 8.8). Only in the cell of Ti/SnO2-Sb-Al, there was no significant accumulation of aromatic compounds throughout the 6 hours 120

Chapter 8: Electrochemical degradation of chlorophenol: Kinetics and pathways of electrolysis. The amount of intermediates in the Pt cell decreased at a slow rate. However, in the Ti/RuO2 cell the decrease of intermediate organics was hardly detected (Figure 8.5). Further degradation of aromatic intermediates produced different kinds of aliphatic acids (Figure 8.6). In the cell of Ti/SnO2-Sb-Al, there was transformation of aliphatic acids, which were reduced to zero by the end of EC treatment. In the cells of the other two electrodes, there was a trend of accumulation of the organic acids, which comprised a large fraction of the TOC. These results further confirmed that the degradation rates of aliphatic acids on anodes Pt and Ti/RuO2 were much slower compared with that on the anode Ti/SnO2-Sb-Al. Muconic acid, a 6-carbon acid, was the major product of oxidation of aromatic ring cleavage. Its concentrations in all three cells of the electrodes were much higher than other intermediates. Its long carbon chain apparently increased the difficulty of its degradation. In the cell of Ti/SnO2-Sb-Al, the non-selective catalytic activity of the anode was demonstrated as all acids were mineralized effectively. For the Pt cell, similar to the phenol treatment test, degradation of the organic acids was actually slower than the destruction of aromatic compounds. In the Ti/RuO2 cell, organic reduction was quite minimal, and the anode hardly showed any catalytic effect toward the oxidation of chlorophenol and its intermediate products.

8.5 Summary

The results of chlorophenol degradation on the three different electrodes were similar to those obtained from the phenol degradation study. The Ti/SnO2-Sb-Al anode was highly effective for the oxidation of chlorophenol and its intermediates. Complete TOC removal from 100 mg/L could be achieved after 6 hours of electrolysis. The Pt anode was less effective for chlorophenol degradation, which gave a partial TOC reduction. The Ti/RuO2 electrode was the least effective anode for chlorophenol treatment with an insignificant TOC removal after 6 hours of electrolysis under the same EC conditions. The current efficiencies in the early phase chlorophenol electrolysis were 55% for anode Ti/SnO2-Sb-Al, 8% for anode Pt and 5% for anode Ti/RuO2. Dechlorination was a critical step in chlorophenol degradation. Dechlorination was not difficult for the Ti/SnO2-Sb-Al electrode, which could be mostly achieved before aromatic ring cleavage. In contrast, dechlorination was more 121

Chapter 8: Electrochemical degradation of chlorophenol: Kinetics and pathways difficult on the Pt anode and much more difficult on the Ti/RuO2 anode. After dechlorination, the organic degradation pathways were comparable to those obtained for EC phenol degradation. The aromatic intermediates were subsequently oxidized rapidly by the Ti/SnO2-Sb-Al anode, but accumulated in the cells of Ti/RuO2, and Pt. There were more chlorohydroquinone found in the Pt cells and both hydroquinone and benzoquinone were found at a high level in the Ti/RuO2 systems. The intermediates according to HPLC analysis included aromatic compounds of chlorohydriquinone, 4-chlorocatechol, 3-chlorocatechol, phenol, hydroquinone, benzoquinone, 1, 2-benzoquinone, 1, 4-benzoquinone and aliphatic acids of muconic acid, oxoglutaric acid, fumaric acid, maleic acid, succinic acid, malonic acid, pyruvic acid, oxalic acid, actic acid and formic acid. The experimental results further indicate that the Ti/SnO2-Sb-Al electrode has a non-selective catalytic capability toward organic oxidation, which is highly suitable for treatment of toxic organic pollutants.

122

Chapter 9: Discussion: Comparison of the surface morphology and properties between different electrodes

Chapter 9

Discussion: Comparison of the surface morphology and properties between different electrodes
9.1 Comparison of surface properties of different electrodes

9.1.1 Coating density The preparation method determined the coating thickness on the electrode. The coating weight density was about 4 mg/cm2 on the electrode Ti/SnO2-Sb-Al. A sufficient thickness of coating could effectively protect the substrate base and prohibit titanium dioxide formation. The optimal range of coating weight was defined by several parameters, such as the substrate, catalytical property, coating materials and also the preparation methodology. 9.1.2 Deactivation of electrode Ti/SnO2-Sb-Al 9.1. 2.1 The mechanism of deactivation Electrode deactivation may be caused two reasons, the formation of titanium dioxide and the peeling away of the coating layer. Firstly, the naked titanium metal substrate uncovered by the coating layer will soon be oxidized under anodic ambient. Moreover, the porous structure and the creaks of the coating film would allow the penetration of active oxygen atoms through the coating layer to reach Ti surface. The oxidation of titanium substrate lowered the electrodes catalytical activity and shortened its service life. Secondly, peeling away of the active coating film was another reason which caused deactivation. Sol-gel attachment and thermal decomposition were selected for adhesion of coating materials and formation of the

123

Chapter 9: Discussion: Comparison of the surface morphology and properties between different electrodes active compounds. However there were mud cracks formed on the electrode coating surface. Small cracks could be easily peeled off, which caused the loss of catalytic coating and made the titanium metal, exposed to the oxidation atmosphere. 9.1.2.2 Images of the titanium plates The fresh titanium plate after etching showed a rough surface, a lackluster gray color under a microscope at 500 (Figure 9.1 A). Comparing to a coated Ti/SnO2-Sb-Al electrode (Figure 9.1 B), the latter one appeared smoother with the coating film. Figure 9.2 provided a clear morphology of the electrode surface by SEM, comparison between untreated and treated surfaces. Figure 9.2 A was a vision of the titanium plate surface after etching. The Ti electrode surface was not as smooth as Pt plate (Figure 9.2 B), and the Ti surface area was effectively. The Ti electrode after coating was shown in Figure 9.2 C. All those holes were filled with coating materials and the plate became smoother.

Figure 9.1 Titanium plate under microscope

9.1.2.3 Deactivation of electrode Ti/SnO2-Sb-Al The first step was the generation of bubble-like points. In Figure 9.2 D, many of small bubble points generated from the original smooth surface. The unevaporated solvent of precursor salts might be oxidized by the penetrating oxygen atom, forming these bubble like points in Figure 9.2 D. More and more cracks generated from these

124

Chapter 9: Discussion: Comparison of the surface morphology and properties between different electrodes sports on the electrodes coating surface during electrolysis (Figure 9.2 E), which accelerated the penetration of oxygen atom. The cracks split the coating surface into small pieces, accompanied by the shear of disturbing, coating materials peeled off from the electrode surface (Figure 9.2 F). In general, the deactivation of the Ti/SnO2Sb-Al anode underwent through a three-step process. The incomplete evaporation of sol-gel solvent could be the major reason. The formation of carbon dioxide might cause the formation of new cracks on the electrode surface and finally led the peeling of the coating materials.

Figure 9.2 SEM images of the electrode in different service stages.

125

Chapter 9: Discussion: Comparison of the surface morphology and properties between different electrodes 9.1.3 Coating surface components 9.1.3.1 XPS analysis Figure 9.3 shows the full scan results of the electrode Ti/SnO2-Sb-Al from zero to 1100 ev for testing the component of the coating materials and their locations. On the electrode surface, the major metal component was Sn, which had the highest response peak during the full scan process. Simultaneously, different peaks of Sn showed its different locations. Elements Sb, C, and O also could be detected. The valence of Sn located at the highest point also was confirmed as 4+, though accompanied with other values (Safonova, 2000; Amanullah, 1998). The major dopant element Sb also could be found by XPS even though the amount was quite small with a low peak. Aluminium could not be detected. Oxygen was another nonmetal component with the metal elements. In general, Figure 9.3 provided an overview of the coating materials on electrode Ti/SnO2-Sb-Al.

Sn3d5/2
Figure 9.3 A typical full-width XPS spectrum of the Ti/SnO2-Sb-Al electrode

Sn3d3/2 O1s Sb3d3/2 C1s

9.1.3.2 EDX analysis 126

Chapter 9: Discussion: Comparison of the surface morphology and properties between different electrodes Electrodes Ti/SnO2-Sb-Al and Ti/RuO2 were tested and the results are shown in Figures 9.4 and 9.5. EDX scan also provided full information of the components of surface coating materials. SEM-EDX combination graphs afforded a 3-dimensional view of the electrode surface. The EDX scan was carried out based on the scan of SEM in the part through the red line. The surface of electrode Ti/SnO2-Sb-Al

Element CK OK Cl K Ti K Ru L Sn L Sb L Totals

Weight% 2.77 20.67 0.53 4.79 7.18 8.37 3.90 48.20

Atomic% 12.73 71.34 0.83 5.52 3.92 3.89 1.77

Figure 9.4 EDX analysis of electrode Ti/RuO2.

127

Chapter 9: Discussion: Comparison of the surface morphology and properties between different electrodes appeared uniform with many evenly distributed small cracks. The coating materials had four kinds of metal elements on the electrode Ti/RuO2, Ti, Ru, Sn, and Sb. The total amount of atomic metal (MT) percentage was 15.10%, and the atomic oxygen percentage was 71.34%. The major components of coated electrode Ti/SnO2-Sb-Al was Sn, with the atomic percentage as high as 17.21%, which is far larger than any element in electrode Ti/RuO2. The high concentration of element Sn apparently increased the catalytic activity of the electrode. 9.1.4 Contact angle and hydrophilicity Contact angles were measured with digital photos (Figure 9.6), which recorded the state of drops when they kept stable on the surface of different electrodes. Higher hydropilicity surface could greatly decrease the contact angle while drops were imbibed of the surface. On the electrode Ti/SnO2-Sb-Al, the water drop has shown a flat shape with a contact angle of 17. The high hydrophilicity of electrode Ti/SnO2-Sb-Al greatly decreased the surface tension of the water drop. The water drops on the electrodes Ti/RuO2 and Pt showed a similar shape. There was not much difference between their contact angles, 67 and 72, respectively. The low hydropolicity of electrode Ti/RuO2 and Pt may prohibit the transformation of organic molecules from water phase to the solid surface. The right column of Figure 9.6 was another method for contact angle measurement. The advancing angle and receding angel of three electrodes were listed in Table 9.1. Electrode Ti/SnO2-Sb-Al differed greatly from the other two electrodes of the advancing angle and receding angles. A plate shape of water drop also has a low value of the advancing angle of 31.75, and a low receding angle even less than 10. Electrode Ti/RuO2 and Pt, with a similar shape of water drop attached on the surface, have the advancing contact angles of 69.75 and 81.5, respectively.
Table 9.1 Contact angle (advancing and receding contact angle) measurement.

Ti/SnO2-Sb-Al Ti/RuO2 Pt 17 67 72

r
8 27.25 28.5

a
31.75 69.75 81.5

128

Chapter 9: Discussion: Comparison of the surface morphology and properties between different electrodes

Element OK Al K Sn L Totals

Weight% 13.32 0.25 20.78 34.35

Atomic% 81.88 0.90 17.21

Figure 9.5 EDX analysis of electrode Ti/SnO2-Sb-Al.

9.1.5 Organic adsorption onto the electrodes The adsorption capability comparison between the three kinds of electrodes is listed in Table 9.2. Here, two aromatic compounds and two aliphatic acids were selected as model substrates for the adsoption tests.

129

Chapter 9: Discussion: Comparison of the surface morphology and properties between different electrodes

II

Sn

Pt

Ru

Figure 9.6 Contact angle measurement.

130

Chapter 9: Discussion: Comparison of the surface morphology and properties between different electrodes Electrode Ti/SnO2-Sb-Al showed better adsorption of aromatic compounds than Ti/RuO2 but worse than Pt. Adsorption maybe also correlated to the contamination of electrode. Electrode Ti/RuO2, lowest adsorption to aromatic acids prohibited further degradation of intermediates, appeared to have a great resistance to contamination. Metal Pt has a high specific adsorption of aromatic compounds, which would limit further oxidation reactions on the electrode surface.
Table 9.2 The quantities of organics adsorbed on different electrodes.

Phenol mg/L Sn Ru Pt 1.0 0.2 7.75 mg/m2 0.25 0.05 1.94

Benzoquinone mg/L 1.10 0.15 3.93 mg/m2 0.28 0.04 0.98

Fumaric acid mg/L 2.45 2.18 0.30 mg/m2 0.61 0.55 0.08

Maleic acid mg/L 2.25 1.94 0.52 mg/m2 0.56 0.49 0.13

9.2 Comparison of different electrodes for organic electrolysis

9.2.1 EC degradation of organic matter On phenol degradation from the initial 100 mg/L, electrode Ti/SnO2-Sb-Al took 3 hrs to achieve complete destruction and 5 hrs by the Pt anode. At the same current input rate to the Ti/RuO2 anode, only 42% phenol removal was achieved after 6 hrs of electrolysis. Complete TOC removal was achieved after 5 hrs on the Ti/SnO2Sb-Al anode. However, just 36% and 12% TOC removals were reached by the Pt and Ti/RuO2 anodes, respectively. The order of organic removal efficiency is: Ti/SnO2Sb-Al > Pt > Ti/RuO2. Also, chlorophenol was destructed nearly completely after 4 hrs by the Ti/SnO2-Sb-Al anode from the initial 100 mg/L. In comparison under the same current input condition, 55% chlorophenol was oxidized by the Pt anode and only 10% chlorophenol was destructed by the Ti/RuO2 anode. Complete TOC removal was achieved after 5 hrs on the Ti/SnO2-Sb-Al anode, and just 30% and 10% TOC removals were produced by the Pt and Ti/RuO2 anodes, respectively. The same order of chlorophenol organic removal capability is: Ti/SnO2-Sb-Al > Pt > Ti/RuO2.

131

Chapter 9: Discussion: Comparison of the surface morphology and properties between different electrodes 9.2.2 Organic degradation pathways In a simple way, the phenol degradation routes could be written as:
1O ] 2O 3O Phenol [ Aromatic Compounds [] Aliphatic Acids [] CO2 Electrode Ti/SnO2-Sb-Al is effective in any step of oxidation [O], and there is no

limiting step from phenol to the final carbon dioxide. Plate Pt is effective in the first two steps of oxidation and stop at aliphatic acid oxidation. The major components are aliphatic acids after 6 hrs of electrolysis. The chemical aliphatic acid degradation on electrode Pt was almost zero except oxalic acid. Complete mineralization is impossible on plate Pt. Phenol degradation on electrode Ti/RuO2, the whole process rate is 1/8 to the electrode Ti/SnO2-Sb-Al, and 1/3 to plate Pt. On chlorophenol degradation, dechlorination is another step that showed the great difference for the three anodes. After 2 hrs of electrolysis on Ti/SnO2-Sb-Al, when 95% chlorophenol was destructured, the chloride concentration reached the maximum. Chloride release was detected at a much slower rate in the chlorophenol solutions electrolyzed by the Pt and Ti/RuO2 anodes. It is apparent that dechlorination was an important limiting step in the electrolysis of chlorophenol by electrodes Pt and Ti/RuO2, which led to much slower chlorophenol destruction and organic degradation. 9.2.3 Surface electrochemical properties Oxygen evolution potential may be closely correlated to the OH radical formation and the electrolysis oxidation process. The Ti/SnO2-Sb-Al anode had a considerably higher oxygen evolution potential of 1.8 V vs. the reference of the saturated Ag/AgCl electrode. This would facilitate the generation and accumulation of OH radicals at the electrode surface under a higher potential condition. The Pt and Ti/RuO2 anodes had a lower oxygen evolution potential, with respective values of around 1.4 and 1.1 V vs. the reference, which was unfavorable to the production of OH radicals. During the LSV tests, the peaks of direct anodic oxidation of phenol were located at about 1.1- 1.5 V for different phenol concentrations to the Ti/SnO2-Sb-Al anode. Phenol could also be oxidized directly at the Pt surface, with smaller peaks observed in the LSV curves. However, the rate of oxidation was much slower than that at the Ti/SnO2-Sb-Al anode. The corresponding phenol oxidation potential at Pt 132

Chapter 9: Discussion: Comparison of the surface morphology and properties between different electrodes was somewhat lower than the potential on the Ti/SnO2-Sb-Al anode. Direct phenol oxidation on the Ti/RuO2 anode was much more difficult. Oxygen evolution was the dominant electrochemical reaction that allowed little increase in potential and hence hindered direct phenol oxidation on the Ti/RuO2 anode surface. Similar results could be found on chlorophenol degradation. 9.2.4 EC oxidation mechanisms Electrode Ti/SnO2-Sb-Al had a much better performance in electrochemical organic degradation likely due to two main reasons. The first was the direct chemical oxidation on the anode surface. Power could be utilized for direct degradation of substrates on the anode. This was indicated by the peaks of current density in the LSV curve for the addition of organic matter in the electrolyte. Except oxygen evolution, organic solution electrolysis allowed a new current peak during the EC test. Another reason could be the generation of hydroxyl radicals from water electrolysis during the EC process. According to the LSV experiments, the Ti/SnO2-Sb-Al had a higher overpotential of oxygen evolution than electrodes Pt and Ti/RuO2. Organics could be easily destructed by the high-energy free radicals in the vicinity of the anode. In the pure electrolyte water discharge showed that the Ti/SnO2-Sb-Al anode had a considerably higher oxygen evolution potential of 1.8 V vs. the reference of the saturated Ag/AgCl electrode. Pt and Ti/RuO2 anodes had lower oxygen evolution potentials, with respective values of around 1.4 and 1.1 V vs. the reference. It is believed that oxygen molecules are formed by radical reactions involving the OH radical generated electrochemically. The high overpotential of the Ti/SnO2-Sb-Al anode suggests that the radical reaction forming molecular oxygen was probably restrained. This would facilitate the generation and accumulation of OH radical at the electrode surface under a higher potential condition. In contrast, oxygen evolution took place at a rapid rate at a lower potential on the Pt and Ti/RuO2 anodes, which was unfavorable to the production and accumulation of OH radicals. The element analysis of surface coating layer illustrate that electrode Ti/SnO2Sb-Al has 17.21% atomic percentage of Sn. The higher proportion of active atomic would increase the catalytic activity of the coating film. The atomic concentration of Ru in the coating of Ti/RuO2 is 3.92%, organic matter degradation was limited by the limiting activator metal atoms. Additionally, the doping element Ti and Sn are 5.52% 133

Chapter 9: Discussion: Comparison of the surface morphology and properties between different electrodes and 3.89%, respectively. On the other hand, the percentage of oxygen of the coating layer is another limiting issue. Anode Ti/SnO2-Sb-Al has a higher percentage of 81.88%, about 10% higher than electrode Ti/RuO2. Higher concentration prohibits the transformation of oxygen from the outside into the crystal lattice of metal oxide, which helps the accumulation of OH on the electrode in favor of organic matter oxidation. 9.2.5 Stability of the electrodes Electrode Ti/SnO2-Sb-Al has far better performance than the other two electrodes for organic degradation. A slow deactivation process is accompanied by a slow increase of voltage on electrode Ti/SnO2-Sb-Al. A well prepared electrode could support continuous 16 hrs of electrolysis, which is the deathful shortcoming of this electrode. Pt fouling is the limiting of its application and its cost as well. Pt is easily be contaminated by substrate matters and their intermediates, even can not sustain several times of LSV scanningf. Electrode Ti/RuO2 has a long service life for its application in alkali industry, and the cycle for electrode replacement is normally ten years.

9.3 Future work

9.3.1 Improvement of the electrode preparation techniques Electrode Ti/SnO2-Sb-Al is good at toxic organic removal while is poor at its service life. More development work should be focused on extension of its service life. Introducing another metal element might increase the stability, and adjusting the method for internal layer preparation might also advance its service life. 9.3.2 Investigation of EC organic oxidation mechanisms The EC oxidation mechanism was only discussed from electrochemical properties, electrochemical reaction direction and the components of surface coating materials. Further discussion from the aspect of energy should be carried out. Electrolysis is a power consumption process and the electrochemical reaction 134

Chapter 9: Discussion: Comparison of the surface morphology and properties between different electrodes accompanied by the energy transfer. Surface materials are also related to the energy matter. The structure of metal oxide crystal lattice and the transformation of electrons and holes involve energy.s. Hydrophilicity discussion is incomplete and also should be advanced by energy calculation.

135

Chapter 10: Conclusion

Chapter 10

Conclusions
10.1 Electrochemical treatment of saline sludge centrate and landfill leachate for denitrification and organic removal

A lab-scale electrochemical (EC) system with Ti-based electrode plates was used for treatment of saline sludge liquor from a sewage treatment works and raw leachate from a landfill. The experimental results demonstrate that the EC process is highly effective for denitrification of the high-strength wastewaters. For the saline sludge with a NH4+-N of around 500 mg/L, complete nitrogen removal could be achieved within a contact time of 1 hr. The rate of EC denitrification increased with the current intensity applied. The best current efficiency for nitrogen removal was obtained with a gap distance between the electrodes at 8 mm, and the power consumption rate for nitrogen removal of the sludge liquor was around 23 kWh/kg N. For the landfill leachate with a chloride content of about 2500 mg/L, nitrogen could be removed from an initial ammonia concentration of more than 1600 mg/L by 3 hrs of electrolysis. The power consumption rates were similar at around 30kWh/kg-N for both the batch reactor and CFSTR for nitrogen removal from the landfill leachate. Electro-chlorination is considered to be the major mechanism for EC denitrification. Nonetheless, there was no significant accumulation of chloramines during the EC reactions. The formation of chlorination by-products (CBPs) appeared to be minimal as the total trihalomethanes (THM) were below 300 g/L. Additional electro-flocculation with a pair of iron needle electrodes could increase the organic removal by sedimentation from less than 30% to more than 70%. Therefore, the EC process including both electro-denitrification and electro-flocculation can be developed as a cost-effective treatment method for saline sludge liquor and landfill leachate.

136

Chapter 10: Conclusion


10.2 Degradation of phenol and chlorophenol by electrolysis with the Ti/SnO2Sb-Al and other electrodes

A double-layer coating technique was employed to make catalytic active anodes in Ti/SnO2-Sb-Al for electrochemical oxidation of toxic organic pollutants. The EC treatment experiments were conducted for phenol and chlorophenol using the Ti/SnO2-Sb-Al electrode, in comparison with the performance of other electrodes including Pt, Ti/RuO2 or Ti/IrO2. Phenol with an initial concentration of 100 mg/L could be readily destructed on the Ti/SnO2-Sb-Al anode; however, its degradation was considerably slower at the Ti/RuO2, Ti/IrO2 and Pt anodes. For organic reduction, complete TOC removal was achieved after 5 hrs on the Ti/SnO2-Sb-Al anode, however, under the same electrolysis conditions at a current density of 20 mA/cm2, just 36% and 12% TOC removals were reached by the Pt and Ti/RuO2 anodes, respectively. The intermediate products of EC phenol degradation, including aromatic compounds and aliphatic acids, were oxidized rapidly by the Ti/SnO2-Sb-Al anode, but accumulated in the cells of Ti/RuO2 and Pt anodes. The HPLC results show almost no considerable accumulation of aromatic compounds in the solution electrolyzed by anode Ti/SnO2-Sb-Al, while more benzoquinone were found in the Pt cell and both hydroquinone and benzoquinone accumulated in the cell of Ti/RuO2. Aliphatic acids were the predominant intermediates produced by EC phenol oxidation and also the subsequent benzoquinone oxidation. Oxoglutaric acid and oxalic acid were the most dominant acids in the Ti/SnO2-Sb-Al cell, and those acids were reduced to zero in concentration after 6 hrs. For the other two cells with the Pt and Ti/RuO2 anodes, however, organic acids accumulated in the solutions. In addition, there was apparent formation of colored polymeric compounds during phenol electrolysis. Similarly, in the electrolysis of chlorophenol, the best treatment performance was obtained with electrode Ti/SnO2-Sb-Al, followed by Pt and then Ti/RuO2. Complete chlorophenol destruction from a concentration of 100 mg/L could be achieved after 4 hrs of electrolysis by the Ti/SnO2-Sb-Al anode. However, only 55% and 10% of chlorophenol could be destructed by the Pt and Ti/RuO2 anodes, respectively, during the same electrolysis period. For organic removal, 100% TOC reduction could be achieved on the Ti/SnO2-Sb-Al anode, while the TOC removals were only 30% and 10% for the Pt and Ti/RuO2 anodes. The chloride ions in solution 137

Chapter 10: Conclusion from dechlorination of chlorophenol apparently increased the rate of chlorophenol electrolysis compared to the tests of phenol degradation. There were more chlorinated aromatic intermediates detected by HPLC, including chlorohydroquinone and chlorocatechol. However, most of chlorophenol was destructed into phenol which was further oxidized. Based on the EC degradation performance and HPLC results modified pathways of EC phenol degradation at different types of electrodes were proposed. It is argued that under the electro-catalytic condition at the Ti/SnO2-Sb-Al anode, the aromatic ring cleavage takes place rapidly. Maleic acid formed from ring cleavage could be oxidized directly to oxalic acid that can be readily mineralized by EC oxidation. However, at the Ti/RuO2 and Pt anodes without significant catalytic functions, ring breakage became more difficult. The accumulation of intermediate aromatic compounds, such as hydroquinone and benzoquinone, would lead to the formation of more stable polymeric chemical compounds.

10.3 Surface properties of electrode Ti/SnO2-Sb-Al and their effects on EC organic degradation

Different electrodes were characterized for their electrochemical behaviors and surface properties. The three types of electrodes performed rather differently during LSV tests. The overpotential of oxygen evolution was 1.9V for the Ti/SnO2Sb-Al anode, which were considerably higher than 1.1V for the Pt and 1.3V for the Ti/RuO2 anodes. Addition of phenol in the testing electrolyte resulted in new peaks in the LSV curves for anode Ti/SnO2-Sb-Al, which suggests direct phenol oxidation on the anode surface. However, for the Pt and Ti/RuO2 anodes, phenol addition did not change their voltammograms significantly. The surface property of the electrodes appeared to affect the reactions involved in EC organic degradation. In comparison to the conventional Pt and Ti/RuO2 anodes, there could be two main reasons for the high effectiveness of the Ti/SnO2-Sb-Al anode in phenol and chlorophenol degradation. First, its high overpotential of oxygen evolution would allow more production and accumulation of hydroxyl radicals near the anode surface under a higher potential condition. Organic matter could be readily destructed by the high-energy, nonselective oxidant of free radicals. In contrast, oxygen evolution took place at a rapid rate at a lower potential on the Pt and Ti/RuO2 anodes, which was unfavorable to the 138

Chapter 10: Conclusion generation of hydroxyl radicals. Second, the peaks in the LSV curves of the Ti/SnO2Sb-Al anode in phenol solution indicated direct organic oxidation on the anode surface. Thus, power could be utilized for direct oxidation of chemical compounds on the anode, which was much difficult to be achieved with the Pt and Ti/RuO2 anodes. In addition, the surface properties, such as morphology, hydrophobicity and chemical adsorption potential, could also affect various reaction steps of EC organic degradation. The Ti/SnO2-Sb-Al electrode appeared to have a high affinity with organic substances than electrode Ti/RuO2. Better adsorption of the organic on the anode surface would improve electron transfer for direct organic oxidation and allow a more effective indirect oxidation by hydroxyl radicals generated on the anode surface. The element analysis of surface coating layer illustrate that electrode Ti/SnO2-Sb-Al had 17.2% atomic Sn. The higher proportion of active metal atomic would increase the catalytic activity of the coating film toward organic oxidation. In brief, specific anode surface treatment such as the SnO2-Sb-Al coating provided the anode with an apparent catalytic function for rapid organic degradation. The effective EC organic oxidation was probably brought about by both direct electron transfer of the organic on the anode surface and the attack of hydroxyl radicals generated from anodic water electrolysis. However, some anode materials, such as Ti/RuO2 and Pt, were less favorable for use in the EC treatment of organic pollutants, because their lack of the electro-catalytic capacity and the production of some intermediates that were more refractory to EC and additional treatment.

139

Reference

References
Agustina T.E., Ang H.M. and Vareek V.K. (2005). A review of synergistic effect of photocatalysis and ozonation on wastewater treatment.

Journal

of

Photochemistry and Photobiology C: Photochemistry Reviews. 6(4): 264-273.


Alejandra A., Medina F., Salagre P., Fabregat A. and Sueiras J. E. (1998). Characterization and activity of copper and nickel catalysts for the oxidation of phenol aqueous solutions. Applied Catalysis B: Environmental. 18(3-4): 307-315 Alves V.A,, Silva L.A. da and Boodts J.F.C. (1998). Electrochemical impedance spectroscopic study of dimensionally stable anode corrosion. Journal of

applied electrochemistry. 28: 899-905.


Amit S., Rupali G. (2004). Developments in wastewater treatment methods.

Desalination. 167: 55-63.


Amor L., Eiroa M., Kennes C. and Veiga M.C. (2005). Phenol biodegradation and its effect on the nitrification process. Water Res. 39: 2915-2920. Amvrossios C. B., Nada A., Rene C. (2006). Effect of chaotic mixing on enhanced biological growth and implications for wastewater treatment: A test case with Saccharomyces cerevisiae. Journal of Hazardous Materials. 136:130-136. Andreescu S., Andreescu D. and Sadik O.A. (2003). A new electrocatalytic mechanism for the oxidation of phenols at platinum electrodes. Electrochem. 5: 681-688. Arana J., Rendon E.T., Rodriguez J.M.D., Melian J.A., Diaz O. and Pena J. (2001). Highly concentrated phenolic wastewater treatment by the photo-fenton reaction, mechanism study by FTIR-ATR. Chemosphere. 44(5): 1017-1023. Arikawa T., Murakami Y., Takasu Y. (1998). Simulatneous determination of chlorine and oxygen evolving at RuO2/Ti and RuO2/Ti anodes by differential electrochemical mass spectroscopy. Journal of Applied Electrochemistry. 28: 511-516. Arnold E., Boehm B. and Wilderer P.A. (2000). Application of activated sludge and biofilm sequencing batch reactor technology to treat reject water from sludge dewatering systems: a comparison. Water Science and Technology. 41(1): 115-122.

140

Reference Azbar N., Yonar T. and Kestioglu K. (2004). Comparison of various advanced oxidation processes and chemical treatment methods for COD and color removal from a polyester and acetate fiber dyeing effluent. Chemosphere. 55 (1): 35-43. Bauer, R. (1994). Applicability of solar irradiation for photochemical wastewater treatment. Chemosphere. 29: 1225. Beline F., Martinez J., Marol C. and Guiraud G. (2001). Application of the
15

technique to determine the contributions of nitrification and denitrification to the flux of nitrous oxide from aerated pig slurry. Water Res. 35: 2774-2778. Bhatkhande, D.S., Pangarkar, V.G., Beenackers, A.A.C.M. (2002). Photocatalytic degradation for environmental applications: a review. J. Chem. Technol.

Biotechnol. 77 (1): 102.


Bidga R.J., (1995). Consider fenton chemistry for waste-water treatment. Chem Eng

Prog. 91(12): 62-66.


Breivik K., Alcock R., Li Y.F., Bailey R.E., Fiedler H., Pacyna J.M. (2004). Primary sources of selected POPs: regional and global scale emission inventories,

Environ. Pollution. 128: 3-16.


Bonfatti F., Ferro S., Levezzo F., Malacarne M., Lodi G. and Battisti A. D. (1999). Electrochemical incineration of glucose as a model organic substrate. I. Role of the electrode material, J. Electrochem. Soc. 146: 2175. Brillas E., Cabot P.L., Garrido J., Montilla M., Rodrguez R.M. and Carrasco J. (1997). Faradaic impedance behaviour of oxidized and reduced poly (2,5-di(2-thienyl)-thiophene) films. Journal of Electroanalytical Chemistry. 430(1-2): 133-140. Brillas E., Montilla M., Carrasco J. and Otero T. F. (1995). Electrochemical behaviour of poly(2,5-di-(-2-thienyl) -thiophene): properties of the reduced polymer. Journal of Electroanalytical Chemistry. 418(1-2): 123-130. Brillas E., Mur E., Sauleda R., Sanchez L., Peral J., Domenech X. and Casado J. (1998). Aniline ineralization by AOP's: anodic oxidation, photocatalysis, electro-Fenton and photoelectro-Fenton processes. J. Appl. Catal. B: Environ.
16: 31-42.

Camel V. and Bermond A. (1998). Review Paper: The use of ozone and associated oxidation processes in drinking water treatment. Water Res. 32: 3208-3222.

141

Reference Cao G.M., Zhao Q.X., Sun X.B. and Zhang T. (2003). Integrated nitrogen removal in a shell-and-tube co-immobilized cell bioreactor. Process. Biochem. 39:1 2691273. Zanta C.L.P.S., Michaud P.A., Comninellis C., Andrade A.R.D; Boodts J.F.C. (2003). Electrochemical oxidation of p-chlorophenol on SnO2Sb2O5 based anodes for wastewater treatment. Journal of Applied Electrochemistry. 33: 1211-1215. Carrera J., Baeza J.A., Vicent T. and Lafuente J., Biological nitrogen removal of highstrength ammonium industrial wastewater with two-sludge system. Water Res.
37(2003): 4211-4221.

Chamarro E., Marco A. and Esplugas S. (2001). Use of fenton reagent to improve organic chemical biodegradability. Water Res. 35(4): 1047-1051. Chen G.H. (2004). Electrochemical technologies in wastewater treatment. Separation

and Purification Technology. 38: 11-41.


Chen G.H., Chen X.M., Yue P.L. (2002). Electrochemical behavior of novel Ti/IrOxSb2O5-SnO2 anodes. Journal of Physical Chemical B. 106: 4364-4369. Chiang L.C., Chang J.E., and Wen T.C. (1995). Indirect oxidation effect in electrochemical oxidation treatment of landfill leachate, Wat. Res. 29(2): 671678. Choisnet J.; Bizo L.; Retoux R.; Raveau B. (2004). Antimony and antimonytin doped indium oxide, IAO and IATO: promising transparent conductors Solid

state sciences. 6: 1121-1123.


Comninellis C. and Pulgarin C. (1991). Anodic-oxidation of phenol for waste-water treatment. Journal of Applied Electrochemistry. 21: 703-708. Comninellis C. and Pulgarin C. (1993). Electrochemical oxidation of phenol for wastewater treatment using SnO2, anodes.

Journal

of

Applied

Electrochemistry. 23: 108-112.


Comninellis C. (1994). Electrocatalysis in the electrochemical conversion/combustion of organic pollutants for waste-water treatment. Electrochim. Acta. 39: 18571862. Cornell A., Hkansson B. and Lindbergh G. (2003). Ruthenium based DSA in chlorate electrolysiscritical anode potential and reaction kinetics. Electrochimica Acta. 48(5): 473-481. Correa-Lozano B., Comninellis Ch., De Battisti A. (1997). Service life of Ti/SnO2Sb2O5 anodes. Journal of Applied Electrochemistry. 27: 970-974. 142

Reference Cossu R., Raga R. and Rossetti D. (2003). The PAF model: an integrated approach for landfill sustainability. Waste Management. 23(1): 37-44. Czapla A., Kusior E. and Bucko M. (1989). Optical properties of non-stoichiometric tin oxide films obtained by reactive sputtering. Thin Solid Films, Volume.
182(1-2, 20): 15-22.

Devlin, H. and Harris, 1. (1984). Mechanism of the Oxidation of Aqueous Phenol with Dissolved Oxygen. Ind Eng. Chem. Fundam. 23(4): 387-392. Diao H.F., Li X.Y., Gu J.D., Shi H.C. and Xie Z.M. (2004). Electron microscopic investigation of the bactericidal action of electrochemical disinfection in comparison with chlorination, ozonation and Fenton reaction. Process

Biochemistry. 39: 1421-1426.


Duprez D., Delano F., Barbier J., Jr , Isnard P. and Blanchard G. (1996). Catalytic oxidation of organic compounds in aqueous media. Catalysis Today, 29(1-4): 317-322. El-Geundi M.S. (1991). Colour removal from textile effluents by adsorption techniques. Water Res. 25: 271-273. Fallmann H., Krutzler T., Bauer R., Malato S. and Blanco J. (1999). Applicability of the photo-fenton method for treating water containing pesticides. Catal Today.
54(2-3): 309-319.

Farre M., Barcelo D. (2003). Toxicity testing of wastewater and sewage sludge by biosensors, bioassays and chemical analysis. TRAC Trends Anal Chem.
22(5): 299-310.

Faust B.C. and Zepp R.G., (1993). Photochemistry of aqueous iron (III) polycarboxylate complexes: roles in the chemistry of atmospheric and surface waters. Environ Sci Technol. 27: 2517-2522. Feng C., Sugiura N., Shimada S. and Maekawa T. (2003). Development of a high performance electrochemical wastewater treatment system. Journal of

Hazardous Materials. 103(1-2): 65-78.


Feng J.R. and Johnson D.C. (1991). Electrocatalysis of anodic oxygen-transfer reactions: titanium substrates for pure and doped lead dioxide films. J.

Electrochem. Soc. 138: 33283337.


Feng J.R., Houk L.L., Johnson D.C., Lowery S.N. and Carey J.J. (1995). Electrocatalysis of anodic oxygen-transfer reactions: the electrochemical incineration of benzoquinone. J. Electrochem. Soc. 142: 3626-3632. 143

Reference Feng, Y. J., Li, X. Y. (2003). Electro-catalytic oxidation of phenol on several metaloxide electrodes in aqueous solution. Water Res. 37: 2399-2407. Gattrell M. and Kirk D.W. (1993). A study of the oxidation of phenol on platinum and preoxidised platinum surfaces. J. Electrochem. Soc. 140: 1534-1540. Gerger I. and Haubner R. (2005). Cyclic voltammetry measurements on boron- and nitrogen-doped diamond layers. Diamond and Related Materials. 14(3-7): 369-374. Gogate P.R. and Pandit A.B. (2004). A review of imperative technologies for wastewater treatment I: oxidation technologies at ambient conditions. Advances in Environmental Research. 8(3-4): 501-551. Gomes H.T., Samant P.V., Serp Ph., Kalck Ph., Figueiredo J.L. and Faria J.L. (2004). Carbon nanotubes and xerogels as supports of well-dispersed Pt catalysts for environmental applications. Applied Catalysis B: Environmental. 54(3): 175182. Grahl T. and Markl H. (1996). Killing of microorganisms by pulse electric fields.

Applied Microbiology and Biotechnology. 45: 148-157.


Greet L., Jan P., Marc V., Luc P. and Willy V. (2003). Electrochemical degradation of surfactants by intermediates of water discharge at carbon-based electrodes.

Electrochimica. Acta. 48: 1656.


Grimm J.H.; Bessarabov D.G.; Simon U.; Sanderson R.D. (2000). Characterization of dopted tin dioxide anodes prepared by a sol-gel technique and their application in an SPE-reaactor. Journal of Applied Electrochemistry. 30: 293-302. Gu L., Wang B., Ma H. and Kong W. (2006). Catalytic oxidation of anionic surfactants by electrochemical oxidation with CuOCo2O3PO43 modified kaolin. Journal of Hazardous Materials. (April) Habazaki H., Hayashi Y. and Konno H. (2002). Characterization of electrodeposited WO3 films and its application to electrochemical wastewater treatment. Electrochimica Acta, 26: 4181-4188. Hostachy J.C., Lenon G., Pisicchio J.L., Coste G. and Legay C. (1997). Reduction of pulp and paper mill pollution by ozone treatment. Water Sci. Technol. 35: 261268. Houk L.L., Johnson S.K., Feng J.R., Houk R.S. and Johnson D.C. (1998). Electrochemical incineration of benzoquinone in aqueous media using a

144

Reference quaternary metal oxide electrode in the absence of a soluble supporting electrolyte. J. Appl. Electrochem. 28(1998): 1167-1177. Huber M.M., Canonica S., Park G.-Y. and Gunten U. (2003). Oxidation of pharmaceuticals during ozonation and advance oxidation processes. Environ.

Sci. Technol. 37: 1016-1024.


Huber M.M., Ternes T.A. and Gunten U. (2004). Removal of estrogenic activity and formation of oxidation products during ozonation of 17-ethinylestradiol.

Environ. Sci. Technol. 38: 5177-5186.


Huber M.M., Gbel A., Joss A., Hermann M., Lffler D., McArdell C.S., Ried A., Siegrist H., Ternes T.A. and Gunten U. (2005). Oxidation of pharmaceuticals during ozonation of municipal waste water effluents: a pilot study. Environ.

Sci. Technol. 39: 4290-4299.


Hu J.Y., Wang S., Ng W.J. and Ong S.L. (1999). The effect of water treatment processes on the biological stability of potable water. Water Res. 33: 25872592. Iniesta J., Michaud P.A., Panizza M., Cerisola G., Aldaz A. and Comninellis Ch. (2001). Electrochemical oxidation of phenol at boron-doped diamond electrode. Electrochim. Acta. 46: 3573-3578. Itokawa H., Hanaki K. and Matsuo T. (2001). Nitrous oxide production in highloading biological nitrogen removal process under low COD/N ratio condition.

Water Res. 35: 657-664.


Janos P., Buchtova H. and Ryznarova M. (2003). Sorption of dyes from aqueous solutions onto fly ash. Water Res. 37: 4938-4944. Jetten M.S.M., Horn S.J. and van Loosdrecht M.C.M. (1997). Towards a more sustainable municipal wastewater treatment system. Water Science and

Technology. 35(9): 171-180.


John, C. and Robert, L. (1985). Nitrogen removal in a low-loaded single tank sequencing batch reactor. J. Water Pollut. Control. Fed. 57 (1): 82-86. Johnson S.K., Houk L.L., Feng J.R., Houk R.S. and Johnson D.C. (1999). Electrochemical incineration of 4-chlorophenol and the identification of products and intermediates by mass spectrometry. Environ. Sci. Technol. 33: 2638-2644.

145

Reference Joo Hung-Soo, Hirai Mitsuyo and Shoda Makoto. (2006). Piggery wastewater treatment using Alcaligenes faecalis strain No. 4 with heterotrophic nitrification and aerobic denitrification. Water Research. Jung J.Y., Chung Y.C., Shin H.S. and Son D.H. (2004). Enhanced ammonia nitrogen removal using consistent biological regeneration and ammonium exchange of zeolite in modified SBR process. Water Res. 38: 347-354. Kadam M. R., Vittal N., Karekar R. N. and Aiyer R. C. (1990). Electrical characteristics of SnO2-based thick film resistors loaded with SnCl2 . Thin

Solid Films. 187(2): 199-208.


Kajitvichyanukul P., Lu M.C., Liao C.H., Wirojanagud W. and Koottatep T. (2006). Degradation and detoxification of formaline wastewater by advanced oxidation processes. Journal of Hazardous Materials. 135(1-3): 337-343. Karvelas M., Katsoyiannis A., Samara C. (2003). Occurrence and fate of heavy metals in the wastewater treatment process. Chemosphere. 53: 1201-1210. Katsoyiannis A., Samara C. (2004). Persistent Organic Pollutants (POPs) in the conventional activated sludge treatment process: Occurrence and removal.

Water Research. 38(11): 2685-2698.


Khelifa A., Moulay S. and Naceur A.W. (2005). Treatment of metal finishing effluents by the electroflotation technique. Desalination. 181(1-3): 27-33. Kim J.H., Chen M., Kislida N., Sudo R. (2004). Integrated real-time control strategy for nitrogen removal in swine wastewater treatment using sequencing batch reactors. Water Research. 38: (14-15) 3340-3348. Kim K.W., Kuppuswamy M. and Savinell R.F. (1990). Electrochemical oxidation of benzene at a glassy carbon electrode. J. Appl. Chem. 30: 543. Kolisch G. and Rolfs T. (2000). Integrated sidestream treatment for enhanced enlargement of sewage plants. Water Science and Technology. 41(9): 155-162. Krsi L., Papp S., Meynen V., Cool P., Vansant E.F. and Dkny I. (2005). Preparation and characterization of SnO2 nanoparticles of enhanced thermal stability: The effect of phosphoric acid treatment on SnO2nH2O. Colloids and

Surfaces A: Physicochemical and Engineering Aspects. 268(1-3): 147-154.


Ktz, R., Stucki, S., Carcer, B. (1991). Electrochemical waste-water treatment using high overvoltage anodes I: physical and electrochemical properties of SnO2. J.

Appl. Electrochem. 21: 14-20.

146

Reference Kositzi M., Poulios I., Malato S., Caceres J. and Campos A. (2004). Solar photocatalytic treatment of synthetic municipal wastewater. Water Res. 38(5): 1147-1154. Krutzler T. and Bauer R. (1999). Optimization of a photo-fenton prototype reactor.

Chemosphere. 38(11): 2517-2532.


Krysa J. and Mraz R. (1994). Experimental investigation of the double layer capacity, X-ray diffraction and the relative surface content of TiH2 during pretreatment of titanium used for the preparation of dimensionally stable anodes with RuO2 and/or IrO2 coating. Electrochimica Acta. 40(12): 1997-2003. Lambert S.D. and Graham N.J.D. (1995). Removal of non-specific dissolved organic matter from upland potable water suppliesII. Ozonation and adsorption.

Water Res. 29(10): 2427-2433.


Lau W.C.I. Removal of Refractory Chemical Landfill Leachate by UASB and Advanced Oxidation Processes. (PhD thesis, The University of Hong Kong. 2000) Lee B.D. and Hosomi M. (2001). Fenton oxidation of ethanol-washed distillationconcentrated benzo(a)pyrene: reaction product identification and biodegradability. Water Res. 35(9): 2314-2319. Lee S.I., Koopman B., Park S.K. and Cadee K. (1995). Effect of fermented waster on denitrification in activated sludge. Water Environ. Res. 67(7): 1119-1122. Lee S.I., Parkm J.H., Ko K.B. and Koopman B. (1997). Effect of fermented swine waster on biological nutrient removal in sequencing batch reactors. Water Sci.

Tech. 31(7): 1807-1812.


Levec J. and Pintar A. (1995). Catalytic oxidation of aqueous solutions of organics. An effective method for removal of toxic pollutants from waste waters .

Catalysis Today. 24(1-2): 51-58.


Li Puma, G. and Yue, P.L. (1998a). A laminar falling film slurry photocatalytic reactor part Imodel development. Chem. Eng. Sci. 53: 2993. Li Puma, G. and Yue, P.L. (1998b). A laminar falling film slurry photocatalytic reactor part IIexperimental validation of model. Chem. Eng. Sci. 53: 3007. Li X.Y., Ding F., Lo P.S.Y. and Sin S.H.P. (2002). Electrochemical disinfection of saline wastewater effluent. J Environmental Engineering, ASCE. 128: 697-704.

147

Reference Li X.Y., Cui Y.H., Feng Y.J., Xie Z.M., Gu J.D. (2005). Reaction pathways and mechanisms of the electrochemical degradation of phenol on different electrodes. Water Research. 39: 1972-1981. Lin S. H. and Ho S. J. (1996). Catalytic wet-air oxidation of high strength industrial wastewater. Applied Catalysis B: Environmental. 9(1-4): 133-147. Lin S.H., Shyu C.T. and Sun M.C. (1998). Saline wastewater treatment by electrochemical method. Water Research. 32(4): 1059-1066. Lin S. H. and Wu C. L. (1996). Electrochemical removal of nitrite and ammonia for aquaculture Water Research, 30(3): 715-721. Lipp L. and Pletcher D. (1997). The preparation and characterization of tin dioxidecoated titanium electrodes. Electrochimina Acta. 42: 1091-1099. Liu Y.-Q., Tay J.-H., Ivanov V., Moy B.Y.-P., Yu L. and Tay S.T.-L. (2005). Influence of phenol on nitrification by microbial granules. Process Biochem.
40: 3285-3289

Lund H. and Baizer M. Eds. 1991. 3rd Edition. Practical Problem in Electrolysis.

Organic Electrochemistry, An lntroductionand Guide. Marcel Dekker: New


York. Maier J. (1993). Electrical sensing of complex gaseous species by making use of acidbase properties. Solid State Ionics. July 1993, 62(1-2): 105-111. Manzanares M. I., Pavese A.G. and Solis V. M. (1991). Comparative investigation of formic acid and formaldehyde electro-oxidation on palladium in acidic medium : Effect of surface oxides. Journal of Electroanalytical Chemistry.
310(1-2): 159-167.

Manzanares M. I., Sols V. and Rita H. R. (1996). Effect of cyclodextrins on the electrochemical behaviour of ascorbic acid on gold electrodes. Journal of

Electroanalytical Chemistry. 407(1-2): 141-147.


Maugans C.B. and Akgerman A. (1997). Catalytic wet oxidation of phenol over a Pt/TiO catalyst. Water Research. 31(12): 3116-3124. Mengoli G. and Musiani M.M., (1986). Protective coatings on iron by anodicoxidation of phenols in oxalic-acid medium. Electrochim. Acta. 31: 201-210. Milstein O., Haars A., Krause F. and Huettermann A. (1991). Decrease of pollutant level of bleaching effluents and winning valuable products by successive flocculation and microbial growth. Water Sci. Technol. 24: 199-206.

148

Reference Mobius C.H. and Helble A. (2004). Combined ozonation and biofilm treatment for reuse of paper mill wastewaters. Water Sci. Technol. 49: 319-323. Moiseev A., Schroeder H., Kotsaridou-Nagel M., Geissen S.U. and Vogelpohl A. (2004). Photocatalytical polishing of paper-mill effluents. Water Sci. Technol.
49: 325-330.

Mukherjee, P.S., Ray, A.K. (1999). Major challenges in the design of a large scale photocatalytic reactor for water treatment. Chem. Eng. Technol. 22(3): 253. Oliveros E., Legrini O., Hohl M., Mueller T. and Braun A. (1997). Industrial waste water treatment: large scale development of a light-enhanced fenton reaction.

Chem Eng Process. 36(5): 397-405.


Panizza M., delucchi M. and Cerisola G. (2005). Electrochemical degradation of anionic surfactants, J. Appl. Electrochem. 35: 358. Park J.C., Lee M.S., Lee D.H., Park B.J., Han D.W., Uzawa M. and Takatori K. (2003). Inactivation of bacteria in seawater by low-amperage electric current.

Applied and Environmental Microbiology. 69: 2405-2408.


Patermaraxis G. and Fountoukidis E. (1990). Disinfection of water by electrochemical treatment. Water Res. 24: 1491-1496. Peir A.M., Aylln J.A., Peral J. and Domnech X. (2001). TiO2-photocatalyzed degradation of phenol and ortho-substituted phenolic compounds. Appl. Catal.

B: Environ. 30: 359-373.


Pea-Alvarez A., Daz L., Medina A., Labastida C., Capella S. and Vera L.E. (2003). Characterization of three Agave species by gas chromatography and solidphase microextractiongas chromatographymass spectrometry.

J.

Chromatogr. A 1027: 131-136.


Petalaa M., Tsiridisa V., Samarasb P., Zouboulisc A., Sakellaropoulosa G.P. (2006). Wastewater reclamation by advanced treatment of secondary effluents.

Desalination. 195: 109-118.


Pham T., Proulx S. (1997). PCBs and PAHs in the Montreal urban community (Quebec, Canada) wastewater treatment plant and in the effluent plume in the St. Lawrence river. Water Research. 31(8): 1887-1896. Pilla A.S., Cobo E.O., Duarte M.M.E., Salinas D.R. (1997). Evaluation of anode deactivation in chlor-alkali cell. Journal of Applied Electrochemistry. 27: 1283-1289.

149

Reference Pintar A. and Levec J. (1994). Catalytic oxidation of aqueous p-chlorophenol and pnitrophenol solutions. Chemical Engineering Science. 49(24): 4391-4407. Pintar A., Besson M. and Gallezot P. (2001). Catalytic wet air oxidation of Kraft bleach plant effluents in a trickle-bed reactor over a Ru/TiO2 catalyst. Applied

Catalysis B: Environmental. 31(4): 275-290.


Plummer J.D. and Edzwald J.K. (1998). Effect of ozone on disinfection by-product formation of algae. Water Sci. Technol. 37: 49. Polcaro A.M. and Palmas S. (1997). Electrochemical oxidation of chlorophenols, Ind.

Eng. Chem. Res. 36: 1791.


Polcaro, A. M.; Palmas, S.; Renoldi, F.; Mascia, M. (1999). On the performance of Ti/SnO2 and Ti/PbO2 anodes in electrochemical degradation of 2-chlorophenol for wastewater treatment. Journal of Applied Electrochemistry. 29: 147-151. Polcaro A.M.; Palmas S.; Renoldi F.; Mascia M. (2000). Three-demensional electrodes foe the electrochemical combustion of organic pollutants.

Electrochimica Acta. 46: 389-394.


Popova L.I., Michailov M.G., Gueorguiev V.K. and Shopov A. (1990). Structure and morphology of thin SnO2 films. Thin Solid Films. 186(1): 107-112. Ra C.S., Lo K.V., Shin J.S., Oh J.S. and Hong B.J. (2000). Biological nutrient removal with an internal organic carbon source in piggery wastewater treatment. Water Res. 34: 965-973. Rajeshwar K., Ibanez J.G. and Swain G.M. (1994). Electrochemistry and the environment, J. Appl. Electrochem. 24: 10771091. Ray, A.K. (1999). Design, modeling and experimentation of a new large-scale photocatalytic reactor for water treatment. Chem.Eng. Sci. 54: 3113. Reemtsma T., Gnirss R. and Jekel M. (2000). Infiltration of combined sewer overflow and tertiary municipal wastewater: an integrated laboratory and field study on nutrients and dissolved organics. Water Research. 34: 1179-1186. Rieger P.H., Electrochemistry, Prentice-Hall, Englewood Cliffs, NJ (1987) Rodgers J.D., Jedral W.J. and Bunce N.J., (1999). Electrochemical oxidation of chlorinated phenols. Environ. Sci. Technol. 33: 1453-1457. Rodgers J.D. Electrochemical Treatment of Recalcitrant Waste: A Study of Chlorophenols and Nitroaromatic Compounds. (PhD thesis, The University of Guelph, 2000).

150

Reference Rodrigues E.C.P.E. and Olivi P. (2003). Preparation and characterization of Sb-doped SnO2 films with controlled stoichiometry from polymeric precursors. Journal

of physics and chemistry of solids. 64: 1105-1112.


Rostron W.M., Stuckey D.C. and Young A.A. (2001). Nitrification of high strength ammonia wastewaters: comparative study of immobilization media. Water Res.
35: 1169-1178.

Rossi A., Boodts J.F.C. (2002). Ir based oxide electrodes: oxygen evolution reaction from mixed solvents. Journal of Applied Electrochemistry. 32: 735-741. Savant D.V., Abdul-Rahman R. and Ranade D.R. (2006). Anaerobic degradation of adsorbable organic halides (AOX) from pulp and paper industry wastewater.

Bioresource Technology. 97(9): 1092-1104.


Schumann U. and Grundler P. (1998). Electrochemical degradation of organic substrates at PbO2 anodes: monitoring by continuous CO2 measurements.

Water Res. 32: 2835-2842.


Sedlak, D. and Andren, A. (1991b). Aqueous-Phase Oxidation of Polychlorinated Biphenyls by Hydroxyl Radicals. Environ Sci. Technol. 25(8): 1419-1427. Shen Y.S., Ku Y. and Lee K.C. (1995). The effect of light absorbance on the decomposition of chlorophenols by UV radiation and UV/H2O2. Water Res. 29: 907-914. Simond O.and Comninellis Ch. (1997). Anodic oxidation of organics on Ti/IrO2 anodes using Nafion as electrolyte. Electrochimica Acta. 42: 2013-2018. Simond O., Schaller V. and Comninellis Ch. (1997). Theoretical model for the anodic oxidation of organics on metal oxide electrodes. Electrochimica Acta. 42: 2009-2012. Stoner G.E. and Cahen G.L. (1982). The mechanism of low frequency AC electrochemical disinfection. Bioelectrochemistry and Bioengineering. 9: 229243. Stucki F., Erbudak M. and Kostorz G. (1987). The initial oxidation of solid and liquid aluminium. Applied Surface Science. 27(4): 393-400. Stucki S., Ktz R., Carcer B. and Suter W. (1991). Electrochemical waste-water treatment using high overvoltage anodes II: anode performance and application. J. Appl. Electrochem. 21: 99-104. Su J.J., Liu B.Y. and Liu C.Y. (2001). Comparison of aerobic denitrification under high oxygen atmosphere by Thiosphaera pantotropha ATCC 33512 and 151

Reference

Pseudomonas stutzeri SU2 newly isolated from the activated sludge of a


piggery waste water treatment system. J. Appl. Microbiol. 90: 457-462. Tahar N.B. and Savall A. (1999). Electrochemical degradation of phenol in aqueous solution on bismuth doped lead dioxide: a comparison of various electrode formulations. J. Appl. Electrochem. 29: 277-283. Tanaka S., Nakata Y., Kimura T., Yustiawati, Kawasaki M. and Kuramitz H. (2002). Electrochemical decomposition of bisphenol A using Pt/Ti and SnO2/Ti anodes. J. Appl. Electrochem. 32: 197-201. Ternes T.A., Stber J., Hermann N., McDowell D., Ried A., Kampmann M. and Teiser B. (2003). Ozonation: a tool for removal of pharmaceuticals, contrast media and musk fragrancies from wastewater. Water Res. 37: 1976-1982. Thornton G. (1992). Poisons and promoters on TiO2 surfaces. Vacuum. 43(11): 11331135. Trasatti S. (2000). Electrocatalysis: understanding the success of DSA.

Electrochimica Acta. 45(15-16): 2377-2385.


Ulrike D. (2003). The surface science of titanium dioxide. Surface Science Reports.
48(5-8): 53-229.

Uygur A. and Kargi F. (2004). Biological nutrient removal from pre-treated landfill leachate in a sequencing batch reactor. J. Environ. Manage. 71: 9-14. Vaidya P.D. and Mahajani V.V. (2002). Insight into heterogeneous catalytic wet oxidation of phenol over a Ru/TiO2 catalyst. Chemical Engineering Journal.
87(3): 403-416.

Valds Hctor and Zaror C.A. (2006). Heterogeneous and homogeneous catalytic ozonation of benzothiazole promoted by activated carbon: Kinetic approach.

Chemosphere.
Vanlangendonck Y., Corbisier D. and Lierde A. V. (2005). Influence of operating conditions on the ammonia electro-oxidation rate in wastewaters from power plants (ELONITA technique) Water Res.. 39(13): 3028-3034. Vlyssides A. G., Karlis P. K., Rori N. and Zorpas A. A. (2006). Electrochemical treatment in relation to pH of domestic wastewater using Ti/Pt electrodes.

Journal of Hazardous Materials.


Vlyssides A. G., Loizidous M., Karlis P.K., Zorpas A.A. and Papaioannou D. (1999). Electrochemical oxidation of a textile dye wastewater using a Pt/Ti electrode,

J. Hazard. Mater. B. 70: 41.


152

Reference Wang Y.H., Chan K.Y., Li X.Y. and So S.K. (2006). Electrochemical degradation of 4-chlorophenol at nickelantimony doped tin oxide ectrode. Chemosphere. Yao W.X., Kennedy K.J., Tam C.M. and Hazlett J.D. (1994). Pretreatment of kraft pulp bleach plant effluent by selected ultrafiltration membranes. Canad. J.

Chem. Eng. 72: 991-999.


Murakami Y.; Kond T.; Shimoda Y.; Kaji H.; Zhang, X.G.;Yoshio Takasu. (1997). Porous ruthenium oxide electrode prepared by adding lanthanum chloride to the coating solution. Journal of Alloys and Compounds. 261: 176-181. Yawalkar, A.A., Bhatkhande, D.S., Pangarkar, V.G., Beenackers, A.A.C.M. (2001). Solar assisted photochemical and photocatalytic degradation of phenol. J.

Chem.T ech. Biotech. 76(4): 363.


Yonar Taner, Kestioglu Kadir and Azbar Nuri. (2006). Treatability studies on domestic wastewater using UV/H2O2 process. Applied Catalysis B:

Environmental.
Kim Y.M., Park D., Lee D. S. and Park J. M. (2006). Instability of biological nitrogen removal in a cokes wastewater treatment facility during summer. Journal of

Hazardous Materials.
Zanta C.L.P.S., Michaud P.A; Comninellis C. (2003). Electrichemical oxidation of pchlorophenol on SnO2-Sb2O5 based anodes for wastewater treatment. Journal

of applied electrochemistry. 33: 1211-1215.


Zhang J., Giorno L., Driolib E. (2006). Study of a hybrid process combining PACs and membraneoperations for antibiotic wastewater treatment. Desalination.
194: 101-107.

153

Anda mungkin juga menyukai