04.11.2014 Views

Universidad del Turabo Heterogeneous Catalysis Applied To ...

Universidad del Turabo Heterogeneous Catalysis Applied To ...

Universidad del Turabo Heterogeneous Catalysis Applied To ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>Universidad</strong> <strong>del</strong> <strong>Turabo</strong><br />

<strong>Heterogeneous</strong> <strong>Catalysis</strong> <strong>Applied</strong> <strong>To</strong> Advanced Oxidation Processes (AOPs) For<br />

Degradation of Organic Pollutants<br />

By<br />

María <strong>del</strong> Carmen Cotto-Maldonado<br />

BS, Biology, University of Puerto Rico<br />

BS, Chemistry, Interamerican University<br />

MS, Environmental Health, University of Puerto Rico<br />

Dissertation<br />

Submitted to the School of Science and Technology<br />

in partial fulfillment of the requirements for<br />

the degree of Doctor of Philosophy<br />

in Environmental Science<br />

(Chemistry Option)<br />

Gurabo, Puerto Rico<br />

May, 2012


<strong>Universidad</strong> <strong>del</strong> <strong>Turabo</strong><br />

A dissertation submitted in partial fulfillment of<br />

the requirement for the degree of<br />

Doctor of Philosophy<br />

4/24/2012<br />

<strong>Heterogeneous</strong> <strong>Catalysis</strong> <strong>Applied</strong> <strong>To</strong> Advanced Oxidation Processes (AOPs) For<br />

Degradation of Organic Pollutants<br />

Maria <strong>del</strong> Carmen Cotto-Maldonado<br />

Approved:<br />

________________________________<br />

Francisco M Marquez Linares, PhD<br />

Research Advisor<br />

______________________________<br />

Marlio Paredes, PhD<br />

Member<br />

________________________________<br />

Jose J Duconge, PhD<br />

Member<br />

______________________________<br />

Angel L Morales Cruz, PhD<br />

Member<br />

________________________________<br />

Santander Nieto, PhD<br />

Member<br />

________________________________<br />

Fred C Schaffner, PhD<br />

Associate Dean, Graduates Studies<br />

and Research<br />

______________________________<br />

Teresa Lipsett, PhD<br />

Dean<br />

ii


© Copyright 2012<br />

María <strong>del</strong> Carmen Cotto-Maldonado. All Right Reserved.


Dedications<br />

<strong>To</strong> my angels in Heaven and Earth…<br />

<strong>To</strong> my Family<br />

iii


Acknowledgments<br />

One day a friend said to me that sometimes it is necessary to touch the thorn of<br />

the rose to reach the flower and I wish to say thanks to all of those people that helped<br />

and supported me during this journey.<br />

I wish to say thanks to Dr Francisco M Marquez-Linares for your mentoring. I<br />

read in some place that a mentor is someone that not only helps the student to direct the<br />

investigation project, but is part of it. Thanks for your patience, support and help, to<br />

teach me what is a good professor and a human being, and as it is said in Puerto Rico:<br />

“pasar conmigo la zarza y el guayacán”. It has been an honor being your student.<br />

Thanks to the members of my dissertation committee; Dr Jose J Duconge, Dr<br />

Santander Nieto Ramos, Dr Marlio Paredes and Dr Angel L Morales Cruz for your<br />

support and trust during this process.<br />

Thanks to Dr Fred Schaffner, Associated Dean for the Graduate Studies and<br />

Research at the <strong>Universidad</strong> <strong>del</strong> <strong>Turabo</strong> for your guide and advice during all of this time.<br />

Many people collaborated in different forms to develop this research. Thanks to<br />

Dr Carmen Morant, Dr Eduardo Elizalde and Ms Teresa Campo at the <strong>Universidad</strong><br />

Autónoma de Madrid for all of your collaboration. I really appreciated all of your help<br />

and support.<br />

<strong>To</strong> Dr Angel Rivera Collazo and Dr. Angel Ojeda at <strong>Universidad</strong> <strong>del</strong><br />

<strong>Turabo</strong>, for your time.<br />

<strong>To</strong> my friends Mr Abraham (Kike) E Garcia, Ms Carmen Bonilla Rivera, Ms<br />

Veronica Castro Simmons and Ms Gloria M Herrera for all of your help, thanks to share<br />

with me the best of you during the long hours of laboratory work. Finally, but no less<br />

important, thanks to my friend Ms Karlo Malave-Llamas for the “phone call” that initiated<br />

this journey.<br />

iv


Curriculum Vitae<br />

Maria <strong>del</strong> C Cotto-Maldonado<br />

Education<br />

2004-Present<br />

PhD in Environmental Science, University of <strong>Turabo</strong>, Gurabo,<br />

Puerto Rico.<br />

2002-2006 BS in Chemistry, Inter American, Metropolitan Campus, Río<br />

Piedras, Puerto Rico<br />

1994-1997 MS in Environmental Health, University of Puerto Rico, Medical<br />

Science Campus, Río Piedras, Puerto Rico<br />

1987-1992 BS in Biology, University of Puerto Rico, Río Piedras Campus,<br />

Río Piedras, Puerto Rico<br />

Academic Honors, Awards and Achievements<br />

2011 Scientific Authors Award, Vice-Chancellor Office of Academic Affairs,<br />

University of <strong>Turabo</strong>, Gurabo Campus, Gurabo, Puerto Rico<br />

2010 Minigrant Award, Associate Dean Office of Graduate Studies, School of<br />

Science and Technology, University of <strong>Turabo</strong>, Gurabo Campus, Gurabo,<br />

Puerto Rico<br />

2009 Scientific Authors Award, Vice-Chancellor Office of Academic Affairs,<br />

University of <strong>Turabo</strong>, Gurabo Campus, Gurabo, Puerto Rico<br />

Minigrant Awards, Associate Dean Office of Graduate Studies, School of<br />

Science and Technology, University of <strong>Turabo</strong>, Gurabo Campus, Gurabo,<br />

Puerto Rico<br />

v


2008 Minigrant Award, Dean Office of Graduate Studies, School of Science<br />

and Technology, University of <strong>Turabo</strong>, Gurabo Campus, Gurabo, Puerto<br />

Rico<br />

Scientific Authors Awards, Vice-Chancellor Office of Academic Affair,<br />

University of <strong>Turabo</strong>, Gurabo Campus, Gurabo, Puerto Rico<br />

2006 Graduate with Honors of the InterAmerican University, Metropolitan<br />

Campus<br />

2005 Founder Member of the Environmental Science Doctoral Student<br />

Association at University of <strong>Turabo</strong><br />

2003 Second Place in the Inorganic Advance Chemistry Competitions 2003<br />

(Olimpiadas de Química 2003), InterAmerican University, Metropolitan<br />

Campus<br />

1995 Founder Member of the Environmental Health Student Association at the<br />

University of Puerto Rico, Medical Science Campus<br />

1993 Graduate with Honors (Cum Laude) of the University of Puerto Rico, Río<br />

Piedras Campus<br />

Scientific Meetings<br />

12. Cotto M, Campo T, Elizalde E, Morant C, Marquez F. 2011. Hydrothermal<br />

Synthesis and Photocatalytic Activity of Titanium Oxide Nanowires Poster<br />

session presented at: 43rd IUPAC World Chemistry Congress, August 2011. San<br />

Juan PR.<br />

11. Cotto M, Masa A, Garcia A, Duconge J, Campo T, Elizalde E, Morant C,<br />

Márquez F.2011. ZnCd Based Photocatalysts for Hydrogen Production from<br />

Water under Visible-UV Light. Poster session presented at: 43rd IUPAC World<br />

Chemistry Congress, August 2011. San Juan PR.<br />

vi


10. Herrera GM, Campo T, Cotto M, Sanz JM, Elizalde E, Morant C, Marquez F.<br />

2011. Synthesis and Characterization of Hollow Magnetite Microspheres. Poster<br />

session presented at: 43rd IUPAC World Chemistry Congress, August 2011. San<br />

Juan PR.<br />

9. Herrera GM, Campo T, Cotto M, Sanz JM, Elizalde E, Morant C, Marquez F.<br />

2011. Preparation of Hollow Magnetite Microspheres and their Applications as<br />

Drugs Carriers. Poster session presented at: 43rd IUPAC World Chemistry<br />

Congress, August 2011. San Juan PR.<br />

8. Campo T, Marquez F, Cotto M, Elizalde E, Morant C. 2011. Silicon Nanowires<br />

grown from Silicon Substrates for Ion-Li Batteries Applications. Poster session<br />

presented at: 43rd IUPAC World Chemistry Congress, August 2011. San Juan<br />

PR.<br />

7. Duconge J, Bonilla C, Garcia A, Herrera GM, Cotto M, Campo T, Elizalde E,<br />

Morant C Marquez F. 2012. Synthesis and Characterization of Copper Oxide<br />

Nanowires. Poster session presented at: 43rd IUPAC World Chemistry<br />

Congress, August 2011. San Juan PR.<br />

6. Duconge J, Bonilla C, Cotto M, Herrera GM, Campo T, Elizalde E, Morant C,<br />

Marquez F. 2012. Hydrothermal Synthesis of Crystalline CuO Nanorods. Poster<br />

session presented at: 43rd IUPAC World Chemistry Congress, August 2011. San<br />

Juan PR.<br />

5. Herrera GM, Felix H, Campo T, Cotto M, Sanz JM, Elizalde E, Morant C,<br />

Hernández-Rivera S, Márquez F. 2012. Synthesis and characterization of Au<br />

coated TiO2 nanowires as ERS solid substrates. Poster session presented at:<br />

43rd IUPAC World Chemistry Congress, August 2011. San Juan PR.<br />

vii


4. Cotto-Maldonado MC, Roque-Malherbe R, Nieto S, Duconge J. 2008. Phenol<br />

Decomposition by Mechanical Activation of Rutile. Poster session presented at:<br />

28th Congreso Latinoamericano de Quimica. 2008. San Juan, Puerto Rico.<br />

3. Cotto MC, Malave K. 2003. Environmental Health Risk Communication for<br />

Hispanic Communities. Presented at: 2003 ATSDR Partners in Public Health<br />

Meeting. March 3-5 2003, Atlanta, Georgia.<br />

2. Cotto MC. Risk Communication. Presented at: University of Puerto Rico. [Rio<br />

Piedras (PR)]:University of Puerto Rico, Medical Science Campus.<br />

1. Marcantoni C, Cotto MC. 1998. Growth of Microbial Population in the<br />

Schmutzdecke of a Slow Sand Filter and its Relationship with Treated Water.<br />

Poster session presented at: AWWA Annual Conference June 21-25 1998,<br />

Dallas, Texas.<br />

Scientific Papers<br />

10. Cotto M, Duconge J, Campo T, Elizalde E, Morant C, Márquez F. Hydrothermal<br />

Synthesis and Catalytic Activity of TiO 2 nanowires. (<strong>To</strong> be submitted to J Catal).<br />

9. Marquez F, Masa A, Cotto M, Bonilla C, Garcia A, Duconge J, Campo T, Elizalde<br />

E, Morant C. Photocatalytic Hydrogen Production by Water Splitting using<br />

ZnCdFeS nanoparticles under UV-Vis Light Irradiation. (<strong>To</strong> be submitted to Int J<br />

Hydrogen Energy).<br />

8. Marquez F, Cotto M, Campo T, Elizalde E, Morant C. Photocatalytic degradation<br />

of Rhodamine B on different nanostructured catalyst. (<strong>To</strong> be submitted to Soft<br />

Nanoscience Lett).<br />

7. Marquez F, Cotto M, Bonilla C, Duconge J, Campo T, Elizalde E, Morant C. High<br />

Catalytic Activity of CuO Nanorods Synthesized by an Hydrothermal Approach.<br />

(Submitted to J Catal)<br />

viii


6. Marquez FM, Herrera GM, Campo T, Cotto MC, Duconge J, Sanz JM, Elizalde E,<br />

Perales O, Morant C. 2012. Preparation of hollow magnetite microspheres and<br />

their applications as drug carriers. Nanoscale Res Lett 7: 210.<br />

5. Marquez F, Campo T, Cotto M, Polanco R, Roque R, Fierro P, Sanz JM, Elizalde<br />

E, Morant C. 2011. Synthesis and Characterization of Monodisperse Magnetite<br />

Hollow Microsphere. Soft Nanoscience Lett: 25-32.<br />

4. Malave K, Cotto-Maldonado MC. 2010. Community environmental risk in<br />

developing countries. Environmental and Human Health: Risk Management in<br />

Developing Countries. Taylor and Francis Group.<br />

3. Cotto MC, Emiliano A, Nieto S, Duconge J, Roque-Malherbe R. 2009.<br />

Degradation of Phenol by Mechanical Activation of a Rutile Catalyst. J Colloid<br />

Interf Sci. 339: 133-139.<br />

2. Marcantoni C, Cotto MC. 1998. Growth of Microbial Population in the<br />

Schmutzdecke of a Slow Sand Filter and its Relationship with Treated Water.<br />

Proceedings of the AWWA Annual Conference. June 21-25 1998, Dallas<br />

(Texas). Vol. C, p. 751-776.<br />

1. Marcantoni C, Maldonado E, Cotto, MC. 1996. Determinación de las densidades<br />

poblacionales microbiológicas existentes en el schmutzdecke de un filtro de<br />

arena lento y su relación con la calidad <strong>del</strong> efluente [master’s thesis].[Rio Piedras<br />

(PR)]:University of Puerto Rico, Medical Science Campus.<br />

ix


Table of Contents<br />

Page<br />

List of Tables ................................................................................................................ xiv<br />

List of Figures ................................................................................................................ xv<br />

List of Appendices ........................................................................................................xxii<br />

Abstract ....................................................................................................................... xxiii<br />

Resumen .....................................................................................................................xxiv<br />

Chapter One. Introduction ............................................................................................... 1<br />

ChapterTwo. Experimental Techniques ....................................................................... 276<br />

2.01. X-Ray Diffraction (XRD) ................................................................................. 276<br />

2.02. Magnetic Susceptibility ..................................................................................... 29<br />

2.03. Thermogravimetric Analysis (TGA) .................................................................. 32<br />

2.04. Specific Surface Area (BET) ............................................................................ 34<br />

2.05. Raman Spectroscopy ....................................................................................... 36<br />

2.06. X-Ray Photoelectron Spectroscopy (XPS) ....................................................... 40<br />

2.07. Field Emission Scanning Microscopy (FE-SEM)............................................... 42<br />

2.08. <strong>To</strong>tal Organic Carbon Analysis ......................................................................... 45<br />

2.09. UV-Visible Spectroscopy .................................................................................. 46<br />

2.10. Fluorescence Spectroscopy .............................................................................. 49<br />

Chapter Three. Synthesis Procedures ........................................................................... 52<br />

3.1. Synthesis of Titanium Oxide Nanowires ............................................................. 52<br />

3.2. Synthesis of Zinc Oxide ..................................................................................... 53<br />

3.3. Synthesis of Titanium Oxide@Multiwalled Carbon Nanotubes ........................... 53<br />

3.3.1. Carbon Nanotubes Modification ................................................................... 53<br />

3.3.2. Synthesis and Incorporation of the Titanium Oxide on the MWCNT ............. 54<br />

3.4. Synthesis of Capped Magnetite Nanoparticles ................................................... 54<br />

3.5. Synthesis of Iron Oxide Nanowires .................................................................... 55<br />

Chapter Four.Material Characterization ..........................................................………….57<br />

4.1. Photocatalysis .................................................................................................... 57<br />

x


Page<br />

4.1.1. Titanium Oxide (TiO 2 , Rutile Phase) ............................................................ 57<br />

4.1.2. Titanium Oxide (TiO 2 , Anatase Phase) ........................................................ 61<br />

4.1.3. Titanium Oxide Nanowires ........................................................................... 65<br />

4.1.4. Titanium Oxide @Multiwalled Carbon Nanotubes ........................................ 68<br />

4.1.5. Zinc Oxide ................................................................................................... 73<br />

4.2. Fenton Catalysts ................................................................................................ 76<br />

4.2.1. Iron Oxide Nanowires (Fe 2 O 3 NWs) .............................................................. 76<br />

4.2.2. Capped Magnetite Nanoparticles (Fe 3 O 4 ) .................................................... 81<br />

4.2.3. Ferrous Chloride (FeCl 2 ).............................................................................. 86<br />

4.2.4. Copper Oxide (CuO).................................................................................... 88<br />

Chapter Five. Results and Discussion ........................................................................... 91<br />

5.1. Defining the Experimental Parameters ............................................................... 91<br />

5.1.1. Effects of the Concentration ......................................................................... 91<br />

5.1.2. Effects of the pH .......................................................................................... 94<br />

5.1.3. Effects of Temperature ................................................................................ 95<br />

5.2. Photochemical degradation ................................................................................ 96<br />

5.2.1. Description of the Photocatalytic System ................................................... 100<br />

5.3. Sono-Fenton Process ...................................................................................... 117<br />

5.3.1. Description of the Sono-Fenton System ..................................................... 117<br />

5.4. Photo-Fenton Process ..................................................................................... 124<br />

5.4.1. Description of the Photo-Fenton System .................................................... 124<br />

5.5. Statistical analysis ............................................................................................ 129<br />

Chapter Six. Conclusion .............................................................................................. 132<br />

Literature Cited ............................................................................................................ 134<br />

Appendix One. Dyes Solutions .................................................................................... 151<br />

Appendix Two. Photocatalytic Process ........................................................................ 152<br />

Appendix Three. Sono-Fenton Process ....................................................................... 182<br />

xi


Appendix Four. Photo-Fenton Process……………………………………………………..206<br />

xii


List of Tables<br />

Page<br />

Table 1.1. Characteristics of the most important dyes classes 19<br />

Table 5.1. Basic information of the studied organic compounds 103<br />

Table 5.2.<br />

Degradation percent of dye solutions during the<br />

Photocatalytic Process 107<br />

Table 5.3.<br />

Kinetic reaction rates and R2 values for the degradation<br />

reaction of the organic compounds during the<br />

photocatalytic process 109<br />

Table 5.4.<br />

Kinetic reaction rates and R2 values for de degradation<br />

reaction of the organic compounds during the sono-Fenton<br />

process 120<br />

Table 5.5.<br />

Kinetic reaction rates and R2 values for de degradation<br />

reaction of the organic compunds during the sono-Fenton<br />

process 122<br />

Table 5.6.<br />

Degradation percent of dye solutions during the Photo-<br />

Fenton Process 127<br />

Table 5.7.<br />

Kinetic reaction rates and R2 values for the degradation<br />

reaction of the organic compounds during the photo-<br />

Fenton process 129<br />

xiii


List of Figures<br />

Page<br />

Figure 1.01.<br />

Schematic diagram for the photoexcitation process in a<br />

semiconductor via photon irradiation 8<br />

Figure 1.02. Molecular structure of Methylene blue 20<br />

Figure 1.03. Molecular structure of Rhodamine B 21<br />

Figure 1.04. Molecular structure of the Methyl Orange 22<br />

Figure 1.05.<br />

Molecular structure of Crystal Violet and the molecular<br />

structure of Methyl Violet 23<br />

Figure 1.06. Molecular structure of p-amino benzoic acid (pABA) 24<br />

Figure 2.01.<br />

Schematic representation of the diffraction process from<br />

atoms in a crystalline lattice 28<br />

Figure 2.02. Images of the PANalytical XRD system 29<br />

Figure 2.03. Typical hysteresis loop of capped magnetite nanoparticles 30<br />

Figure 2.04. VSM components 31<br />

Figure 2.05. Lake Shore-7400 Vibrating Sample Magnetometer 32<br />

Figure 2.06. Schematic illustration of the TGA instrument 33<br />

Figure 2.07. Thermal Gravimetric Analysis (TGA), TA instrument, Q500 34<br />

Figure 2.08.<br />

Micromeritics ASAP 2020 Accelerated Surface Area<br />

and Porosimetry 36<br />

Figure 2.09. Raman vibrational and scattering modes 38<br />

Figure 2.10. Image of the micro Raman scattering equipment 39<br />

Figure 2.11. Perkin-Elmer PHI 3027 spectrometer and VG Escalab 210<br />

Spectrometer 42<br />

Figure 2.12. Main components of a FE-SEM instrument 44<br />

xiv


Figure 2.13. FE-SEM JEOL JM-6400 microscope 44<br />

Figure 2.14.<br />

TOC analyzer (Tekmar Dohmann Phoenix 8000 UV-<br />

Persulfate TOC Analyzer 46<br />

Figure 2.15.<br />

Image of the Leco CHNS 932 analyzer and scheme of its<br />

different components parts 46<br />

Figure 2.16.<br />

Representation of the different electronic transitions<br />

generated during the absorption process under UV-Visible<br />

irradiation 48<br />

Figure 2.17.<br />

Image of the fluorescence spectrophotometer Varian Cary<br />

Eclipse and a diagram of the fluorescence spectrometer 51<br />

Figure 3.01.<br />

Image of the CVD system and scheme of the CVD system<br />

and thermal treatment used for the synthesis of the Fe 2 O 3<br />

nanowires 56<br />

Figure 4.01.<br />

FE-SEM image of the titanium oxide (rutile phase) at a<br />

magnification of 5000x 58<br />

Figure 4.02. TGA scan of titanium oxide (rutile phase) 58<br />

Figure 4.03. Raman spectrum of TiO 2 sample (rutile phase) 59<br />

Figure 4.04. XPS spectrum corresponding to the O1s region of the TiO 2<br />

catalyst (rutile phase) 60<br />

Figure 4.05. XPS spectrum corresponding to the Ti2p region of the TiO 2<br />

catalyst (rutile phase) 60<br />

Figure 4.06.<br />

XRD diffraction pattern for TiO 2 -Anatase, TiO 2 -Rutile,<br />

TiO 2 @MWCNTs and TiO 2 NWs 61<br />

Figure 4.07.<br />

FE-SEM image of the titanium oxide (anatese phase) at a<br />

magnification of 50 000x 62<br />

Figure 4.08. TGA scan of titanium oxide (anatase phase) 62<br />

xv


Figure 4.09.<br />

Raman spectrum of titanium oxide catalyst (anatase<br />

phase) 63<br />

Figure 4.10.<br />

XPS spectrum of Ti2p peak on titanium oxide (anatase<br />

phase) 64<br />

Figure 4.11.<br />

XPS spectrum of TiO 2 showing the O1s transition<br />

(anatase phase) 64<br />

Figure 4.12.<br />

FE-SEM image of the as-synthesized TiO 2 NWs at different<br />

Magnifications 65<br />

Figure 4.13. TGA analysis of the as-synthesized TiO 2 NWs 66<br />

Figure 4.14. Raman spectrum of the as-synthesized TiO 2 NWs 67<br />

Figure 4.15. XPS spectrum of Ti2p region of the as-synthesized TiO 2 NWs 67<br />

Figure 4.16. XPS spectrum of O1s region of the as-synthesized TiO 2 NWs 68<br />

Figure 4.17. FE-SEM image of the as-synthesized TiO 2 @MWCNTs 69<br />

Figure 4.18. TGA analysis of the as-synthesized TiO 2 @MWCNTs 70<br />

Figure 4.19. Raman spectrum of the as-synthesized TiO 2 @MWCNTs 71<br />

Figure 4.20.<br />

XPS spectrum corresponding to the C1s region of the as-<br />

synthesized TiO 2 @MWCNTs catalyst 72<br />

Figure 4.21.<br />

XPS spectrum corresponding to the Ti2p region of the as-<br />

synthesized TiO 2 @MWCNTs catalyst 72<br />

Figure 4.22.<br />

XPS spectrum corresponding to the O1s region of the as-<br />

synthesized TiO 2 @MWCNTs catalyst 73<br />

Figure 4.23.<br />

FE-SEM image of the as-synthesized ZnO particles at<br />

different magnification 74<br />

Figure 4.24. TG curve of the as-synthesized ZnO particles 75<br />

Figure 4.25. Raman spectrum of the as-synthesized ZnO particles 75<br />

Figure 4.26. XRD diffraction pattern of the as-synthesized ZnO particles 76<br />

xvi


Figure 4.27.<br />

FE-SEM image of the as- as-synthesized iron oxide<br />

nanowires (Fe 2 O 3 NWs) at different magnifications 77<br />

Figure 4.28. TG curve of raw Fe 2 O 3 NWs 77<br />

Figure 4.29.<br />

XPS spectrum corresponding to the Fe2p region of the as-<br />

synthesized Fe 2 O 3 NWs 78<br />

Figure 4.30.<br />

XPS spectrum corresponding to the O1s region of the as-<br />

synthesized Fe 2 O 3 NWs 79<br />

Figure 4.31. Raman spectrum of the as-synthesized Fe 2 O 3 NWs particles 79<br />

Figure 4.32.<br />

Magnetic susceptibility of the as-synthesized Fe 2 O 3 NWs,<br />

measured at room temperature 80<br />

Figure 4.33.<br />

XRD diffraction patterns of Fe 2 O 3 NWs synthesized at<br />

600 °C and 700 °C at atmspheric pressure and in flowing<br />

Oxygen 81<br />

Figure 4.34.<br />

FE-SEM images of the as-synthesized capped magnetite<br />

nanoparticles (Fe 3 O 4 ) at different magnifications 83<br />

Figure 4.35.<br />

XPS spectrum corresponding to the Fe2p region, of the as-<br />

synthesized capped magnetite nanoparticles (Fe 3 O 4 ) 84<br />

Figure 4.36.<br />

XPS spectrum corresponding to the O1s region, of the as-<br />

synthesized capped magnetite nanoparticles (Fe 3 O 4 ) 84<br />

Figure 4.37.<br />

Raman spectrum of the as-synthesized capped magnetite<br />

Nanoparticles 85<br />

Figure 4.38.<br />

Temperature effect on the magnetite properties of the<br />

magnetite at different temperatures 85<br />

Figure 4.39. TG curve of the ferrous chloride 86<br />

Figure 4.40.<br />

XPS spectrum corresponding to the Cl2p region, of the<br />

FeCl 2 catalyst 87<br />

xvii


Figure 4.41.<br />

XPS spectrum corresponding to the Fe2p region, of the<br />

FeCl 2 catalyst 87<br />

Figure 4.42. XRD diffraction pattern of the FeCl 2 catalyst 88<br />

Figure 4.43. FE-SEM images of CuO at different magnification 89<br />

Figure 4.44. TG curve of the cupric oxide catalyst 89<br />

Figure 4.45. XRD diffraction pattern of the CuO catalyst 90<br />

Figure 5.01.<br />

Effects of the concentration of anatase on the<br />

photodegradation process of RhB 92<br />

Figure 5.02.<br />

Effects of the catalyst and hydrogen peroxide on the<br />

photodegradation process of RhB 93<br />

Figure 5.03.<br />

Effects of the pH of the reaction mixture on the<br />

photodegradation process of RhB 94<br />

Figure 5.04.<br />

Effects of the temperature of the solution on the<br />

photodegradation process of RhB 96<br />

Figure 5.05.<br />

Experimental setup used during this research, without<br />

irradiation and during the irradiation 101<br />

Figure 5.06. Dye solution used during the investigation 101<br />

Figure 5.07. Methylene blue visible absorption spectrum 102<br />

Figure 5.08.<br />

Visible absorbance abd fluorescence spectrum of MB in<br />

presence of rutile 104<br />

Figure 5.09.<br />

Possible degradation intermediates of RhB during the<br />

photocatalytic process 105<br />

Figure 5.10.<br />

Graphic of the percent of degradation of the different<br />

organic compounds by the Photocatalytic process 107<br />

Figure 5.11.<br />

Regression curve of the Methylene Blue (MB) with rutile<br />

under photochemical process 108<br />

xviii


Figure 5.12.<br />

Possible processes involved in the degradation reaction<br />

using TiO 2 as catalyst 112<br />

Figure 5.13.<br />

Absorption spectrum corresponding to the degradation of<br />

Rhodamine B by TiO 2 @MWCNTs under photochemical<br />

Process 113<br />

Figure 5.14. Four principal by-products of the MO degradation process 115<br />

Figure 5.15.<br />

Possible intermediates of degradation of MO during the<br />

photocatalytic degradation 116<br />

Figure 5.16.<br />

Schematic diagram of the sonochemical generation of the<br />

degradation radicals 118<br />

Figure 5.17.<br />

Degradation curves of RhB; UV-vis absorbance , TOC,<br />

fluorescence and dye solution before and after the sono-<br />

Fenton degradation process 119<br />

Figure 5.18.<br />

Graphic of percent of degradation of the organic<br />

compounds by the Sono-Fenton process 120<br />

Figure 5.19. Regression curve of the Methylene Blue (MB) with Fe 3 O 4<br />

under sono-Fenton process 121<br />

Figure 5.20.<br />

Scheme of the different areas of interest during the<br />

sonochemical process 124<br />

Figure 5.21.<br />

Degradation curves of MO; UV-vis absorbance, TOC,<br />

fluorescence and dye solution before and after the photo-<br />

Fenton degradation process 126<br />

Figure 5.22.<br />

Graphic of percent of degradation of the organic<br />

compounds by the Photo-Fenton process 128<br />

Figure 5.23. Regression curve of the Methylene Blue (MB) with Fe 3 O 4<br />

during photo-Fenton degradation process 128<br />

xix


Figure 5.24.<br />

Graphic of comparison between the Photocatalytic process<br />

and the Photo-Fenton process 130<br />

Figure 5.25.<br />

Graphic of comparison between the Photocatalytic process<br />

and the Sono-Fenton process for MB, RhB and MO 131<br />

xx


List of Appendices<br />

Page<br />

Appendix One Dyes Solutions 151<br />

Appendix Two Photocatalytic Process 152<br />

Appendix Three Sono-Fenton Process 182<br />

Appendix Four Photo-Fenton Process 206<br />

xxi


Abstract<br />

María <strong>del</strong> Carmen Cotto-Maldonado (PhD, Environmental Science)<br />

<strong>Heterogeneous</strong> <strong>Catalysis</strong> <strong>Applied</strong> <strong>To</strong> Advanced Oxidation Processes (AOPs) For<br />

Degradation of Organic Pollutants<br />

(April/2012)<br />

Abstract of a doctoral dissertation at the <strong>Universidad</strong> <strong>del</strong> <strong>Turabo</strong><br />

Dissertation supervised by Professor Francisco M Marquez Linares<br />

No. of pages in text 260<br />

Water is an essencial resource for humankind and biomes.<br />

Actually, the<br />

pollution of the water resources, specially the contamination of the fresh water is great<br />

concern in our society. Develop of new and more efficient method for degradation of<br />

pollutant in water increase the research in this area, especially in the AOPs. During this<br />

investigation a comparison between different AOPs methods (photocatalysis, sono-<br />

Fenton and photo-Fenton) to determine the most efficient process of them was done. <strong>To</strong><br />

reach our goal, different catalysts, namely TiO 2 nanowires, TiO 2 @CNTs, ZnO<br />

nanoparticles, Fe 2 O 3 nanowires and magnetite nanoparticles were synthesized and<br />

characterized by different techniques including FE-SEM, TGA, specific surface area<br />

(BET), XRD, Raman spectroscopy, XPS and magnetic susceptibility. Commercial and<br />

synthesized catalysts were used in photocatalysis, sono-Fenton and photo-Fenton<br />

processes for the degradation of mo<strong>del</strong> organic compounds (Methylene Blue,<br />

Rhodamine B, Methyl Orange, Gential Violet, Methyl Violet and p-aminobenzoic acid).<br />

According with the experimental results, no significant differences were observed<br />

between the photo-Fenton and sono-Fenton processes when the same catalysts were<br />

used. For the photocatalytic process, the more effective catalyst was TiO 2 NWs and for<br />

the sono-Fenton and photo-Fenton processes, the more effective catalyst was FeCl 2 .<br />

xxii


Resumen<br />

María <strong>del</strong> Carmen Cotto Maldonado (PhD, Environmental Science)<br />

<strong>Heterogeneous</strong> <strong>Catalysis</strong> <strong>Applied</strong> <strong>To</strong> Advanced Oxidation Processes (AOPs) For<br />

Degradation of Organic Pollutants<br />

(Abril/2012)<br />

Resumen de disertación doctoral en <strong>Universidad</strong> <strong>del</strong> <strong>Turabo</strong><br />

Disertación fue supervisada por el Profesor Francisco M Marquez Linares<br />

No. de páginas 260<br />

El agua es un recurso esencial para la vida humana y los biomas. Actualmente,<br />

la contaminación de los recursos acuáticos, especialmente de la contaminación de los<br />

abastos de agua potable ha creado una gran preocupación en nuestra sociedad. El<br />

desarrollo de nuevos y más eficientes métodos para la degradación de los<br />

contaminantes en agua se ha incrementado, especialmente en la utilización de los<br />

procesos de oxidación avanzada (AOPs, por sus siglas en ingles).<br />

Durante esta<br />

investigación se llevó a cabo una comparación de la eficiencia entre diferentes procesos<br />

de “AOPs” (fotocatálisis, sono-Fenton y foto-Fenton). Para alcanzar la meta de nuestra<br />

investigación se han sintetizado diferentes catalizadores como nanohilos de TiO 2 , TiO 2<br />

depositado sobre nanotubos de carbono, partículas de ZnO, nanohilos de Fe 2 O 3 y<br />

nanopartículas de magnetita. Estos materiales han sido caracterizados mediante<br />

diferentes técnicas entre las que se incluyen microscopia electrónica de barrido con<br />

emisión de campo (FE-SEM, por sus siglas en inglés) análisis termogravimétrico,<br />

determinación de área específica (BET), difracción de rayos X (DRX), espectroscopia<br />

Raman, espectroscopia fotoelectrónica de rayos X (XPS, por sus siglas en inglés) y<br />

susceptibilidad magnética. Estos catalizadores de síntesis y otros comerciales fueron<br />

utilizados en los procesos de degradación estudiados (fotocatálisis, sono-Fenton y foto-<br />

xxiii


Fenton). El Azul de metileno, Rodamina B, Naranja de metilo, Cristal Violeta y ácido p-<br />

amino benzoico fueron los compuestos orgánicos mo<strong>del</strong>os utilizados durante los<br />

procesos de degradación.<br />

Según los resultados experimentales, no se observan<br />

diferencias significativas entre los procesos sono-Fenton y foto-Fenton cuando los<br />

mismos catalizadores son utilizados. En el proceso fotocatalítico de degradación, el<br />

fotocatalizador que presentó mayor eficiencia fue el correspondiente a nanohilos de<br />

óxido de titanio. Durante los procesos sono-Fenton y foto-Fenton, el catalizador más<br />

activo en la degradación de los compuestos estudiados fue el FeCl 2 .<br />

xxiv


Chapter One<br />

Introduction<br />

Water is an important resource in our society. Less than a 0.7% of the total of<br />

water in the Planet is fresh water and only 0.01% is accessible to be used (Garriga I<br />

Cabo 2007). This resource is essential for sustaining the basic human functions as<br />

health, agriculture and the integrity of the biomes (Garriga I Cabo 2007, UNEP et al.<br />

2002). One of the human basic rights, especially children, is access to safe water for<br />

drinking and other uses (UNEP et al. 2002) because biological and chemical<br />

contaminants compromise the water quality in the world. <strong>To</strong>day, some of the most<br />

discussed issues around the world are the sanitation, soil and water chemical pollution,<br />

air pollution, the degradation of water sources and natural resources (Garriga I Cabo<br />

2007, UNEP et al. 2002). Organic, inorganic, bionutrients and microorganisms are some<br />

of the most common contaminants in water (Garriga I Cabo 2007). One of the facts<br />

mentioned by UNEP et al. (2002) said:<br />

“at the dawn of the 21 st Century, about 18 per cent of the<br />

world’s population do not have access to safe drinking<br />

water, and nearly 40 per cent lack adequate sanitation”.<br />

In many regions of the world, water is a scarce resource, and in these places the reuse<br />

of the water is a relevant issue (Marin et al. 2007).<br />

The production and use of synthetic chemical products have experienced an<br />

important increase during the last century. These products imply a challenge to the<br />

environment (UNEP et al. 2002), due to the fact that the environment does not have the<br />

1


2<br />

required mechanisms to promote the degradation, and these contaminants can become<br />

highly toxic to many species including the human being. Humankind is responsible for<br />

the release of the pollutants to the environment in many of their normal activities like<br />

industrial processes, wastewater discharges, excessive use of pesticides, fertilizers, etc.<br />

Many contaminants could move through the trophic chains and be accumulated in the<br />

organisms (UNEP et al. 2002).<br />

This situation highlights the importance of more<br />

epidemiological studies to understand the effect (synergistic or antagonist) of the<br />

population exposure to environmental contaminants.<br />

According to the OAS (2010), one of the objectives of the “Sustainable<br />

development of the Americas” is the protection of the public health by keeping the<br />

drinking water free of microorganisms, heavy metals and hazardous pollutants and trying<br />

to strengthen the development and implementing laws, regulations and policies.<br />

Organizations in different countries as “Alianza para el Desarrollo Sostenible” (ALIDES),<br />

“Organización Panamericana de la Salud” (OPS), the “Comité Coordinador de<br />

Instituciones de Agua Potable y Saneamiento de Centroamerica”, the Environmental<br />

Protection Agency (EPA), etc. work together to establish laws and regulations to protect<br />

the environment (OAS 2010). Examples of countries that are working with to achieve a<br />

better quality of water and environment protection are Belize, Costa Rica, Guatemala,<br />

etc. (OAS 2010). Chile is another example of a Latin American country that presents a<br />

relevant interest in the protection of the water sources and the environment (UNEP<br />

[undated]). According to the report of the UNEP et al. (2002), in 2002 the 30% of the<br />

industrial effluents were discharged into sewage systems without the appropriate<br />

discharge treatments.<br />

Another example of the strong interest of the countries for the conservation and<br />

management of the water resources was the UNEP Conference entitled “Greening<br />

Water Law in Africa: Managing Freshwater Resources for People and the Environment”


3<br />

held in Kampala, Uganda. The main objective of this meeting was the analysis of the<br />

socio-economic infrastructure with the environmental protection and conservation of the<br />

resources in the African continent (UNEP 2010).<br />

The Clean Water Act (CWA) of 1977 described in the 33 USC §1251 et seq.<br />

establishes all the basis and regulations to avoid the pollution of the waters of the United<br />

States of America regulating the discharges. The CWA enables the Environmental<br />

Protection Agency to implement the regulatory standards for water quality and<br />

discharges through the National Pollutant Discharge Elimination System (NPDES).<br />

However, the NPDES only regulates specific discharges, including industrial and<br />

treatment plant effluents. Non specific discharges, as for instance the septic systems,<br />

are not regulated by NPDES.<br />

In Puerto Rico, Environmental Quality Board (EQB)<br />

establishes some discharge regulations to protect the quality of the waters according<br />

with the uses.<br />

The EPA has proposed a “Strategic Plan” for the fiscal years of 2011 to 2012 to<br />

protect and restore the waters in the USA, and specifically to protect the human and<br />

aquatic ecosystem health (USEPA [undated]). In the “Notice of Final 2010 Effluent<br />

Gui<strong>del</strong>ines Program Plan”, the effluents gui<strong>del</strong>ines and pretreatment standards are<br />

evaluated to maintain the integrity of the water sources (FR 2010). Meanwhile, Best<br />

Available Technology Economically Achievable (BAT) are promulgates for reach the<br />

EPAs goal as suggest in the Federal Register as a form to increase the effectivity of the<br />

treatments processes.<br />

The United States Geological Survey (USGS) as part of the US Department of<br />

the Interior is developing different studies to determine the level of contamination of<br />

different streams in the US. A survey from 1999 to 2000 of the USGS (Barnes et al.<br />

2002a, 2002b) has demonstrated the presence of 82 of the 95 organic wastewater<br />

contaminants analyzed. A total of 80% of the samples taken from 30 of the states were


4<br />

positive for the presence of at least 1 of the contaminants including antibiotics,<br />

hormones, detergents, plasticizers, disinfectants, insecticides, fire retarded (using during<br />

fibers synthesis), and antioxidants. Most commonly detected products are steroids, nonprescript<br />

drugs and insect repellents.<br />

Emerging Contaminants Project of the USGS (USGS 2011a) has the goal of<br />

provide information (analytical methods, environmental occurrence, pathways and<br />

ecological effects) of the contaminants that are not monitored due to the lack of<br />

regulation but have the potential to reach the environment in significant amounts, having<br />

adverse effects on the biosphere and specifically in humans.<br />

Different studies<br />

developed in New York and New Jersey were undertaken to determine the occurrence<br />

and concentration of emerging contaminants after treatment processes (USGS 2008).<br />

The coordinator of the USGS <strong>To</strong>xic Substances Hydrology Program, Herb Buxton,<br />

declared that:<br />

“The wastewater treatments are not really designed to remove<br />

those trace-organic chemicals”<br />

and these contaminants are normally released to the environment (USGS 2011b).<br />

Among the emerging contaminants found in the environment, the group of<br />

cosmetic and personal care products (PPCs) deserves special attention. The UV filters<br />

is one of the products commonly used and several studies demonstrate the presence in<br />

different water samples. The UV filters also possess potential risks derived from the<br />

presence of single or multiple aromatic groups in their structures and these substances<br />

are normally used in sunscreen lotions and many cosmetics. Comparison studies in<br />

Switzerland, between river and lake fish as Salmo trutta fario, Coregonus spp and R<br />

rutilus demonstrated the presence of UV-filters in muscle tissue of the fishes (Buser et<br />

al. 2006).<br />

Another study of Schlumpf et al. (2008) demonstrated the presence of<br />

sunscreen compounds from analyses of human milk.


5<br />

These results demonstrate the need for efficient water treatment technologies<br />

able to remove or degrade hazardous contaminants present in the effluents, making the<br />

water resources both safe and potable to human consumption. For example, to maintain<br />

the aesthetic and diminish the environmental impact of industrial effluents is necessary<br />

the discoloration of the wastewaters (Hussein et al. 2008). Currently, the most used<br />

treatment methods for the removal of contaminants from water are the reverse osmosis,<br />

ion-exchange technology, precipitation of materials and adsorption of the contaminants,<br />

especially using activated carbon (charcoal) and biological degradation (Gupta et al.<br />

2004; Mezohegyi et al. 2007). Other processes as Fenton, photochemistry, radiolysis or<br />

sonolysis generate highly reactive hydroxyl radicals for bleaching; finally arising to the<br />

mineralization of recalcitrant compounds (Ozen et al. 2005). In the boom of the ecoconservation<br />

and the eco-friendly techniques to degrade the pollutants in water and<br />

wastewater, the Advance Oxidation Processes or AOPs are seen as alternative<br />

techniques (Gupta et al. 2004) to the traditional processes.<br />

Techniques as hydrogen peroxide oxidation, ozonation, photolysis, Fenton<br />

process, photocatalytic oxidation, wet-air oxidation and ultrasonic sonication are<br />

considered as part of the AOPs used for the degradation of contaminants (Gupta et al.<br />

2004; Garriga I Cabo 2007). The AOPs use chemical procedures based on the use of<br />

catalysts or photochemical compounds which generate highly reactive transient species<br />

as the hydroxyl radical which possesses high affectivity for the oxidation of organic<br />

compounds (Marin et al. 2007). The AOPs are defined as “processes that involve in situ<br />

generation of free radicals” (Priyas and Madras 2006) with a highly potential oxidant<br />

such as the hydroxyl radicals (•OH) (Priyas and Madras 2006; Ai et al. 2007a; Garriga I<br />

Cabo 2007; Marin et al. 2007; Mosteo et al. 2008) being non-selective chemical oxidant<br />

processes (Ai et al. 2007a; Mosteo et al. 2008). These radicals are produced by the<br />

combination of the hydrogen peroxide, UV radiation, ozone and a semiconductor as


6<br />

titanium oxide or the combination of hydrogen peroxide with iron ions (Fenton reaction)<br />

(Marin et al. 2007). The radicals (OH • ) produced during the AOPs are powerful oxidants<br />

because they have high oxidative potential (E 0 OH/H2O=2.8 V) when compared with the<br />

normal hydrogen electrode (Ab<strong>del</strong>malek et al. 2006).<br />

AOPs have many advantages as: the complete mineralization of the pollutant,<br />

are non-selective process, can be used in low concentration of contaminants and can be<br />

combined with other methods (Garriga I Cabo 2007). The use and development of<br />

photocatalytic processes for the removal of harmful contaminants, as a treatment for<br />

wastewater and air pollutants is becoming increasingly popular (Yin et al. 2009).<br />

<strong>Heterogeneous</strong> photocatalysis is one of the AOP’s and is based on the direct or indirect<br />

absorption of photons from ultraviolet (UV) or visible light by a semiconductor that<br />

possesses the appropriate energy gap. According with Ruan and Zhang (2009) the “UV<br />

–driven photocatalytic activity of the sample is much higher than the visible light –driven<br />

photocatalytic activity” because the shorter the wavelength the higher quantum yield.<br />

Velegraki and Mantvinos (2008) describe the importance of the heterogeneous<br />

photocatalytic degradation as<br />

“organic compounds can then undergo both oxidative degradation<br />

through their reactions with valence band holes, hydroxyl and<br />

peroxide radicals and reductive cleavage through their reactions<br />

with electrons yielding various by-products and eventually mineral<br />

end-products.”<br />

The excitation of the semiconductor can take place by two different ways: i) the direct<br />

excitation of the semiconductor (direct absorption of the photons by the surface of the<br />

semiconductor) or ii) the excitation of molecules previously adsorbed on the surface of<br />

the semiconductor which transfer the electrons to the semiconductor (Marin et al. 2007).<br />

The direct absorption process of the photon causes the excitation of the surface or


7<br />

interface region between the solid and the liquid avoiding any chemical change in the<br />

catalyst (Marin et al. 2007). A distinctive characteristic of the interface is the charge<br />

redistribution to both sides of the interface (Marin et al. 2007). Vinu and Madras (2010)<br />

define photocatalysis as<br />

“the acceleration of the rate of chemical reactions<br />

(oxidation/reduction) brought about by the activation of a catalyst,<br />

usually a semiconductor oxide, by UV or visible radiation”.<br />

Other authors (Aarthi and Madras 2007) argue that<br />

“in aqueous environment, the holes created under the UV<br />

irradiation are scavenged by the hydroxyl groups present on the<br />

surface, generating OH• radicals, which promote the oxidation of<br />

the organics”.<br />

The semiconductor for the photocatalyst should be chemical and biological inert,<br />

stable, inexpensive, of easy synthesis and production, and without human and<br />

environmental risks (Garriga I Cabo 2007). When a dye is used, the mechanism of<br />

photodegradation involves the excitation of the dye and the transference of the electrons<br />

to the conduction band of the photocatalyst (i.e. TiO 2 ) to generate the dye radicals.<br />

These radicals react with the oxygen on the surface of the catalyst generating oxygen<br />

radical species as O •- 2 , H 2 O 2 and •O 2 remaining the valence band unaffected (Yin et al.<br />

2009).<br />

Another form to simplify this complex process is considering that the<br />

photocatalyst (i.e. titanium oxide) absorbs a photon having energy greater than or equal<br />

to the band gap (hv≥ E BG ); This energy absorption implies the promotion of an electron<br />

from the valence band to the conduction band of the photocatalyst. This promotion<br />

leaves a “hole” (positive charge) in the valence band giving place to the formation of


8<br />

“electron-hole” pairs. If the pairs migrate to the surface of the metal they can react with<br />

the solution (Prakash et al. 2009; Vinu and Madras 2010).<br />

Figure 1.01 shows a<br />

schematic view of the photoexcitation process experienced by a semiconductor.<br />

Figure 1.01. Schematic diagram for the photoexcitation process<br />

in a semiconductor via photon irradiation (Adapted from: Hu et al.<br />

2010; Vinu and Madras 2010).<br />

Considering this photocatalytic mechanism, the photodegradation process should<br />

be affected by the light source (irradiation energy), dye concentration, catalyst<br />

concentration and the presence of other organic substances or ions in the solution<br />

(Aarthi and Madras 2007, Yin et al. 2009). Some of the most common photocatalysts<br />

include TiO 2 , ZnO, ZnS, CdS, WO 3 , SrTiO 3 , and SnO 2 (Sokmen et al. 2000; Priya and<br />

Madra 2006). Additionally, catalysts with perovskite structure (Yu et al. 2009; <strong>To</strong>rres<br />

Martínez et al. 2010) are also used for photochemical reactions. For degradation on wet<br />

oxidation, different types of catalysts are used including heterogeneous catalysts of<br />

metal oxides (ZnO, CuO, MnO 2 , SeO 2 , TiO 2 , ZrO 2 , etc.), noble metals on alumina


9<br />

support and metal impregnation on activated carbon (Cu, Co, Bi, Fe, Mn) (Ma et al.<br />

2007).<br />

The photocatalysis is an AOP commonly used because is able to mineralize<br />

organic pollutants at low cost (Yu et al. 2009). Other important AOP is the ultrasonic<br />

irradiation. At the beginning of the 20 th century, Richards and Loomis described the use<br />

of the ultrasound irradiation technique as driving force for a chemical transformation<br />

(Priya and Madras 2006).<br />

Degradation process using ultrasound irradiation in<br />

heterogeneous catalysis can be increased due the formation of radicals as •OH during<br />

the cavitation process (Shimizu et al. 2007). Cavitation phenomenon in liquids includes<br />

the nucleation, growth and collapse of small bubbles (Shimizu et al. 2007). According to<br />

the authors (Shimizu et al. 2007), cavitation is fundamental for the chemical and<br />

mechanical process occurring during the ultrasound irradiation. This process can induce<br />

the increase of temperature in hot spots of thousands of Kelvin (T=4000 K) in an<br />

adiabatic heating, and pressures in the scale of hundreds of atmospheres (313 atm)<br />

leading the dissociation of the water molecules producing hydrogen atoms and hydroxyl<br />

radicals (•OH) (Priyas and Madras 2006; Shimizu et al. 2007). These radicals can<br />

produce many chemical reactions (sonochemical reactions) (Shimizu et al. 2007;<br />

Kavitha and Palanisamy 2011).<br />

The use of the sonochemical reactions could be<br />

potentially used in environmental processes as wastewater treatments (Shimizu et al.<br />

2007). Semiconductor catalysts commonly used in ultrasonic degradation are Fe 2 O 3 ,<br />

TiO 2 , ZnO and CuO (Priya and Madra 2006).<br />

Priya and Madra (2006) point out that one of the main advantages of the<br />

ultrasonic irradiation with respect to the photocatalytic process is the elimination of the<br />

“spatial limitation” over the catalyst, because cavitation process increases the generation<br />

of radicals and is extended along the solution and it is not exclusively limited to the


10<br />

catalyst surface. This process increases the surface area, avoiding the occlusion of the<br />

active sites on the surface, reducing the mass-transfer limitations. Authors describe the<br />

“spatial limitation” as a problem caused by light screening effects produced during the<br />

photocatalytic reaction that reduces the excitation area on the surface of the catalyst<br />

(Priya and Madra 2006). Use of the ultrasonic irradiation for the degradation process of<br />

a dye could be affected by many factors as the concentration of the catalyst and the dye,<br />

the presence of anions, the pH and presence of scavenger agents in the solution<br />

(Shimizu et al. 2007). Some authors (Shimizu et al. 2007) studied the synergistic effect<br />

of the photochemical process on the sonochemical process (ie. degradation of salicylic<br />

acid).<br />

The Fenton reaction as part of the AOPs generates hydroxyl radicals. This<br />

process is clearly non selective (Ai et al. 2007b) and represents a viable technique to<br />

degrade hazardous organic compounds. Horstman-Fenton and Jackson (1899)<br />

demonstrated the importance of iron and hydrogen peroxide during the oxidation of<br />

some substances. A characteristic of the Fenton process is that the reaction requires<br />

acid conditions to work more efficiently (pH ranging from 2 to 3) (Ai et al. 2007b). The<br />

development of a Fenton process working efficiently in a neutral pH should be an<br />

advance, because it is unnecessary the decrease of the pH before the reaction takes<br />

place, decreasing the generation of sludges during the process and increasing the<br />

possibility to recuperate the Fenton reagent (iron) from the media (Ai et al. 2007b).<br />

<strong>To</strong>xicological studies with L gibba demonstrated the degradation of substances as<br />

sulfonamides (antibiotics) using anodic Fenton treatment (AFT) in solutions with<br />

concentration of 100 μM (Neafsey et al. 2010).<br />

In a biological process, Hotta et al. (2010) demonstrated that the addition of Fe 2+<br />

ion stimulated the cell growth of Sphingomonas spp increasing the biodegradation<br />

activity of the alkylphenol polyethoxylates or APEO n (used as detergents, emulsifiers and


11<br />

pesticides) increasing the production of endocrine active metabolites. According with<br />

the authors (Hotta et al. 2010) three possible classifications of the microbial degradation<br />

of man-made compounds in the environment are possible; biodegradation rate increases<br />

by stimulation of the cell growth by a chemical substance, by minerals presents in the<br />

media and by enzyme induced and/or stimulation of the minerals.<br />

It is relevant to define some important terms concerning these reactions (i.e.<br />

Fenton and Fenton-like reagents). According to Ai et al. (2007a), the Fenton reagent<br />

can be defined as the combination of hydrogen peroxide and iron (II) (Fe 2+ /H 2 O 2 ).<br />

Fenton-like reagent does not include iron (II) species and normally this term is used for<br />

the combination of Fe 3+ /H 2 O 2 although both reagents (Fenton and Fenton-like) are<br />

present during the reaction because both iron species are in equilibrium during the<br />

reaction. The Fenton-like reagent is capable of oxidizing organic substrates, but it is<br />

somewhat less reactive than Fenton reagent. Similarly to Fenton reactions produced by<br />

Fe 2+ /Fe 3+ in presence of hydrogen peroxide, Randorn et al. (2004) demonstrated that<br />

analogous processes are also observed with other transition metals as titanium. The<br />

reactive titanium species involved in the Fenton reaction are Ti 3+ /Ti 4+ .<br />

Use of the Fenton and Fenton-like reactions has two principal disadvantages for<br />

the use in large scale; the first one is the high cost of the reagents required (H 2 O 2 ) and<br />

their instability in solution, and the second one is the concerns involved in the use of<br />

acidic pH (pH < 4) and with a narrow range of pH values during the process (Ai et al.<br />

2007b). As Fenton reagents (Ai et al. 2007b) different compounds including hematite,<br />

goethite, clays, iron hydroxide and iron supported on different materials have been<br />

evaluated.<br />

An alternative to the typical Fenton reactions based on the use of soluble<br />

Fe 2+ /Fe 3+ species is the use of Fe 0 phase as a supported or immobilized catalyst and<br />

hydrogen peroxide as oxidizer (Ai et al. 2007a). Examples of possible reagents for


12<br />

environmental remediation include Fe 0 , Fe 3 O 4 (magnetite) and Fe 2 O 3 (maghemite) (Ai et<br />

al. 2007a). The Fe 0 is used to remove organic compounds from the soil and Fe 3 O 4 and<br />

Fe 2 O 3 are normally used for the degradation of organic compounds in solution. Ai et al.<br />

(2007a) demonstrated the efficiency of the Fe@Fe 2 O 3 core-shell nanowires for the<br />

degradation of RhB.<br />

These materials have been synthesized by using different<br />

procedures as chemical vapor deposition, different metal oxidation processes and wet<br />

chemistry (Huh et al. 2010). Some authors point out the disadvantage of using zinc<br />

compounds because they are easily oxidized and the zinc oxide is weak and instable<br />

forming zinc hydroxide in some solutions (Randorn et al. 2004).<br />

Magnetites and related materials attracted a great deal of attention when a<br />

Martian meteorite was analyzed and these materials where found as one of their main<br />

components (Nyiro-Kosa et al. 2009). The magnetite is a versatile material due to their<br />

interesting applications in different fields such as catalysis, information storage,<br />

optoelectronics and biomedical applications that include magnetic bioseparation,<br />

magnetic resonance imaging contrast enhancement and targeted drug (Marquez et al.<br />

2011, 2012). For these applications, the particle size of the magnetites should range<br />

from 30 to 120 nm (Nyiro-Kosa et al. 2009). Proteins of magnetotactic bacteria can be<br />

used to biomimetic the natural process in the lab (Nyiro-Kosa et al. 2009. The magnetite<br />

size is influenced by different parameters, including the concentration of the reagents,<br />

temperature, pH of the solution and the reaction time (Nyiro-Kosa et al. 2009). Among<br />

the possible methods for the synthesis of magnetites the co-precipitation, pyrolysis,<br />

ultrasound irradiation, hydrothermal or electrochemical approach can be considered as<br />

the most useful and with higher yields than other processes (Nyiro-Kosa et al. 2009).<br />

The use of the nanomaterials for environmental and energy applications has<br />

experienced an important increase due to the development of new synthesis processes<br />

to manufacture these new materials at atomic and molecular scale.<br />

As a result,


13<br />

materials can be designed to have different chemical and/or physical properties<br />

according to the interest of the investigator (Hu et al. 2010).<br />

The electronic band<br />

structures finally determine the properties of the inorganic catalyst (Osterloh 2008). The<br />

use of transition metals is relevant due to the presence of d orbitals in their electronic<br />

configuration (Randorn et al. 2004). Many synthesis techniques as co-precipitation, solgel,<br />

microemulsions, freeze drying (or lyophilization), hydrothermal processes, chemical<br />

vapor deposition, etc. are commonly used to control the morphology, size and the<br />

uniformity of the structured nanoparticles so grown (Hu et al. 2010). Cao (2004) studied<br />

the electrophoretic deposition for the synthesis of titanium oxide nanorods synthesis.<br />

The use of heterogeneous catalysis, by using nanosized catalysts as TiO 2 ,<br />

demonstrated the complete mineralization of the hazardous substances to CO 2 and<br />

water by means of •OH radicals generated during the photochemical process (Shimizu<br />

et al. 2007). The TiO 2 has important applications in green chemistry because it is<br />

commonly used as a catalyst for the synthesis of pharmaceutical products, reducing the<br />

traditional large amount of waste because can be recoverable, increasing the yield of<br />

products (Prakash et al. 2009). An example of this improvement in the pharmaceutical<br />

production is the modification in the Biginelli’s reaction on the synthesis of<br />

dihydropyrimidin-2 (1H)-ones (Prakash et al. 2009). The titanium oxide is inexpensive,<br />

non-toxic in nature, stable under ambient conditions, environmental friendly, able to use<br />

the solar radiation (Randorn et al. 2004, Marin et al. 2007, Yin et al. 2009, Velegraki and<br />

Mantvinos 2008), antibacterial activity (Parthasarathi and Thilagavathi 2009), interesting<br />

optical and electronic properties, low cost, abundance (Velegraki and Mantvinos 2008)<br />

and it is appropriate for some oxidation or reduction reactions in aqueous solutions<br />

(Prakash et al. 2009). Titanium oxide nanowires are other interesting structures of the<br />

oxide and have been used, among other applications, for the degradation of pollutants<br />

by photocatalysis and for the production of hydrogen by a photocatalytic water splitting


14<br />

process (Huh et al. 2010). The TiO 2 has a band gap of 3.2 eV (Marin et al. 2007;<br />

Prakash et al. 2009) which is relevant for the photocatalytic activity. Other advantages<br />

of the titanium nanowires include the high specific surface area and an easy recovery<br />

process by filtration, centrifugation, etc. (Huh et al. 2010). The titanium oxide is wi<strong>del</strong>y<br />

studied because it possesses photocatalytic and photoconductor characteristics; is used<br />

for the degradation of azo dyes, volatile organic compounds and others (Hernandez<br />

Enriquez et al. 2008) and could be recoverable after the process (Rahmani et al. 2008).<br />

Another pollutant studied was phenol by titanium oxide (anatase) (Rahmani et al. 2008).<br />

Titanium oxide nanostructured films have also been used for the degradation of stearic<br />

acid (Takahashi et al. 2011).<br />

In the recombination process between the valence band and conduction band a<br />

low quantum yield should be observed; to resolve this situation some authors<br />

recommend the use of the transition metal and their oxide to create an electron trap to<br />

increase the efficiency (Li and Shang 2010; Zhou et al. 2010).<br />

The use of PdO<br />

nanoparticles on titanium oxide nanotubes is an alternative to create an electron trap to<br />

increase the lifetime of charge carriers and subsequently improve the photoactivity (Li<br />

and Shang 2010). Several materials as the Pt@TiO 2 NWs synthesized by hydrothermal<br />

process are an example of materials in which the Schottky effect for the degradation<br />

process is relevant (Wang et al. 2010). Synthesis in gas phase of titanium oxide doped<br />

with SiO 2 and the synthesis of Ag 2 O/TiO 2 are also known (Remnev et al. 2009, Zhou et<br />

al. 2010).<br />

Other types of oxide catalysts with catalytic applications are the<br />

nanostructured Mn 2 O 3 (Su et al. 2010) and CuO-MoO 3 -P 2 O 5 materials (Ma et al. 2007).<br />

According to Marin et al. (2007), the use of a sol-gel approach for the synthesis<br />

of TiO 2 over different supports (i.e. glass) is a good method because the synthesized<br />

product is obtained as a stable and homogeneous sheet of titanium oxide, catalytically<br />

active and resistant. Randorn et al. (2004) mentioned the importance of some thermal


15<br />

treatments in this catalyst (i.e. calcination in presence of oxygen) to increase the<br />

interactions between OH - from water to the surface of the catalyst.<br />

The carbon nanotubes (CNTs) are another important group of materials in the<br />

development of nano-optical and electronic devices as quantum memory elements,<br />

magnetic storage media and semiconducting devices due to their internal structures,<br />

high surface area, low density and chemical stability (Hussein Sharif Zein and<br />

Boccaccini 2008).<br />

These materials can be modified adding other materials to the<br />

surface of the CNTs (Hussein Sharif Zein and Boccaccini 2008). The carbon nanotubes<br />

are “cylindrical molecules formed by one or more sheets of carbon atoms rolled one over<br />

one” (Anson Casaos 2005) with a diverse range of diameters and lengths. This material<br />

has been extensively studied during the last years due to its special geometry and<br />

amazing properties (Anson Casaos 2005).<br />

The carbon nanotubes are classified in two main groups: multiple concentric<br />

nanotubes precisely nested within one another namely multi walled carbon nanotubes<br />

(MWCNTs) and nanotubes with a single wall (SWCNTs) (Lopez-Fernandez 2009). The<br />

nanotubes are composed by sheets of graphene. The graphene structure is composed<br />

by carbon atoms having a hexagonal arrangement of carbons with sp 2<br />

hybridization<br />

(Anson Casaos 2005). In the SWCNTs the graphene sheet is rolled to form the tube,<br />

meanwhile in the case of MWCNTs the structure is formed by concentric cylinders<br />

(Lopez-Fernandez 2009, Hernández Rueda 2010). A secondary classification of the<br />

nanotubes is based on the chirality (as “zig-zag”, “armchair” and chiral), diameter and<br />

quantity of walls (Anson Casaos 2005; Hernandez Rueda 2010). The chirality defines<br />

the possible behavior of the nanotube; i.e. metallic behavior defines electrical properties<br />

of the material (Lopez-Fernandez 2009). The surface of carbon nanotubes can be<br />

related to the possible uses in some applications (Anson-Casaos 2005) including<br />

hydrogen storage and fuel cells (Anson Casaos 2005).


16<br />

The nanotubes are commonly as aggregates. <strong>To</strong> disperse these materials is<br />

necessary to use mechanical dispersion, ultrasound and functionalization techniques<br />

(Hernandez Rueda 2010). These materials can be functionalized by two principal<br />

approaches; the covalent and supramolecular functionalization.<br />

These methods<br />

preserve the structural and electronic integrity of the materials (Lopez-Fernandez 2009).<br />

Treatments as purification with acids (acid reflux), thermal oxidations (in air and high<br />

temperature) or chemical activations can partially modify the structure of the nanotubes<br />

(Anson Casaos 2005). The most relevant structural modification consists in opening the<br />

ends of the nanotubes because the original structure is closed as a capsule (Anson<br />

Casaos 2005).<br />

The use of mo<strong>del</strong> contaminants is relevant for the study of many processes. The<br />

most common substances used as mo<strong>del</strong> contaminants are the organic dyes. Until the<br />

XIX century, synthetic dyes were used as inks (Confortin et al. [unknown date]). The<br />

dyes have different applications in paper industries, leather, cosmetics, drugs,<br />

electronics, plastics and printing (Vinu and Madras 2009). According to the authors<br />

(Vinu and Madras 2009) 80% of the synthetic dyes are consumed by the textile industry.<br />

Some authors have determined that the annual discharge of waters containing dyes<br />

ranges from 30 000-150 000 tons (Vanhulle et al. 2008). These wastewaters also<br />

contain other chemicals used during the processes (Vinus and Madras 2009). <strong>To</strong>rres<br />

Martinez et al. (2010) point out that according to some statistical results approximately<br />

12% of the synthetic textile dyes used during a year are “lost” during the manufacturing<br />

and operational procedures and from that 12%, the 20% will be finally released to the<br />

ecosystem through the industrial water discharges.<br />

In the textile industry, more than 10 000 different dyes and pigments are<br />

available in the market and 20-30% of them are reactive dyes (Karadag et al. 2006;<br />

Dafnopatidou et al. 2007). These dyes are characterized by their brilliant colors, high


17<br />

wet fastness, easy application and a minimum of energy applied during the process.<br />

These dyes have, as part of the structure, azo, anthraquinone, phthalocyanine, formazin<br />

or oxazine functional groups (Karadag et al. 2006). Approximately 60% of the reactive<br />

dyes used contain an azo group (Karadag et al. 2006).<br />

Dyes are nonbiodegradable compounds (Mahanta et al. 2008).<br />

Industrial<br />

wastewaters that contain biorefractory compounds are normally limited to the use of<br />

chemical treatments because the chemicals are toxic to the microorganisms used in the<br />

conventional biological treatments (Barrera-Diaz et al. 2009). Potential human exposure<br />

to wastewater which contains dyes is a concern because are carcinogenic compounds,<br />

showing high resistance against biological, physical and chemical reactions (Vanhulle et<br />

al. 2008). Different processes are employed to remove color from wastewaters including<br />

the use of activated carbon, membrane filtration, ultrafiltration, coagulation-flocculation,<br />

electrocoagulation, UV light and ozone (Barrera-Diaz et al. 2009).<br />

The effluents of the textile industry have high concentrations of organic and<br />

inorganic dyes which are strongly colored, have high chemical oxygen demand (COD),<br />

present important fluctuations in the pH, and are toxic to the organism (Ab<strong>del</strong>malek et al.<br />

2006). The common techniques used to remove the dyes include chemical, physical<br />

and biological processes (Dafnopatidou et al. 2007). Nevertheless, these conventional<br />

processes for the treatment of sewage waters including the degradation of residual<br />

dyestuffs are inefficient because these compounds have high molecular weight and<br />

biochemical stability (aromatic rings) (Panizza et al. 2006; Ma et al. 2007).<br />

The<br />

adsorption process using activated carbon to eliminate the contaminants has the<br />

advantage that is very easy to use but this method is expensive (Gupta et al. 2004) and<br />

produces another problem during the disposition of the contaminated material. Another<br />

method is the adsorption of the dye by polymers and other materials (Karadag et al.<br />

2006; Mahanta et al. 2008). The conventional treatments do not reduce the toxicity of


18<br />

the dyes (Barrera-Diaz et al. 2009). One of the principal disadvantages of the physical<br />

methods as coagulation, precipitation and adsorption is the sludge formation, possible<br />

toxic by-products and the chemical processes are expensive (Panizza et al. 2006; Ma et<br />

al. 2007; Hernandez Enriquez et al. 2008; Mahanta et al. 2008).<br />

Most of the dyes are organic or organometallic compounds characterized by<br />

having aromatic rings. This characteristic necessarily implies the use of treatments by<br />

unconventional methods (<strong>To</strong>rres Martinez et al. 2010). The decomposition of many<br />

organic compounds as pesticides, dyes, aromatics, halogenated aliphatic compounds,<br />

metallurgical residuals, oil and chemical compounds derived from steel processes are<br />

based on photocatalytic degradation processes (Sokmen et al. 2000). Meanwhile, to<br />

maintain the aesthetic and reduce the environmental impacts of the industrial effluents is<br />

necessary the discoloration of wastewaters (Hussein Sharif Zein et al. 2008).<br />

The degradation process of the organic dyes could be defined in two different<br />

ways: one is the discoloration and the other one is the mineralization (Vinu and Madras<br />

2010). The authors (Vinu and Madras 2010) also clarify the difference between<br />

discoloration process (reduction of the parent dye) and mineralization (complete removal<br />

of the organic components and their transformation in CO 2 ). Intermediates that are<br />

generated during the degradation process could be colored (Vinu and Madras 2010).<br />

The total organic carbon (TOC) analysis helps to determine the carbon content and its<br />

variation during the degradation process.<br />

Different dyes exposed to photochemical<br />

degradation under visible light show the following degradation order: indigo<br />

≈phenanthrene > triphenylmethane > azo ≈ quinoline > xanthenes ≈ thiazine ><br />

anthraquinone. The order of the light sources is: natural sunlight >> 90 W halogen flood<br />

light > 150 W spotlight (Vinu and Madras 2010).


19<br />

The organic dyes are classified according to their functional groups as: azoic,<br />

anthraquinonic, heteropolyaromatic, aryl methanes, xanthenes, indigo, acridine, nitro,<br />

nitroso, cyanine and stilbene (Vinu and Madras 2010).<br />

Table 1.1 Characteristics of the most important dyes classes (Adapted from: Parshetti et<br />

al. 2006; Vanhulle et al. 2008).<br />

Classes of<br />

Dyes<br />

Type of Fiber<br />

Chemical Class<br />

Acid Polyamide, wool<br />

and nylon<br />

Anthraquinones, azo, triarymethanes azo<br />

or metal complex azo, phtalocyanines<br />

Reactive<br />

Disperse<br />

Cellulose, polyester,<br />

acetates<br />

Small azo or nitro compounds, multi azo,<br />

phtalocyanines, stilbenes<br />

Direct vat Cellulose, rayon Indigoids, diarylmethanes, triarylmethanes<br />

Basic Sulfur Acrylic, polyester,<br />

cellulose<br />

Polymer with S-containig heterocycles,<br />

azo<br />

Different mo<strong>del</strong> organic dyes have been selected for this research due to their<br />

different structures (functional groups) and their presence in the environment. According<br />

to Vinu and Madras (2009), the degradation reaction of a dye by a hydroxyl radical<br />

generated by UV irradiation of ultrasonic is as follow:<br />

TiO 2 (OH·) ads – D ads + TiO 2 – D ads (or D) → intermediates (P) → CO 2 + H 2 O


20<br />

Methylene blue (MB) (Figure 1.02) is a hetero-polyaromatic dye (Ma et al. 2007)<br />

commonly used for printing cotton, as textile tannin and for coloring leather. MB is also<br />

used in chemistry as a base-acid indicator and in the medical field as an antiseptic<br />

(Gupta et al. 2004).<br />

During the photocatalytic degradation process of MB, some<br />

transients were detected when nanosized TiO 2 was used as catalyst, including 3-<br />

dimethylamino aniline, benzene sulfonic acid, phenol and hydroxylated products of<br />

amino and sulfoxide groups (Vinu and Madras 2010). Huh et al. (2010) developed a<br />

study on the degradation of MB using visible light. Authors used a standard white light<br />

bulb (100 mW•cm -2 ) as the visible light source. Starting concentration of the MB was<br />

10 -5 mol/L.<br />

H C<br />

Cl -<br />

3<br />

N<br />

S +<br />

H 3 C<br />

N<br />

CH 3<br />

CH3<br />

N<br />

Figure 1.02. Molecular structure of Methylene<br />

blue.<br />

Rhodamine B (RhB) is a dye that belongs to a class of compounds called<br />

xanthenes (Figure1.03), extensively used as mo<strong>del</strong> compound because it shows a<br />

strong absorption band in the visible region of the electromagnetic spectrum (555 nm)<br />

and this dye is characterized by having a high stability at different pH values. Ai et al.<br />

(2007), argue two possible competitive mechanisms during the degradation of the RhB.<br />

The first one is the N-demethylation and the second one is the breakdown of the<br />

xanthene structures. This dye is currently used as dye laser material (Aarthi and Madras<br />

2007) and is part of the triphenylmetane family of dyes that contain four N-ethyl groups<br />

at both sides of the xanthene rings (Yu et al. 2009). Also it is stable in aqueous solution


21<br />

(Yin et al. 2009). The RhB is also used as a dye for wool and as analytical reagent<br />

during the determination of metals in solution, especially alkali, and alkaline earth<br />

metals. This dye is used in the textile, food and cosmetic industries but can cause<br />

aesthetic pollution in the aquatic environments showing high resistance to biological and<br />

chemical degradation (Yu et al. 2009). Currently this dye has been prohibited for the<br />

use as food color because it is suspected that RhB could be a carcinogenic substance<br />

(Gupta et al. 2004).<br />

H 3 C<br />

C<br />

H 3<br />

N<br />

O N + CH 3<br />

CH 3<br />

COOH<br />

Figure 1.03. Molecular structure of Rhodamine B.<br />

Azo colorants are released to the environment by many industrial sources as<br />

textile, pharmaceutical, paper and cosmetic. They are very important pollutants because<br />

are very recalcitrant and even at low concentrations can affect the water sources giving<br />

an undesirable color, which reduce the sunlight penetration through the water column<br />

(Mezohegyi et al. 2009). Additionally, another important problem derives from the fact<br />

that their degradation products could have toxic or even mutagenic properties<br />

(Mezohegyi et al. 2009). Some azo dyes are commonly used in the food industries<br />

although some studies reveal that these dyes can cause hyperactivity in children<br />

(Mezohegyi et al. 2009) and, some of them, during the hydrolysis process, can produce<br />

by-products potentially dangerous, including carcinogenic amines (Ozen et al. 2005).<br />

Many research groups have studied the use of biological methods for the degradation of


22<br />

the azo dyes but these processes are normally very slow (Mezohegyi et al. 2009) and<br />

sometimes need red-ox mediators to accelerate the degradation rate.<br />

Azo dyes under reductive conditions could be excised in one of the 22 potencial<br />

carcinogenic aromatic amines, which are categorized as dangerous substances<br />

involving kidney, urinary bladder and liver (Vanhulle et al. 2008). Example of some azo<br />

dyes are the methyl orange, methyl red, phenolphthalein and 1,10-Phenanthroline (Hong<br />

et al. 2009).<br />

Methyl orange (MO) (Figure 1.04) is commonly used as a dye in the textile<br />

industry and in chemistry as an acid-base indicator (Marin et al. 2007). MO is not a<br />

biodegradable substance when is in aqueous solution. The azo dyes possess basically<br />

two aryl groups (benzene rings) connected by the azo group (-N=N-) as a bridge<br />

between the aryl rings; these structures conform the chromophore (Ozen et al. 2005). If<br />

one protic group is conjugate to the azo a tautomer is formed (azo-hydrazine<br />

tautomerism) (Ozen et al. 2005). The hydroxyl radical reactions experienced by azo<br />

compounds include the addition to the aryl ring, hydrogen removal or one-electron<br />

oxidation (Ozen et al. 2005). Bi 2 Fe 4 O 9 nanosheets are another type of photocatalyst for<br />

degradation of MO (Ruan and Zhang 2009).<br />

O S Na 3<br />

N N N CH 3<br />

CH 3<br />

Figure 1.04. Molecular structure of the Methyl<br />

Orange.<br />

Crystal violet dye (Hexamethyl pararosaniline chloride or CV) is part of the<br />

triphenyl methane dye group, commonly named Basic Violet 3, and is used as a DNA<br />

label (Ma et al. 2011), in textile, ball point pens, on artist pallet, in paper industry, as a


23<br />

fungicidal, human anti-parasitic and also in veterinary medicine (Ab<strong>del</strong>malek et al. 2006)<br />

but affect the aquatic life acting as a mutagenic agent because affects the mitotic<br />

process (Pattapu et al. 2008; Confortin et al. [unknown date]). The triphenylmethane<br />

dyes are carcinogenic to animals (Parshetti et al. 2006). Many studies were performed<br />

for the degradation of CV in aqueous media under aerobic conditions (Pattapu et al.<br />

2008). Some authors studied the degradation process of the CV under aerobic<br />

conditions using MnO 2 as a catalyst and they conclude that the degradation process<br />

could be affected by many factors as the presence of possible ions, the catalyst and dye<br />

concentration, pH of the solution and other factors (Pattapu et al. 2008). Also, the<br />

kinetic of degradation for the reaction is a first order (Pattapu et al. 2008).<br />

The CV and the Methyl Violet (MV) have similar structures. The only difference<br />

between both structures is the presence of one NHCH 3 group (in MV) instead of an<br />

CH 2 (CH 3 ) 2 in CV (see Figure 1.05a and Figure 1.05.b).<br />

Figure 1.5. Molecular structure of Crystal Violet (a) and the molecular<br />

structure of Methyl Violet (b).<br />

Aromatic compounds constitute an important source of environmental pollution<br />

reaching the atmosphere and groundwaters because there are wi<strong>del</strong>y used as<br />

intermediates in the production of pesticides, synthetic polymers and dyes (Huang et al.


24<br />

2010a). The presence of these substances in the environment is a concern because<br />

possess carcinogenic, teratogenic and toxic properties (specially the azo dyes),<br />

decrease the light penetration through the water column, and affect aesthetically<br />

(Karadag et al. 2006; Vanhulle et al. 2008; Huang et al. 2010b) damaging the<br />

environment (Dafnopatidou et al. 2007).<br />

The organic pollutant selected for comparison purposes has been the p-amino<br />

benzoic acid (pABA) (see Figure 1.06). The yeast Saccharomyces cerevisiae uses the<br />

p-aminobenzoic acid as a precursor in some biosynthesis processes but in mammalian<br />

cells (human and rats cells) pABA competes with synthesis processes inhibiting the<br />

biosynthesis of some enzymes (Marbois et al. 2010).<br />

Based aminobenzoic acid<br />

compounds are present in clinical, pharmaceutical, anesthetic drug metabolite,<br />

cosmetics, sunscreen products and ammunition waste (Schmidt et al. 1997).<br />

O<br />

OH<br />

NH 2<br />

Figure 1.06. Molecular structure of p-amino<br />

benzoic acid (pABA).<br />

The presence of anthropogenic substances in fresh waters is a concern. Some<br />

authors (Gaulke et al. 2009) studied different methods for treatment of this type of<br />

compounds. Presence of exogenous estrogenic substances, for example, in aquatic<br />

system in concentrations less than 1 ng L -1 affects aquatic species as fishes because in


25<br />

the organism is an endocrine disruptor. The mayor sources of these substances are the<br />

municipal wastewater treatment plants and the operation of animal feeding areas. Other<br />

important contaminants in the aquatic environments as some antibiotics (sulfonamides)<br />

commonly used in the agriculture and p-ABA which is used as a component of<br />

sunscreens, have a very similar structure with a substituent in the C 1 (Neafsey et al.<br />

2010).<br />

Eichenseher (2006) developed different studies that evidence the presence of<br />

compounds used as UV filters in the lipid tissue of fishes. In this study was stated that<br />

these compounds enter into the environment when people use sunscreens swimming in<br />

the rivers or lakes. The UV filters are present in lip balms, sunscreen lotions and many<br />

cosmetic and personal care products (PPCPs) but some of the UV filters compounds are<br />

endocrine disrupter and can alter the reproductive functions of the organism. The 4-<br />

methylbenzylidene camphor (4-MBC) and octocrylene (OC), for example, are UV filters<br />

which can bioaccumulate in the aquatic food chain and are biologically degraded<br />

although they get degraded very slowly in the environment and persist for a long time.<br />

(Eichenseher 2006).<br />

The fundamental research question for this study is to determine which of the<br />

possible degradation processes, including catalytic photodegradation, photo-Fenton or<br />

Sono-Fenton is more efficient for the degradation of organic contaminants dissolved in<br />

water. <strong>To</strong> reach this goal, many objectives should be previously satisfied as: i) the<br />

synthesis of different catalysts as TiO 2 nanowires, TiO 2 @CNTs, ZnO nanoparticles,<br />

Fe 2 O 3 nanowires and magnetite nanoparticles, used in different catalytic processes<br />

(Sono-Fento, Photo-Fento and photocatalysis reactions) to degrade mo<strong>del</strong> compounds<br />

(dyes) as Methylene Blue, Rhodamine B, Methyl Orange, Gential Violet and Methyl<br />

Violet and an organic contaminant as p-aminobenzoic acid, ii) determine the rate of<br />

reaction of the different processes, and iii) finally, to establish which degradation


26<br />

processes (photocatalysis, Sono-Fenton and Photo-Fenton) are more effective for the<br />

degradation of organic compounds as possible alternatives as wastewater treatments.


ChapterTwo<br />

Experimental Techniques<br />

In this chapter we report on the experimental techniques and synthesis methods<br />

used during the development of this experimental work.<br />

An important part will be<br />

focused on the synthesis of the catalysts used for the degradation of organic<br />

compounds. The experimental setup for the irradiation (photo and sono irradiation) of<br />

organic compounds in solution will be explained. Spectroscopy (UV-vis, Fluorescence,<br />

Raman and XPS), <strong>To</strong>tal Organic Carbon (TOC), magnetometry (VSM) and microscopy<br />

(FESEM) techniques used during the present work are included in this chapter. Some of<br />

the instruments described here are located at the School of Science and Technology<br />

(<strong>Universidad</strong> <strong>del</strong> <strong>Turabo</strong>), Autonomous University of Madrid (Spain) and at the<br />

Complutense University of Madrid (Spain). Magnetometry measurements were carried<br />

out with the research group of Prof. Óscar Perales at Department of General<br />

Engineering, University of Puerto Rico in Mayagüez. Raman experiments were<br />

performed at the University of Puerto Rico in Rio Piedras.<br />

2.01. X-Ray Diffraction (XRD)<br />

The X-ray diffraction or XRD allows obtaining relevant information on solid<br />

samples. Among the different information to be obtained by XRD to be mentioned are<br />

the crystalline structure, the averaged particle size, the unit cell dimensions, and the<br />

constituents of the cell.<br />

Additionally, this information can be obtained in situ to<br />

characterize the different transitions (crystallinity, particle size or even the variation of<br />

the chemical composition) during a chemical reaction (Thomas and Thomas 1997). This<br />

technique is based on the principle of the diffraction or dispersion of light waves when an<br />

X-ray beam bombards a sample.<br />

26


27<br />

X-rays are electromagnetic radiation with photon energies typically in the range<br />

of 100 eV - 100 keV. For diffraction applications, only short wavelength X-rays (hard X-<br />

rays) in the range of 10 to 0.01 nanometers (1 - 120 keV) are used. Because the<br />

wavelength of X-rays is comparable to the interatomic spacing, they are ideally suited for<br />

probing the structural arrangement of atoms and molecules in a wide range of materials.<br />

X-rays are generated after bombardment by an electron beam of a stationary or<br />

rotating solid target.<br />

Electrons collide with atoms in the solid target producing a<br />

continuous spectrum of X-rays (Bremsstrahlung radiation). Common solid targets used<br />

in X-ray tubes include Cu and Mo, which emits 8 keV and 14 keV X-rays with<br />

corresponding wavelengths of 1.54 Å and 0.8 Å, respectively. According with the Bragg<br />

theorem, the diffraction pattern is done by the “constructed interference of the waves<br />

scattered from the successive lattice planes in the crystal” and this occurs when the<br />

difference of the path is equal to an integer number of the wavelength (Gersten and<br />

Smith 2001). The Bragg Equation which describes the diffraction process is defined as:<br />

nλ 2dsinθ<br />

In a solid material, the deviation angle Ф is defined as 2Θ, and Θ is the angle<br />

done by the beam with respect to the crystalline plane and d is the distance between<br />

consecutives planes (see Figure 2.01) (Gersten and Smith 2001; Flewitt and Wild 2003).<br />

Using this equation, the crystal spacing can be measured (Gersten and Smith 2001).


28<br />

2ϴ<br />

d<br />

Figure 2.01. Schematic representation of the<br />

diffraction process from atoms in a crystalline<br />

lattice (Adapted from Flewitt et al. 2003).<br />

Diffractograms consist of a plot of reflected intensities against the detector angle<br />

2-theta. In powder samples, all possible diffraction directions of the lattice should be<br />

attained due to the random orientation of the powdered material.<br />

Using β as the full width at half maximum or FWHM of a broad diffraction peak,<br />

the averaged particle sizes can be estimated (Thomas and Thomas 1997; Hadj Salah et<br />

al. 2004; Sridevi and Rajendra 2009) by applying the Scherrer’s equation:<br />

Kλ<br />

β <br />

Dcosθ<br />

where λ is the X-ray wavelength, Θ is the Bragg’s angle and K is the Scherrer constant<br />

that depends on the peak shape (Thomas and Thomas 1997, Hong et al. 2009).<br />

According to Gersten and Smith (2001), there are four different ways available to<br />

perform XRD experiments. The first one is to use a broadband (non monochromatic) X-<br />

ray source and to analyze the back reflection. The second one is the use of a diverging<br />

(noncollimated) X-ray beam. The third one consists of using a non monochromatic


29<br />

source and non collimated beam and in this case the diffraction conditions are reached<br />

during the rotation of the crystal and finally, the fourth way consists of using a<br />

monochromatic X-ray source.<br />

X-ray powder diffraction patterns (XRD) were collected using an X´Pert PRO X-<br />

ray diffractometer (PANalytical, The Netherlands) in Bragg-Brentano goniometer<br />

configuration. X-ray radiation source was a ceramic X-ray diffraction Cu anode tube<br />

type Empyrean of 2.2 kW.<br />

Angular measurements (θ - 2θ) were made with<br />

reproducibility of: ±0.0001 degree, applying steps of 0.05 degrees from 5 to 60 degrees.<br />

Figure 2.02-a shows an image of the XRD diffractometer. A detailed view of the<br />

goniometer is shown in Figure 2.02-b.<br />

Figure 2.02. Images of the PANalytical XRD system used in<br />

this research (a) and detail of the goniometer (b).<br />

2.02. Magnetic Susceptibility<br />

Magnetometry has been wi<strong>del</strong>y used for determining magnetic properties of<br />

materials. This technique is one of the most appropriate to study and characterize<br />

magnetic materials due to the vast information that can be obtained by the hysteresis<br />

cycles. A hysteresis cycle shows the relationship between the induced magnetic flux


30<br />

density (B) and the magnetizing force (H). It is often referred to as the B-H loop. An<br />

example of hysteresis loop is shown in Figure 2.02.<br />

Figure 2.03. Typical hysteresis loop of capped magnetite<br />

nanoparticles (Marquez et al. 2012, discussion about the<br />

results of "Dimensionality effects on the magnetization<br />

processes in magnetite nanoparticles".<br />

This loop is obtained by measuring the magnetic flux of the magnetic material<br />

under scanning of the magnetizing force. Any magnetic material that has never been<br />

previously magnetized or has been thoroughly demagnetized will follow a similar<br />

hysteresis loop. Greater the applied magnetic field (G), stronger the magnetization<br />

observed. At the saturation level almost all of the magnetic domains are aligned and<br />

additional increases in G will produce subtle changes in magnetization.<br />

Vibrating sample magnetometer, VSM, is based on Faraday’s law of magnetic<br />

induction, which states that a changing magnetic flux enclosed by a coil induces a<br />

voltage in that coil.<br />

In this technique, an external magnetic field produces the<br />

magnetization of the sample. Magnetic dipole moments in the sample create a magnetic


31<br />

field around the sample (magnetic stray field). In VSM, the sample is vibrating in the z<br />

direction as a function of time and the stray field is determined as a function of time from<br />

pick up coils and converted into electronic data as a voltage output (see Figure 2.04.).<br />

Figure 2.04. VSM components. Sample<br />

is placed between large diameter poles<br />

for obtaining homogeneous magnetic<br />

fields. The arrow indicates the vibrational<br />

motion of the sample.<br />

Materials may be classified according their magnetic susceptibility to an applied<br />

magnetic field.<br />

A paramagnetic material could be defined as a material which is<br />

attracted toward an external magnetic field. In contrast with this, diamagnetic materials<br />

are repulsive when placed in a magnetic field (Gersten and Smith 2001).<br />

The study of magnetic variability was done using a Lake Shore-7400 vibrating<br />

sample magnetometer (VSM) at room temperature (see Figure 2.05).<br />

This VSM<br />

instrument can attain fields up to 3.1 Tesla in the presence of 3 inch gap between<br />

magnets and the sample rod vibrates at 84 Hz. At room temperature the magnetization<br />

sensitivity is 0.1 μemu and the maximum limit is 1000 emu.


32<br />

Figure 2.05. Lake Shore-7400 Vibrating Sample Magnetometer (VSM).<br />

2.03. Thermogravimetric Analysis (TGA)<br />

Thermogravimetric Analysis (TGA) is an experimental test that is performed on<br />

powder samples for determining changes in weight in relation to change in temperature.<br />

Such analysis requires high degrees of precision in three measurements: weight,<br />

temperature, and temperature change. The weight change observed during a specific<br />

temperature range can be correlated with the composition of the sample and thermal<br />

stability.<br />

This technique is extensively used to determine the composition, thermal<br />

stability, oxidative stability, moisture and volatile content, lifetime and kinetics of<br />

decomposition or dehydration of samples (TA [undate]; Anson Casaos 2005).<br />

Figure 2.06. shows a schematic illustration of a TGA instrument. This instrument<br />

is composed by a sensitive analytical balance, a furnace, a purge gas system, and the<br />

microprocessors to control and display the data. The balance cell is the most important


33<br />

part of any TGA system and consists of a high-precision balance with a pan loaded with<br />

the sample.<br />

The sample is placed within an electrically heated oven with a<br />

thermocouple to accurately measure the variations of temperature.<br />

During the<br />

measurement, the atmosphere in contact with the sample is controlled by flowing pure<br />

nitrogen as inert purge gas. Analysis is carried out by raising the temperature gradually<br />

and plotting weight against temperature.<br />

Microprocessor<br />

Lamp<br />

Photodiode<br />

Thermal Balance<br />

Curve<br />

Thermocuple<br />

Purge gas outlet<br />

Tare pan<br />

Purge gas inlet<br />

Sample holder<br />

Furnace<br />

Figure 2.06. Schematic illustration of the TGA instrument<br />

(Adapted from TA [unknown date]).<br />

This method has the advantage that only a small amount of substance is needed<br />

(around 10-20 mg). Nevertheless, the main disadvantage of the TGA method is the<br />

limited information that can be obtained from this technique, due to the fact that only<br />

information concerning to the lost or gain of weight by the sample is obtained. The<br />

curves obtained during the analysis are just a behavior pattern and not a fingerprint of<br />

the materials because any small change in the parameters as temperature rampage,<br />

purge gas, sample size or even the sample morphology can affect the shape of the<br />

curve (TA [unknown date]).


34<br />

The thermogravimetric analyses were done with a TGA Q-500 instrument (TA<br />

Instruments) under an inert atmosphere of nitrogen. The heating rampage was of 20<br />

°C/min from 100 to 600 °C.<br />

Figure 2.07 shows the instrument used for these<br />

measurements.<br />

Figure 2.07. Thermal Gravimetric Analysis<br />

(TGA), TA instrument, Q500.<br />

2.04. Specific Surface Area (BET)<br />

The study of the surface area of catalysts has a great relevance to determine the<br />

activity of the catalysts because the rate of the product formation can be directly related<br />

to the surface area available. According to Thomas and Thomas (1997), the synthesis<br />

procedure of metal oxide catalysts could have relevant effects on their catalytic<br />

properties due to the different surface area or even the presence of open pore structures<br />

that are appropriate to control the catalytic behavior. The determination of the surface<br />

area is important because can be used to determine catalyst poisoning, thermal<br />

deactivation and other degradation effects over time and also to predict the performance


35<br />

of the catalyst. Three methods are commonly used to determine the surface area; the<br />

volumetric method, the gravimetric method and the dynamic method (Thomas and<br />

Thomas 1997).<br />

The volumetric method was selected to determine the surface area of the<br />

catalysts used in this research.<br />

According to Thomas and Thomas (1997), the<br />

monolayer capacity may be identified either by noting the ordinate value of the volume<br />

(when V is plotted against p) as the isotherm bends over sharply or by applying the<br />

Brunauer-Emmett-Teller (BET) theory (Chandras et al. 2010).<br />

Brunauer, Emmett and Teller derived a theory from a statistical and gas-kinetic<br />

mo<strong>del</strong> based on the principle that the increase of the adsorbate partial pressure over a<br />

dry powder sample corresponds to the increase of multi-layers on the sample surface.<br />

This technique is normally based on the physical adsorption of nitrogen at low<br />

temperature. This technique measures gas uptake (corresponding to the adsorption<br />

process) under increasing the partial pressure of nitrogen in contact with the powder<br />

sample and the release of nitrogen (desorption process) (Garriga I Cabo 2007).<br />

The BET equation is commonly used when the isotherm curve is well defined<br />

(Thomas and Thomas 1997). The equation is defined as:<br />

V(p<br />

p<br />

<br />

1<br />

c 1<br />

<br />

p<br />

0<br />

p) Vmc<br />

Vmc<br />

p0<br />

If<br />

p<br />

V(p is plotted against<br />

0<br />

p)<br />

p (where p<br />

0<br />

is the vapor pressure of the absorbate at<br />

p 0<br />

the adsorption temperature) a straight line is obtained. Using the slope and the intercept<br />

the<br />

V<br />

m<br />

can be finally calculated (Thomas and Thomas 1997; Anson Casaos 2005;<br />

Lopez-Fernandez 2009).<br />

The specific surface areas of the catalysts used in the present research were<br />

determined by the BET method using a Micromeritics ASAP 2020 (Figure 2.08). The


36<br />

micropore volume, WMP [cm 2 /g], was measured using the Barrett-Joyner-Halenda (BJH)<br />

approach (Barret et al. 1951; Marquez et al. 2012).<br />

Figure 2.08. Micromeritics' ASAP 2020<br />

Accelerated Surface Area and<br />

Porosimetry, used in this research.<br />

2.05. Raman Spectroscopy<br />

The most common vibrational spectroscopies are the infrared (IR) and Raman.<br />

Both techniques can be used to assess the molecular motion and to identify species and<br />

functional groups in a sample (Hernandez Rueda 2010).<br />

Raman spectroscopy is a technique based on the Raman Effect, consisting in an<br />

inelastic scattering process discovered in 1928 by the Indian physicist C.V. Raman. In<br />

this process, a monochromatic beam of light is focused onto the sample and the energyshifted<br />

fraction of the scattered light is detected and measured (Schwartz [unknown


37<br />

date]). Raman Effect can be easily explained using an electro-dynamical or a quantummechanical<br />

mo<strong>del</strong>.<br />

According to the electro-dynamical mo<strong>del</strong>, when an electromagnetic radiation<br />

collides with a body, most of the scattered light appears at the same wavelength of the<br />

incident laser. This effect is due to the fact that this incident light does not undergo any<br />

interaction with molecular vibrations of the sample. This scattered light is produced by<br />

an elastic scatter and is called Rayleigh peak. However, an extremely low fraction of the<br />

excitation light may inelastically interact with atomic vibrations, producing the Raman<br />

scattering. During the Raman Effect, the high frequency vibration of the electric field<br />

vector of the laser source induces a time-dependent dipole moment. The interaction<br />

between the dipole moment and the electromagnetic wave is controlled by the<br />

polarizability of the excited molecule. Only those vibrations leading to variations of the<br />

polarizability are responsible for Raman transitions.<br />

From the quantum-mechanical point of view, all molecules are characterized by<br />

having vibrational states with a limited number of allowed discrete energies. When a<br />

molecule in its ground state is excited by an input of energy this molecule is promoted to<br />

an excited vibrational state. Nevertheless, excitation photons can only be absorbed<br />

when their energy is equivalent to the energy difference between two allowed vibrational<br />

levels. This absorption is possible when excitation radiation is in the infrared range. On<br />

the contrary, excitation radiation of higher energies (i.e. visible or ultraviolet) cannot be<br />

absorbed because its energy is much higher than that concerning to vibrational<br />

transitions. Hence, in most cases no interaction occurs and, in this way, the molecule<br />

does not experience vibrational changes producing a scattered peak at the same energy<br />

as the excitation laser (Rayleigh transition). Only with an extremely low probability, the<br />

Raman scattering is observed at higher or lower vibrational state than before the<br />

interaction is reached.


38<br />

A Raman spectrum is a plot of the detected light intensity (usually given in counts<br />

or arbitrary units) as a function of the photon energy (Raman shift). The Rayleigh line is<br />

observed at zero Raman shift.<br />

Anti-Stokes and Stokes Raman bands appear at<br />

negative and positive Raman shifts, respectively. In general, only the most intense<br />

Raman bands (Stokes) are used for characterizing materials (Figure 2.09).<br />

Figure 2.09. Raman vibrational and scattering modes (Adapted from Flewitt<br />

et al. 2003; Hernandez Rueda 2010).<br />

The bands in a Raman spectrum represent the interaction of the incident light<br />

with specific vibrations of the nuclei. These vibrations clearly depend on the sizes,<br />

masses and valences of the atoms, the bond forces and the symmetry of the material<br />

and, for this reason interpretation of Raman spectra provides relevant information about<br />

the sample.<br />

A conventional problem of Raman spectroscopy is the fluorescence emission<br />

that is simultaneously produced by laser beam excitation; this fluorescence emission can<br />

mask the Raman signal.<br />

This problem can be avoided by using Raman excitation<br />

wavelengths in a spectral range that is not affected by the luminescence signal (i.e.<br />

infrared radiation). Other possible artifacts are caused by the increase of the local


39<br />

temperature due to the high power density and very high absorptivity of the sample,<br />

which may result in alteration or decomposition of the sample.<br />

Raman spectroscopy can be used in a vast number of applications including<br />

pharmaceutics, forensic science, polymer science, semiconductor physics, and<br />

chemistry of materials. Raman spectroscopy is highly specific for a certain type of<br />

samples (for example, carbon nanotubes of fullerenes) and for this reason, this<br />

technique is used for the identification and structural characterization of materials.<br />

Raman spectroscopy is a non-destructive technique that can be applied to study solids<br />

as well as liquids (even in aqueous solution) or gases, having the additional advantage<br />

that no special sample preparation is needed.<br />

Figure 2.10. Image of the micro Raman<br />

scattering equipment used in this research.<br />

During this research, the Raman spectra of the catalyst samples were recorded<br />

using an ISA T64000 triple monochromator (Figures 2.10). <strong>To</strong> focus the line (514.5 nm)<br />

of the Coherent Innova 99 Ar + laser and to collect the backscattered radiation an optical<br />

microscope (Olimpus BH2-UMA) with an 80X magnification was used. This microscope<br />

was equipped with a NEC NC-15 camera.<br />

The scattered light dispersed by the


40<br />

spectrophotometer was detected by a charge-coupled device (CCD) cooled with liquid<br />

nitrogen (by using a 2.5 cm CCD and 1800 grooves mm -1 grating, the spectral resolution<br />

obtained was typically less than 1 cm -1 ) (Dixit 2003).<br />

2.06. X-Ray Photoelectron Spectroscopy (XPS)<br />

X-ray photoelectron spectroscopy (XPS), traditionally called ESCA, is a surface<br />

analytical technique which has proved to be extremely useful for the study and<br />

characterization of the oxidation states. Intensities and positions of photoelectron peaks<br />

depend on over-layer thickness, chemical state of near-surface atoms and the<br />

stoichiometry of the over-layer.<br />

This spectroscopic technique is based on the<br />

photoelectric effect, i.e., the ejection of an electron from a core level by an X-ray photon<br />

of energy hv. The sample is irradiated by photons by using an X-ray gun. In the surface<br />

of the sample, photoelectrons (and Auger electrons) are produced.<br />

Energy of the<br />

emitted photoelectrons is then analyzed by an electron detector (normally a<br />

hemispherical analyzer, HSA, operated using a constant pass energy mode) that is<br />

placed near of the sample surface to detect the kinetic energy of the electrons leaving<br />

the sample. Kinetic energy (KE) of the electrons is the experimental quantity measured<br />

by the spectrometer, although this value will depend on the X-ray energy used to<br />

produce them. The binding energy of the electron (BE) is the standard parameter which<br />

identifies an element specifically. The next equation establishes the relation among<br />

these different parameters:<br />

BE hv - KE -W<br />

where hv is the X-ray energy and W is the work function of the spectrometer (Garriga I<br />

Cabo 2007).<br />

XPS spectra provide chemical information on the sample surface (typically 20-<br />

100 Ǻ) depending on the nature of the specimen and the angle of the incident X-ray<br />

beam. Maximum sampling depth is obtained when the sample is perpendicular to the


41<br />

incident X-ray beam. When the incident angle with respect to the surface is very low<br />

(i.e. Φ < 10º) the incident radiation can be exploited to study changes in sample<br />

composition at depths of only some angstroms from the surface. In this way, spectra<br />

can be obtained using different incident angles and then compared to finally study the<br />

homogeneity of the sample with respect to depth.<br />

The binding energies of core<br />

electrons are directly affected by the energy of the valence electrons. Consequently, if<br />

we consider as an example the core electrons of carbon, the binding energy<br />

corresponding to the C1s transition will depend on the bonded atom: C-H (285·0 eV), C-<br />

Br (286·0 eV), C-Cl (286·5 eV) and C-F (287·9 eV). Due to this effect, it is possible to<br />

distinguish among possible environments around specific atoms. In this example, when<br />

electronegative atoms are bonded to carbon, a δ + charge is generated on the carbon<br />

atom. As a result, the carbon atom holds electrons more tightly producing a higher<br />

binding energy than for the case of C bonded to H. Contrarily, the excess of negative<br />

charge on an atom has the opposite effect, making the electrons easier to remove,<br />

lowering their binding energies.<br />

The XPS measurements were performed on both an ESCALAB 210<br />

spectrometer (equipped with a hemispherical analyzer) and on a Perkin–Elmer PHI 3027<br />

spectrometer (equipped with a double-pass cylindrical mirror analyzer), using a nonmonochromatic<br />

Mg K α (1253.6 eV) radiation of a twin-anode (Figure 2.11). In all cases,<br />

the spectra were recorded at 20 mA and 12 kV in the constant analyzer energy mode<br />

using a pass energy (PE) of 50 eV. The samples were previously degassed at the<br />

preparation chamber of the spectrometer for at least 24 hours before the analysis and<br />

the vacuum during the spectroscopic analysis was better than 5x10 -9 mbar. The binding<br />

energies were corrected using the C-C peak component to remove any charging shifts<br />

and deal with the Fermi edge coupling problems. The C-C peak (at 284.6 eV) used as a<br />

reference peak is originated from the environmental contamination with carbon


42<br />

compounds as CO 2 and hydrocarbons (Corma et al. 1997a, Corma et al. 1997b; Arribas<br />

et al. 1999). In the case of samples with copper and with the aim to avoid the X-ray<br />

induced reduction of Cu 2+ to Cu +1 , samples were maintained at 173 K during the spectral<br />

acquisition and the X-ray power was limited to 200 W (20 mA–10 kV). The spectral<br />

acquisition time was also reduced to the maximum to prevent the damage of the<br />

samples and the possible reduction of Cu 2+ to Cu 1+ .<br />

Figure 2.11. Perkin–Elmer PHI 3027 spectrometer (a), and VG Escalab<br />

210 spectrometer (b) used in this research.<br />

2.07. Field Emission Scanning Microscopy (FE-SEM)<br />

Although optical microscopy is the most conventional and simple solid state<br />

materials characterization technique, this microscopy is clearly limited in its resolution by<br />

the wavelength of light.<br />

This technique uses visible light with wavelengths varying<br />

between 400 and 700 nanometers.<br />

In most optical microscopes, the presence of<br />

spherical aberration limits the resolution to several micrometers. Distinct from optical<br />

microscopy, the images obtained using scanning electron microscopy are generated by<br />

electrons (Garriga I Cabo 2007) instead of visible light and for this reason the resolution<br />

of this microscopy is limited by the wavelength of electrons (as an example, using a


43<br />

standard energy of 5 keV the theoretical resolution is 0.55 nm). Nevertheless, and as<br />

occurs in optical microscopy, the presence of other limiting factors (lens aberration or<br />

astigmatism) is responsible for the decrease of the theoretical resolution to values on the<br />

order of a few nanometers.<br />

Electrons in a SEM carry significant amounts of kinetic energy, and this energy is<br />

dissipated in a variety of events produced by different interactions between the electron<br />

beam and the sample. These events include secondary electrons (SE) responsible for<br />

the SEM images, backscattered electrons (BSE) (responsible for SEM images with<br />

relevant information concerning the chemical nature of the sample), diffracted<br />

backscattered electrons (DBE) used to obtain similar information to that obtained using<br />

X-ray diffraction, photons that are used for elemental analysis, visible light<br />

(cathodoluminescence), and heat. SE and BSE are the most conventional electron<br />

emission techniques used for imaging samples. SE is mainly used to characterize the<br />

morphology of solid samples and BSE is most valuable to characterize differences in<br />

chemical composition (see Figure 2.12). The scheme of a scanning electron microscope<br />

(Field emission SEM) is shown in Figure 2.12.<br />

In this research, Field emission scanning electron microscopy (FE-SEM) images<br />

were obtained using a JEOL JM-6400 microscope. The microscope is a high-resolution<br />

FE-SEM. It can provide beam voltages ranging from 0.2kV to 40 kV and beam currents<br />

from 10 picoamps to 10 microamps. This instrument offers high performance and low<br />

noise at low accelerating voltages.<br />

Resolution of ca. 3 nm is attainable, and<br />

magnifications can be obtained ranging from 10 X to 300,000 X. The cathode is a highbrightness<br />

lanthanum hexaboride (LaB 6 ) source. The SEM is equipped with two-inch<br />

and four-inch airlocks and a Faraday cup for beam current measurements. The sample<br />

stage is computer-driven. Figure 2.13 shows an image of the FE-SEM instrument used<br />

in this research.


44<br />

Cold Cathode Field Emitter<br />

Image<br />

Anodes<br />

Digital<br />

Processor<br />

Electromagnetic<br />

Lenses<br />

Electron detector<br />

(Scincillator)<br />

Sample holder and<br />

Sample<br />

Figure 2.12. Main components of a FE-SEM instrument<br />

(Adapted from Flewitt et al. 2003; NMT Materials Dept<br />

2012).<br />

Figure 2.13. FE-SEM JEOL JM-6400 microscope<br />

(“Centro de Microscopía Luis Bru” at the<br />

Complutense University of Madrid, Spain).


45<br />

2.08. <strong>To</strong>tal Organic Carbon Analysis<br />

Carbon content is one of the most relevant parameters measured in different<br />

types of solutions, including drinking water, industrial wastewater, etc. Carbon analyzers<br />

are instruments devoted to the analysis of organic, inorganic and total carbon content in<br />

these water or liquid solutions. The method is based on the oxidation of the carbon<br />

based compounds to finally produce CO 2 . During the oxidation process, potassium<br />

persulfate, in presence of UV irradiation, initiates a quick reaction to oxidize the<br />

compounds. Sulfate ions and hydroxyl groups act as free radicals reacting with the<br />

organic compounds. The CO 2 produced during the oxidation is carried by the nitrogen<br />

gas to the nondispersive infrared (NDIR) detector and the signal is produced. The<br />

carbon concentration is expressed in mg L -1 , parts per million or ppm.<br />

For a typical analysis, a 10 mL sample is diluted to 40 mL in carbon-free distilled<br />

water.<br />

The sample is taken by an automatic syringe and read in triplicates for<br />

reproducibility. A duplicate and an internal standard were used during each analysis to<br />

standardize the analysis procedure. The method used in this research was the total<br />

organic carbon analysis (TOC) method with a range of 0.01 – 20 ppm C.<br />

The equipments used to determine the TOC concentration were both a Tekmar<br />

Dohomann, Phoenix 8000 UV-Persulfate TOC Analyzer (Figure 2.14) and a Leco<br />

CHNS-932 (Figure 2.15).<br />

This last instrument is commonly used to determine the<br />

carbon, hydrogen, sulfur and oxygen concentrations. The Leco CHNS-932 allows the<br />

detection of carbon in a large concentration range (0.002 to 100%), with a precision of<br />

±0.001. For both instruments used in the present research, the detection method is<br />

based on highly selective, infrared detection systems. The instruments used in this<br />

research can only measure dissolved organic carbon (DOC) (Garriga I Cabo 2007). The<br />

suspended solids in the sample have to be previously removed before the injection into


46<br />

the analyzer and, for this reason, 0.22 μm pore size PTFE syringe-driven filters were<br />

used. After filtration, samples were directly injected and analyzed.<br />

Figure 2.14. TOC analyzer (Tekmar Dohrmann<br />

Phoenix 8000 UV-Persulfate TOC Analyzer).<br />

Figure 2.15. Image of the Leco CHNS 932 analyzer<br />

(a) and scheme of its different component parts (b).<br />

2.09. UV-Visible Spectroscopy<br />

The UV-Visible spectroscopy is commonly used due to its simplicity, versatility,<br />

accuracy and cost-effectiveness.<br />

UV-Visible wavelengths cover a range from


47<br />

approximately 10 nm (far UV irradiation) to 780 nm (visible irradiation). These energies<br />

are sufficient to promote or excite a molecular electron to a higher energy orbital. For<br />

this reason, absorption spectroscopy carried out in this region is also called "electron<br />

spectroscopy". Figure 2.19 shows the different types of electronic transitions that may<br />

occur in organic molecules.<br />

The energy of a photon is defined as:<br />

hc<br />

E <br />

λ<br />

Where h is the Planck’s constant, c is the speed of light in a vacuum and λ is the<br />

wavelength. According to this equation, the energy of the photon decrease when the<br />

wavelength increases. A photon energetically higher is necessary to excite a molecule<br />

and promote an electron to another quantum state.<br />

In a simple way, when a sample is irradiated by using a white light, this irradiation<br />

could be totally reflected and in this case the sample looks white but if all irradiation is<br />

totally absorbed the sample, in this case, looks black. Meanwhile, when only a portion of<br />

the irradiation is absorbed and the remaining portion is reflected, the sample shows<br />

different color. The color observed is the portion of the light reflected; a complementary<br />

wavelength of the absorbed irradiation wavelength. Non-colored samples do not show<br />

absorption spectrum in the UV-visible range, but can absorb in the IR portion of the<br />

spectrum.<br />

The atomic structure and the presence of color in a sample are closely related<br />

because an electronic transition is necessary for the occurrence of the absorption. The<br />

electronic promotion can occur from the ground state to different excited states. The<br />

possible transitions can involve different orbitals (i.e. σ, π, n, σ* and π*) arising in<br />

different electronic transitions (σ → σ*, n → σ*, n → π* and π → π*) (Figure 2.16). The<br />

transition to the first excited state associated with the HOMO (highest occupied


Incident<br />

Radiation<br />

48<br />

molecular orbital) - LUMO (lowest unoccupied molecular orbital) excitation, is normally<br />

characterized by having low energy and high intensity.<br />

σ*<br />

π*<br />

Possible<br />

Excited<br />

States<br />

η<br />

(η→π*)<br />

(η→π*)<br />

π<br />

σ<br />

(σ→σ*)<br />

(π→π*)<br />

Ground<br />

States<br />

Figure 2.16. Representation of the different electronic<br />

transitions generated during the absorption process<br />

under UV-Visible irradiation.<br />

Molecular groups with conjugate insaturations produce a high effect in the<br />

molecular absorption, increasing the λ max and the intensity of the peaks on the<br />

absorption spectrum. The presence of chromophores (color-bearing molecular features)<br />

which are functional groups not conjugated to other groups (i.e. nitro, azo, azo-amine,<br />

carbonyl, etc) and auxochromes, such as OH, NH 2 , CH 3 and NO 2 , have been suggested<br />

to be responsible for important changes in the absorption spectrum. Other relevant<br />

factors that can affect the absorption properties of a UV-Visible spectrum are the<br />

presence of steric effects and the solvent used during the analysis.<br />

concentration:<br />

The equation of Beer-Lambert correlates the absorption of a substance with the<br />

A εcλ


49<br />

Where ε is the molar absortivity, c is the concentration and λ is the wavelength. This<br />

equation has the disadvantage that is true only for monochromatic light and if the<br />

physical and chemical properties of the substance do not change with the change in<br />

concentration.<br />

Finally, during the analysis of a sample using a UV-vis spectrophotometer, the<br />

pass of monochromatic light through the cell and the intensity of the transmitted light<br />

depend both on the pathlength of the cell and the concentration. Transmittance is<br />

defined as:<br />

I 100<br />

I I ;A log<br />

10<br />

log<br />

10<br />

I <br />

<br />

0 T<br />

0<br />

<br />

<strong>To</strong> characterize the absorption properties of our samples and to study the<br />

catalytic degradation of the organic compounds we have used a UV-vis CARY 3 Varian<br />

spectrophotometer.<br />

2.10. Fluorescence Spectroscopy<br />

The Fluorescence Spectroscopy is an important technique that has been used to<br />

determine the degradation process of the organic compounds under different catalytic<br />

processes studied along this research. This technique is complementary to the UV-vis<br />

absorption technique.<br />

This spectroscopy is based on the study of the different transitions between the<br />

first excited singlet state and the ground state. Molecules in the ground state can be<br />

excited by absorption of an appropriate wavelength photon, reaching different excited<br />

states (S 1 *, S 2 *, etc). Two different mechanisms can be observed during desexcitation<br />

process: i) molecules in higher excited states (i.e. S 2 *, S 3 *, etc) experience a rapid non<br />

radiative internal conversion from these excited states to the S 1 *, and ii) molecules in the<br />

first excited singlet state (S 1 *) experience a radiative desexcitation to the ground state,<br />

namely fluorescence. Therefore, the fluorescence is a mechanism to relax an excited<br />

τcl


50<br />

molecule or atom to reach the ground state by emitting a photon. Nevertheless, there<br />

are different alternatives to the light emission consisting in several radiationless<br />

deactivation pathways from the S 1 * state. Among these deactivation processes are the<br />

intramolecular internal conversion (S 1 * →S 0 ), the intersystem crossing (S 1 * →T n ), as well<br />

as the collisional quenching or resonance energy transfer, the most relevant [Albers et<br />

al. 2003].<br />

Frequently, the fluorescence bands are composed by bands at longer<br />

wavelengths than the excitation wavelength.<br />

This process is known as Stokes<br />

displacement but when no change in the wavelength occurs, the process is known as<br />

fluorescence resonance.<br />

If the energy of absorbent during the absorption process,<br />

represented in the excitation spectrum, is similar to the energy released during the<br />

fluorescence process, both spectra are mirror images of each other. In this case, both<br />

spectra are largely overlapping and the resonance line corresponds to the wavelength at<br />

which these spectra cross each other. The fluorescence of a compound is affected by<br />

the quantum yield, which depends on temperature, solvent polarity, molecular structure,<br />

pH and concentration.<br />

In this research, the samples have been characterized by fluorescence<br />

spectroscopy to determine the possible degradation of the organic compound according<br />

with the intensity of the maximum fluorescence emission.<br />

The fluorescence<br />

spectroscopy analysis was performed at room temperature on a Varian Cary Eclipse,<br />

using quartz cells of pathlength 1 cm (Figure 2.17).


51<br />

b)<br />

Polychromatic<br />

UV-Vis Source<br />

Exclusion<br />

Monochromator<br />

Sample<br />

a)<br />

90<br />

Emission<br />

Monochromator<br />

Detector<br />

Figure 2.17. Image of the fluorescence<br />

spectrophotometer Varian Cary Eclipse (a) and a<br />

diagram of a fluorescence spectrometer (Adapted from<br />

Albers et al. 2003).


Chapter Three<br />

Synthesis Procedures<br />

3.1. Synthesis of Titanium Oxide Nanowires<br />

TiO 2 nanowires (TiO 2 NWs) have been synthesized by a novel catalyst-free<br />

hydrothermal procedure. Uniform and size controllable TiO 2 NWs have been obtained<br />

by crystallization of the precursor in acid solution at high pressure and temperature. For<br />

a typical synthesis, 75 mL of concentrated hydrochloric acid (Fisher Scientific, 35%) and<br />

75 mL of DDW (Milli Q) were mixed in a 200 mL Erlenmeyer flask. After the solution has<br />

cooled down to room temperature, 5 mL of the titanium precursor (titanium tetrachloride,<br />

Aldrich Chemical) is added by dripping under agitation at room temperature.<br />

The<br />

mixture was magnetically stirred until all solid particles were dissolved and the material<br />

had a uniform color (approximately 10 min). After that, the solution was placed in 30 ml<br />

Teflon-lined stainless steel autoclaves. Next, flat glass substrates of ca. 15 x 15 mm<br />

(previously cleaned with isopropyl alcohol in an ultrasound bath for 5 min) were<br />

introduced inside the autoclaves, in contact with the acid solution. Autoclaves were<br />

maintained at 150ºC by 4 hours. After that, the autoclaves were left to cool down to<br />

room temperature. The resulting TiO 2 NWs grown on the surface of the glass substrates<br />

were washed at least 5 times with DDW and dried overnight at 60 °C. After drying, the<br />

TiO 2 NWs were separated from the glass substrates and homogeneously pulverized to<br />

facilitate the use in the catalytic tests. Finally, samples were transferred and stored in<br />

sealed vials at room temperature.<br />

52


53<br />

3.2. Synthesis of Zinc Oxide<br />

The synthesis procedure used for obtaining zinc oxide nanoparticles is based on<br />

the procedure described by (Behnajady et al. 2011). In a typical synthesis, 0.2 mol of<br />

Zn(CH 3 COO) 2 (Aldrich, 98+% ACS Reagent) and 0.2 mol of NaOH (Fisher Scientific,<br />

97+% ACS Reagent) were previously dissolved in a few milliliters of water and<br />

subsequently added to a 200 mL Erlenmeyer flask. After that, 100 mL of ethanol (Acros<br />

Organic, 95%) were added to the mixture. The solution was magnetically stirred at room<br />

temperature for approximately 2 hr. The synthesized ZnO nanoparticles were separated<br />

from the solution by centrifugation (7000 rpm) for 10 min, washed five times with<br />

ultrapure water. Next, the powder was dried overnight at 60˚C and maintained in sealed<br />

containers before characterization.<br />

3.3. Synthesis of Titanium Oxide@Multiwalled Carbon Nanotubes<br />

The synthesis of the multiwalled carbon nanotubes covered with titanium oxide in<br />

rutile phase consists principally of two steps. The first one is the modification of the<br />

carbon nanotubes to created actives sites (OH - groups) on the surface of the material.<br />

The second one is the synthesis of the titanium oxide and the incorporation of the<br />

material in the actives sites previously generated on the surface of the carbon<br />

nanotubes.<br />

3.3.1. Carbon Nanotubes Modification<br />

Commercial multiwalled carbon nanotubes, MWNTs Cheap-tubes (95wt%) with<br />

30 – 50 nm OD were modified to be used as support. In a typical synthesis, 5 g of<br />

MWNTs were refluxed in concentrated nitric acid at 100 ºC for 24 hrs. After that, the<br />

nanotubes were separated by centrifugation (6000 rpm, 10 min) and washed repeatedly<br />

with DDW (Milli Q) until the pH rise to neutral. The nanotubes were dried at 60 °C and<br />

maintained in sealed containers.


54<br />

3.3.2. Synthesis and Incorporation of the Titanium Oxide on the MWCNT<br />

The synthesis procedure mainly consists of the incorporation of a titanium oxide<br />

precursor to an acid solution and the subsequent use of high pressure and temperature.<br />

The synthesis of the TiO 2 nanoparticles is produced on the surface of the modified<br />

nanotubes. This growth is initiated exclusively on the OH groups generated during the<br />

acid treatment of the carbon nanotubes, producing small particles whose dimensionality<br />

will depends on the amount of titanium oxide precursor introduced into the reaction<br />

mixture. In a fume hood, 75 mL of concentrated hydrochloric acid (Fisher Scientific,<br />

35%) and 75 mL of DDW (Milli Q) are mixed and magnetically stirred in an Erlenmeyer<br />

flask. When the reaction mixture cool down, 5 mL of the titanium oxide precursor<br />

(titanium tetrachloride, Aldrich Chemical) were carefully added by dropwise. The mixture<br />

was magnetically stirred until any solid particle was observed (approximately 10 min).<br />

<strong>To</strong> synthesize TiO 2 nanoparticles on the MWNTs surface, 0.5 g of the chemically<br />

modified MWNTs were added to this reaction mixture and the solution was magnetically<br />

stirred for 30 min. Next, this solution was transferred to 30 ml Teflon-lined autoclaves.<br />

The autoclaves were closed and introduced in an oven for 4 hours at 150 °C. After<br />

cooling down, the synthesized material, namely TiO 2 @MWNTs, was washed with DDW<br />

(Milli Q) for at least 5 times and finally washed with ethanol. The product was dried<br />

overnight at 60 °C.<br />

3.4. Synthesis of Capped Magnetite Nanoparticles<br />

Linoleic acid capped magnetite nanoparticles were obtained by following a<br />

method previously published in our research group.<br />

Magnetite nanoparticles were<br />

obtained by hydrothermal reaction of (NH 4 ) 2 Fe(SO 4 ) 2 6H 2 O (Fisher Scientific, ACS<br />

Certified) in the presence of KOH in water solution. Ammonium iron sulphate (6 mmol)<br />

was dissolved in 150 mL of distilled water (Milli Q) and this solution was added to 40<br />

mmol of KOH (Aldrich, 99.99%) in 25 mL of distilled water.<br />

Next, ammonium


55<br />

peroxodisulfate (5 mmol) dissolved in 25 mL of water was added to a mixture formed by<br />

toluene (Fisher Scientific, ACS Certified) and isopropyl alcohol (Acros Chemical, 99.8%,<br />

HPLC) (4:1, v/v) with 3 mL of linoleic acid (Fisher Scientific, NF/FCC). This solution was<br />

added to the reaction mixture. The reaction was carried out under reflux for 8 hours.<br />

Then, the reaction mixture was cooled to room temperature and the organic phase<br />

containing the solubilized Fe 3 O 4 nanoparticles was separated. Fe 3 O 4 nanoparticles were<br />

precipitated with ethanol followed by centrifugation. The magnetite nanoparticles were<br />

cleaned with nitric acid (Fisher Scientific) (1 M) and subsequently washed with distilled<br />

water an ethanol, and dried overnight at 50 ºC.<br />

3.5. Synthesis of Iron Oxide Nanowires<br />

Highly crystalline iron oxide nanowires were synthesized by a simple catalystfree<br />

growth procedure. For the synthesis of the iron oxide nanowires a Chemical Vapor<br />

Deposition (CVD) system was used (Figure 3.01).<br />

The pure iron substrates<br />

(Goodfellow, 99.999%) were thermally treated inside a quart tube furnace at<br />

temperatures ranging from 400 to 600 °C in a controlled atmosphere (vacuum and<br />

oxidative/reductive atmosphere). A thermal rampage (400 to 600 °C) was used for the<br />

synthesis (Bonilla et al. 2011).


56<br />

a) b)<br />

c)<br />

Figure 3.01. Image of the CVD system (a and b) and scheme of the CVD<br />

system and thermal treatment (c) used for the synthesis of Fe 2 O 3<br />

nanowires (Adapted from Bonilla et al. 2011).


Chapter Four<br />

Material Characterization<br />

The full characterization of the catalysts used for the different reactions<br />

(photocatalysis, sono-Fenton and photo-Fenton) is described in this section, including<br />

the synthesized and commercial catalysts. As shown in earlier chapters, a variety of<br />

analytical techniques have been used for the characterization of the materials, including<br />

FE-SEM, XRD, TG and Raman. The analysis by magnetometry was only applied to<br />

samples with iron in different oxidation states.<br />

4.1. Photocatalysis<br />

4.1.1. Titanium Oxide (TiO 2 , Rutile Phase)<br />

The use of titanium oxide as a photocatalyst is very common because it is<br />

nontoxic, photostable and has a high oxidant power, but its activity is limited to the UV<br />

region of the spectrum (Velegraki and Mantzavinus 2008; Yu et al. 2009). According to<br />

Hernandez-Enriquez et al. (2008), the specific area of the titanium oxide is related with<br />

the quantity of acid used during the synthesis procedure. According to Hernandez-<br />

Enriquez et al. (2008) the efficiency of the titanium oxide in the photocatalytic reactions<br />

is due to both the specific area of the material and stability of the crystalline phase.<br />

The titanium oxide (rutile phase) is a commercial catalyst (Alfa Aesar, 97%). The<br />

rutile phase has a tetragonal structure with six oxygen atoms around octahedral<br />

arrangement (Garriga I Cabo 2007; Bae et al. 2009). The specific surface area, as<br />

determined by the BET method, was 41 m 2 g -1 . According with the information obtained<br />

from FE-SEM micrographs (50 000x and 20 kV), the particles of rutile were smaller than<br />

1 μm (Figure 4.01) and apparently do not show porosity.<br />

57


Weight Loss (%)<br />

58<br />

Figure 4.01. FE-SEM image of the titanium oxide<br />

(rutile phase) at a magnification of 50 000x.<br />

The TG analysis confirmed the information of the SEM. In TiO 2 (Rutile) only 11%<br />

of weight lost was observed (Figure 4.02). The weight lost is a two-steps process, the<br />

first step is due the absorbed water and the second step could be attributed to the<br />

removal of the hydroxyl groups of the titanium oxide (Niederberger et al. 2002).<br />

100<br />

TiO 2<br />

-Rutile<br />

98<br />

96<br />

94<br />

Weight Loss-11%<br />

92<br />

90<br />

0 100 200 300 400<br />

Temperature ( o C)<br />

Figure 4.02. TGA scan of titanium oxide (rutile<br />

phase).


Intensity (a.u.)<br />

59<br />

In the Raman spectrum (Figure 4.03) two characteristics peaks of titanium oxide<br />

were present (at 450.83 and 608.83 cm -1 ) corresponding to the rutile phase (Jackson<br />

2004).<br />

450.83<br />

608.83<br />

TiO 2<br />

-Rutile<br />

200 400 600 800 1000 1200<br />

Raman Shift (cm -1 )<br />

Figure 4.03.Raman spectrum of TiO 2 sample<br />

(rutile phase).<br />

This catalyst was also characterized by XPS. A very intense peak at 530.01 eV<br />

was observed in the XPS spectrum of the rutile sample (Figure 4.04) (Fundamental XPS<br />

Data 1999). This peak has been associated with the oxygen atoms in the lattice (Liu et<br />

al. 2008) of the TiO 2 . Figure 4.05 shows an intense band at 457.26 eV that has been<br />

unambiguously ascribed to Ti2p (1/2) corresponding to Ti 4+ ions of the crystalline lattice.<br />

This catalyst was also characterized by XRD. The diffractograms obtained for<br />

this sample (rutile phase) show reflections at 27 °, 36 °, 41 °, 44 ° and 57 ° (Figure<br />

4.06b) (Hernandez Enriquez et al. 2008).


CPS<br />

CPS<br />

60<br />

530.01<br />

(O1s)<br />

O1s<br />

TiO 2<br />

-Rutile<br />

534 533 532 531 530 529 528 527 526<br />

Binding Energy (eV)<br />

Figure 4.04. XPS spectrum corresponding to<br />

the O1s region of the TiO 2 catalyst (rutile<br />

phase).<br />

457.26<br />

(2p 1/2<br />

)<br />

463.12<br />

(2p 3/2<br />

)<br />

475 470 465 460 455 450<br />

Binding Energy (eV)<br />

Figure 4.05. XPS spectrum corresponding to the<br />

Ti2p region of the TiO 2 catalyst (rutile phase).


Intensity (cps)<br />

61<br />

TiO 2<br />

NWs<br />

(110)<br />

(101)<br />

(111)<br />

(210)<br />

(211)<br />

(002)<br />

(220)<br />

(d)<br />

(c)<br />

TiO 2<br />

@MWCNTs<br />

TiO 2<br />

-Rutile<br />

TiO 2<br />

-Anatase<br />

(004)<br />

(b)<br />

(101)<br />

(200)<br />

(105) (211)(116)<br />

0 10 20 30 40 50 60 70<br />

2 Theta (Degree)<br />

(a)<br />

Figure 4.06. XRD diffraction patterns for TiO 2 -Anatase (a), TiO 2 -Rutile (b),<br />

TiO 2 NWS (c) and TiO 2 @MWCNTs (d).<br />

4.1.2. Titanium Oxide (TiO 2 , Anatase Phase)<br />

The titanium oxide (anatase phase) is a commercial catalyst. Some drawbacks<br />

of the titanium oxide include the strong absorption capacity of the pollutant or the<br />

intermediate on the actives sites and the fact that the optimal irradiation for anatase is<br />

shorter than 387 nm (Ma et al. 2007; Rahmani et al. 2008). The anatase phase has a<br />

tetragonal structure; six oxygen atoms around one titanium atom in an octahedral<br />

structure (Bae et al. 2009, Garriga I Cabo 2007) with a band gap energy of 3.2 eV<br />

(Rahmani et al. 2008). Anatase is characterized by having a high photoactivity, optimum<br />

band gap, and additionally, this catalyst is easy to synthesize (Vinu and Madras 2009).<br />

The specific surface area (S area ), as determined using the BET method, was 48<br />

m 2 g -1 . The FE-SEM images of the anatase catalyst (Figure 4.07) shows the small size<br />

of the particles (less than 1 μm) and the presence of small aggregates. Smaller particles<br />

or the presence of additional porous structure could not be observed. According with the


Weight Loss (%)<br />

62<br />

TG analysis, the aggregates of the anatase particles could loss approximately the 27.8%<br />

of their weight (Figure 4.08).<br />

This weight lost could be due to the water removal,<br />

possibly water molecules absorbed on the particle surface.<br />

Figure 4.07. FE-SEM image of the titanium<br />

oxide (anatase phase) at a magnification of<br />

50 000x.<br />

100<br />

TiO 2<br />

- Anatase<br />

95<br />

90<br />

85<br />

80<br />

Weight Loss 27.84%<br />

75<br />

70<br />

0 100 200 300 400 500<br />

Temperature ( o C)<br />

Figure 4.08. TGA scan of titanium oxide<br />

(anatase phase).<br />

The nondestructive technique of Raman was applied to elucidate the<br />

characteristics of the materials (Zhou et al. 2006). In the Raman spectra (Figure 4.09) is


Intensity (a.u.)<br />

63<br />

clearly identified the most characteristics peaks of the anatase phase (393.36, 512.59<br />

and 638.05 cm -1 ) revealing that no other phases were present (Jackson 2004).<br />

TiO 2<br />

-Anatase<br />

638.05<br />

512.59<br />

393.36<br />

0 200 400 600 800 1000 1200<br />

Raman Shift (cm -1 )<br />

Figure 4.09.<br />

Raman spectrum of titanium<br />

oxide catalyst (anatase phase).<br />

Anatase was also characterized by XPS. At 457.52 eV was observed the most<br />

characteristic peak of this catalyst that was ascribed to Ti2p (3/2) (Figure 4.10). The peak<br />

observed at 463.26 eV was ascribed to Ti2p (1/2) (Fundamental XPS Data 1999). The<br />

Figure 4.11 shows the XPS spectrum corresponding to the O1s. As can be seen there,<br />

this peak appears at 528.80 eV and has been ascribed to the oxygen atoms in the lattice<br />

(Liu et al. 2008) of anatase (Fundamental XPS Data 1999).


CPS<br />

CPS<br />

64<br />

Ti2p<br />

457.52<br />

(2sp 1/2<br />

)<br />

TiO 2<br />

-Anatase<br />

463.26<br />

(2p3/2)<br />

470 465 460 455<br />

450<br />

Binding Energy (eV)<br />

Figure 4.10.<br />

XPS spectrum of Ti2p<br />

peak on titanium oxide (anatase phase).<br />

529.99<br />

(O1s)<br />

O1s TiO 2<br />

-Anatase<br />

534 532 530 528 526<br />

Binding Energy (eV)<br />

Figure 4.11. XPS spectrum of TiO 2<br />

showing the O1s transition (anatase<br />

phase).<br />

The XRD pattern (Figure 4.06) of anatase was characterized by having different<br />

peaks at 30.93°, 36.44°, 42.77°, 53.73°, 56.72° and 62.65°, corresponding to (101),<br />

(004), (200), (105), (211) and (116) reflections, respectively (Chowdhury et al. 2005).<br />

According to Hernandez Enriquez et al. (2008) the diffraction peaks that characterize the<br />

tetragonal phase of anatase are: 25 °, 37 °, 48 °, 54 °, 55 °, 62 °, 71 ° and 75 °.


65<br />

4.1.3. Titanium Oxide Nanowires<br />

The titanium oxide nanowires (TiO 2 NWs) were synthesized according with the<br />

procedure described in Chapter 3 - Material Synthesis. Figure 4.12 shows different<br />

images obtained by FE-SEM of this catalyst. The wires are composed by smaller wires<br />

of nanometric dimensions (Figure 4.12). The specific surface area (S area ), as determined<br />

by the BET method, was 480 m 2 g -1 . This value is unexpectedly high and could have<br />

relevant effects on the catalytic properties of this material.<br />

Figure 4.12. FE-SEM images of the as-synthesized<br />

TiO 2 NWs at different magnification: 5000x (a), 10 000x (b),<br />

25 000x (c) and 150 000x (d).<br />

Figure 4.13 shows the TG analysis of the as-synthesized TiO 2 NWs. Only a<br />

weight loss of 5.65% was observed during the heating process, indicating the compact<br />

and non-porous structure of the nanowires.


Weight Loss (%)<br />

66<br />

TiO 2<br />

NWs<br />

100<br />

90<br />

Weight Loss 5.65%<br />

0 50 100 150 200 250 300 350 400<br />

Temperature ( o C)<br />

Figure 4.13.<br />

TGA analysis of the assynthesized<br />

TiO 2 NWs.<br />

Raman spectrum of TiO 2 NWs (Figure 4.14), is characterized by having two<br />

peaks at 440.83 cm -1 and 604.73 cm -1 , respectively, that have been ascribed to titanium<br />

oxide as rutile phase (Jackson 2004).<br />

TiO 2 NWs were also characterized by XPS. The obtained XPS spectra were<br />

similar to those previously obtained for TiO 2 as rutile or anatase phase. The peaks<br />

observed at 457.39 eV and 462.98 eV were associated to the Ti2p (3/2) and Ti2p (1/2)<br />

transitions, respectively (Figure 4.15) (Fundamental XPS Data 1999). The XPS peak<br />

corresponding to O1s was observed at 529.80 eV and it was assigned to the oxygen<br />

atoms in the lattice (Figure 4.16) (Liu et al. 2008) (Fundamental XPS Data 1999). As can<br />

be seen there, this peak is not symmetric and could be deconvolved in two components.<br />

An additional peak could appear at ca. 532 eV and could be ascribed to the presence of<br />

CO 2 and other species adsorbed on the surface of the TiO 2 NWs.


CPS<br />

Intensity (a.u.)<br />

67<br />

440.83<br />

604.73<br />

TiO 2<br />

NWs<br />

200 400 600 800 1000 1200<br />

Raman Shift (cm -1 )<br />

Figure 4.14.<br />

Raman spectrum of the assynthesized<br />

TiO 2 NWs.<br />

457.39<br />

(2p 1/2<br />

)<br />

TiO 2<br />

NWs<br />

462.98<br />

(2p 3/2<br />

)<br />

465<br />

460<br />

455<br />

Binding Energy (eV)<br />

450<br />

Figure 4.15. XPS spectrum of Ti2p region of<br />

the as-synthesized TiO 2 NWs.


CPS<br />

68<br />

529.80<br />

(O1s)<br />

O1s TiO 2<br />

NWs<br />

540 538 536 534 532 530 528 526 524<br />

Binding Energy (eV)<br />

Figure 4.16. XPS spectrum of O1s region<br />

of the as-synthesized TiO 2 NWs.<br />

XRD diffraction pattern of the titanium oxide nanowires synthesized and used as<br />

catalyst is shown in Figure 4.06. As can be seen there, the rutile phase with reflections<br />

at 27 °, 36 °, 41 °, 44 ° and 57 ° is the only crystalline phase observed in this catalyst<br />

(Hernandez Enriquez et al. 2008; Cotto et al. 2011). The narrow sharps peaks indicate<br />

the crystalline structure of the nanowires (Li and Liu 2010).<br />

4.1.4. Titanium Oxide @Multiwalled Carbon Nanotubes<br />

The principal forms are: vitreous carbon, carbines, fullerenes and nanotubes<br />

(Ansón-Casaos 2005). In this research, the multiwalled carbon nanotubes were coated<br />

with particles of titanium oxide in rutile phase (TiO 2 @MWCNTs) (Figure 4.17). The<br />

synthesis of this material has been carried out according to the experimental procedure<br />

described in Chapter 3 – Materials Synthesis. Other forms to prepare the TiO 2 @CNTs<br />

include different techniques as, for instance, the electrospray deposition (Doi et al.<br />

2009). The starting material, namely MWCNTs, is forming small clusters or aggregates<br />

whose dimensions can be reduced by treating in ultrasound baths (Bal 2010). After<br />

functionalization treatments of the CNTs the authors observed reduction in the average<br />

length, sidewall disordering and extensive debundling (Wang et al. 2006). The use of


69<br />

MWCNTs as supporting catalyst structure is relevant because some investigations<br />

reveal the flow of photogenerated electrons from the conduction band of the TiO 2 to the<br />

carbon nanotubes (Garriga I Cabo 2007).<br />

TiO 2 @MWCNTs were characterized by FE-SEM (Figure 4.17). As can be seen<br />

there, there are aggregates and clusters of carbon nanotubes coated by TiO 2 .<br />

Additionally, the presence of small aggregates composed exclusively of TiO 2 , as<br />

determined by EDX analysis, can also be observed (arrows in Figure 4.17).<br />

Figure 4.17.<br />

FE-SEM image of the as-synthesized<br />

TiO 2 @MWCNTs at a magnification of 5000x. Arrows<br />

correspond to the presence of small clusters of TiO 2 , as<br />

determined by EDX analysis.<br />

The specific surface area of this hybrid material (S area ), determined by the BET<br />

method, was 620 m 2 g -1 . This high surface area implies that this material could have<br />

interesting applications in different catalytic processes. The TG analysis of this material<br />

shows that approximately the 25.27% of weight is lost during the thermal process<br />

(Figure 4.18). This high weight loss indicates that this is a porous material, as it was


Weight Loss (%)<br />

70<br />

stated from the BET analysis. The interior of the porous structure can contain adsorbed<br />

water or any other chemical substance as a residual of the synthesis.<br />

100<br />

TiO 2<br />

@MWCNT<br />

95<br />

90<br />

Weight Loss 25.27%<br />

85<br />

80<br />

75<br />

70<br />

65<br />

0 100 200 300 400 500 600<br />

Temperature ( o C)<br />

Figure 4.18.<br />

TGA analysis of the assynthesized<br />

TiO 2 @MWCNTs.<br />

Raman spectroscopy plays an important role in the research on carbon<br />

nanotubes because the signals observed in the spectra clearly depend on different<br />

structural parameters, including the diameter and the metallic or semiconductor<br />

character of the nanotubes (Anson Casaos 2005). Raman spectrum obtained from this<br />

hybrid material (TiO 2 @MWCNTs) (see Figure 4.19) confirmed the presence of TiO 2<br />

(rutile phase) on the surface of the multiwalled carbon nanotubes. The two peaks at<br />

444.32 and 603.69 cm -1 are characteristics of the rutile phase (Jackson 2004). The<br />

presence of several peaks ranging from ca. 100 to 300 cm -1 indicates the presence of<br />

carbon nanotubes. According to Anson Casaos (2005), Raman spectra of carbon<br />

nanotubes are characterized by having different peaks at very low Raman shifts (radial<br />

breathing mode or RBMs, at around 150 cm -1 ), and other modes including the tangential<br />

mode, TMs or G band (approx. 1600 cm -1 ), D band (approx. 1300cm -1 ) and the G band<br />

(around 2600cm -1 ), that are not shown in Figure 4.19.


Intensity (a.u.)<br />

71<br />

444.32 TiO 2<br />

@MWCNTs<br />

603.69<br />

200 400 600 800 1000 1200<br />

Raman Shift (cm -1 )<br />

Figure 4.19.<br />

Raman spectrum of the assynthesized<br />

TiO 2 @MWCNTs.<br />

TiO 2 @MWCNTs was also characterized by XPS. The Figure 4.20 shows the<br />

XPS spectrum corresponding to the C1s region. As can be seen there, C1s transition<br />

shows only a peak at 284.79 eV that has been ascribed to the C1s of the MWNTs (sp 2 -<br />

hybridized carbon).<br />

The presence of adsorbed carbon (CO 2 or hydrocarbons) was<br />

practically undetected. The Figure 4.21 shows the XPS spectrum corresponding to the<br />

Ti2p regions. The Ti2p transition is characterized by having the main peak at 457.43 eV<br />

and a secondary peak at ca. 463.04 eV that have been ascribed as Ti2p (3/2) and Ti2p (1/2) ,<br />

respectively, being in agreement with the expected peak positions for the rutile phase.<br />

The Figure 4.28 shows the XPS spectrum corresponding to the O1s region. As can be<br />

seen there, only a peak at ca. 532.39 eV is observed, being assigned to the oxygen at<br />

the TiO 2 lattice (Zhou et al. 2006). Figure 4.22 demostrated the presence of other O1s<br />

atoms, possible attached to the carbon portion of the MWCNTs.


CPS<br />

CPS<br />

72<br />

C1s TiO 2<br />

@MWCNTs<br />

284.79<br />

(C1s)<br />

294 292 290 288 286 284 282 280<br />

Binding Energy (eV)<br />

Figure 4.20. XPS spectrum corresponding to the<br />

C1s region of the as-synthesized TiO 2 @MWCNTs<br />

catalyst.<br />

457.43<br />

(Ti2p 1/2<br />

)<br />

TiO 2<br />

@MWCNTs<br />

463.04<br />

(Ti2p 3/2<br />

)<br />

470 465 460 455 450<br />

Binding Energy (eV)<br />

Figure 4.21. XPS spectrum corresponding to the<br />

Ti2p region of the as-synthesized<br />

TiO 2 @MWCNTs catalyst.


CPS<br />

73<br />

530.05<br />

(O1s)<br />

O1s TiO 2<br />

@MWCNTs<br />

536 534 532 530 528 526<br />

Binding Energy (eV)<br />

Figure 4.22. XPS spectrum corresponding to<br />

the O1s region of the as-synthesized<br />

TiO 2 @MWCNTs catalyst.<br />

This catalyst has been characterized by XRD (Figure 4.06). The presence of a<br />

broad peak ranging from 20° to 35° can make difficult the identification of the peaks<br />

(Figure 4.06). Some of the characteristics peaks that identify the rutile phase and their<br />

lattice planes were observed in this sample (Hernandez Enriquez et al. 2008; Cotto et al.<br />

2011).<br />

4.1.5. Zinc Oxide<br />

Zinc oxide was synthesized according with the experimental procedure described<br />

previously in the Chapter 3-Material Synthesis.<br />

ZnO nanoparticles have been<br />

extensively used as catalysts and in a wide range of applications including: gas sensors,<br />

cosmetics, as anti-virus agents, in the development of piezoelectric transducers, solar<br />

cells and transparent electrodes, etc. (Hong et al. 2009; Sridevi and Rajendra 2009).<br />

The ZnO nanoparticles used in cosmetics could be harmful to people because they can<br />

generate OH radicals, which can affect the cells (Hong et al. 2009). Different synthesis<br />

methods including sol-gel, hydrothermal, homogeneous precipitation, mechanical milling,


74<br />

organometallic synthesis, thermal evaporation, etc. have been used for the synthesis of<br />

this nanomaterial (Hong et al. 2009; Sridevi and Rajendra 2009).<br />

Other additional<br />

methods for the synthesis of ZnO are the precipitation and calcination of different<br />

precursors (Hong et al. 2009). ZnO is an important semiconductor, as the TiO 2 (Hong et<br />

al. 2009), having a wide band gap of 3.37 eV (Sridevi and Rajendra 2009).<br />

ZnO<br />

nanoparticles were characterized by FE-SEM (Figure 4.23). As can be seen there, ZnO<br />

nanoparticles are characterized by having irregular forms and dimensions ranging from<br />

several hundred nanometers to no more than one-micrometer length.<br />

Figure 4.23.<br />

FE-SEM images of the as-synthesized ZnO particles at<br />

different magnification: 25 000x (a), 50 000x (b).<br />

According with the TG curve (Figure 4.24), the weight-loss was approximately<br />

25.27% and possibly corresponds to the loss of water and the removal of surplus<br />

reagents. The specific surface area (S area ), as determined by the BET method, was 68<br />

m 2 g -1 .


Intensity (a.u.)<br />

Weight Loss (%)<br />

75<br />

ZnO<br />

100<br />

90<br />

Weight Loss 25.27%<br />

80<br />

70<br />

100 200 300<br />

Temperature ( o C)<br />

Figure 4.24. TG curve of the as-synthesized<br />

ZnO particles.<br />

Raman spectrum of the as-synthesized ZnO particles (Figure 4.25) shows two<br />

relevant peaks at 326.15 cm -1 and 436.72 cm -1 . Both peaks are characteristic of the<br />

ZnO catalyst (Jackson 2004).<br />

436.72<br />

ZnO<br />

326.15<br />

200 400 600 800 1000 1200<br />

Raman Shift (cm -1 )<br />

Figure 4.25.<br />

Raman spectrum of the assynthesized<br />

ZnO particles.


Intensity (cps)<br />

76<br />

Figure 4.26 shows the XRD diffraction pattern of the as synthesized ZnO<br />

catalyst.<br />

The most characteristic crystallographic lattice planes are present in the<br />

diffragtogram (Hong et al. 2009; Sridevi and Rajendran 2009) and correspond to<br />

reflections of the hexagonal phase (Sridevi and Rajendran 2009).<br />

3000<br />

(002)<br />

ZnO-as synthesized<br />

2500<br />

2000<br />

1500<br />

1000<br />

(100)<br />

(101)<br />

(103)<br />

(110)<br />

500<br />

0<br />

0 10 20 30 40 50 60 70 80 90<br />

2 Theta (Degree)<br />

Figure 4.26. XRD diffraction pattern of the assynthesized<br />

ZnO particles.<br />

4.2. Fenton Catalysts<br />

4.2.1. Iron Oxide Nanowires (Fe 2 O 3 NWs)<br />

Iron Oxide Nanowires (Fe 2 O 3 NWs) were synthesized according with the<br />

procedure previously described (Chapter 3-Material Synthesis). FE-SEM images of the<br />

as-prepared samples are shown in Figure 4.27. The iron oxide nanowires are<br />

characterized by being formed by filaments very long and extremely thin. Nevertheless,<br />

the Fe 2 O 3 NWs become coarser under increasing the temperature (above 600 ºC),<br />

indicating the temperature effect on the morphologies of these nanostructures. The BET<br />

method reveals an unexpectedly high specific surface area (S area ) of 180 m 2 g -1 .


Weight Loss (%)<br />

77<br />

Figure 4.27. FE-SEM images of the as-synthesized iron oxide nanowires<br />

(Fe 2 O 3 NWs) at different magnification: 1000x (a), 2000x (b) and 5000x (c).<br />

TG analysis revealed only a weight loss of 3.62% at 520 ºC (Figure 4.28),<br />

indicating that iron nanowires do not possess porosity able to absorb a measurable<br />

quantity of water and other solvents.<br />

100<br />

Fe 2<br />

O 3<br />

NWs<br />

98<br />

96<br />

94<br />

Weight Loss 3.62%<br />

92<br />

90<br />

0 100 200 300 400 500 600<br />

Temperature ( o C)<br />

Figure 4.28. TG curve of raw Fe 2 O 3 NWs.


CPS<br />

78<br />

Iron oxide nanowires were also characterized by XPS. Figure 4.29 shows the<br />

XPS spectrum of the Fe2p region. As can be seen there, two relevant peaks at 711.02<br />

eV and 724.74 eV have been assigned to the Fe2p (3/2) and Fe2p (1/2) transitions,<br />

respectively. Binding energies observed for the Fe2p region are in agreement with the<br />

presence of Fe 3+ atoms of the oxide (Garriga I Cabo 2007). Figure 4.30 corresponds to<br />

the O1s XPS region. As can be seen there, a main peak at 529.80 eV was measured,<br />

being assigned to the oxygen atoms of the iron oxide. On the other hand, an additional<br />

peak appears in this region. At ca. 532 eV a very low intense peak is observed and<br />

possibly could be due to the presence of structural defects on the nanowire surface<br />

(Fundamental XPS Data 1999).<br />

724.74<br />

(2p1)<br />

711.02<br />

(2p3)<br />

Fe2p Fe 2<br />

O 3<br />

NWs<br />

740 730 720 710 700<br />

Binding Energy (eV)<br />

Figure 4.29. XPS spectrum corresponding<br />

to the Fe2p region of the as-synthesized<br />

Fe 2 O 3 NWs.<br />

Raman spectrum of Fe 2 O 3 NWs was also analyzed during this research (Figure<br />

4.31). As can be seen there, two main peaks at low Raman shift (ca. 210 and 276 cm -1 )<br />

have been measured. Both peaks were ascribed to Fe-O vibrations (Chandra et al.<br />

2010).


Intensity (a.u.)<br />

CPS<br />

79<br />

529.80<br />

(O1s)<br />

O1s TiO 2<br />

NWs<br />

540 538 536 534 532 530 528 526 524<br />

Binding Energy (eV)<br />

Figure 4.30. XPS spectrum corresponding<br />

to the O1s region of the as-synthesized<br />

Fe 2 O 3 NWs.<br />

Fe 2<br />

O 3<br />

NWs<br />

200 400 600 800 1000 1200<br />

Raman Shift (cm -1 )<br />

Figure 4.31.<br />

Raman spectrum of assynthesized<br />

Fe 2 O 3 NWs.


Moment/Mass(emu/g)<br />

80<br />

Magnetic properties of as-synthesized Fe 2 O 3 NWs were also analyzed. Figure<br />

4.32 shows the hysteresis loop of the iron nanowires. The coercivity value obtained for<br />

this material was 20.485 G, with saturation magnetization (M s ) of 22.376 emu g -1 ,<br />

indicating a high ferromagnetic behavior (Wu et al. 2010).<br />

30<br />

nanowires Fe<br />

20<br />

Coercivity (Hci):20.485 G<br />

Magnetization (Ms): 22.376 emu/g<br />

10<br />

0<br />

-10<br />

-20<br />

-30<br />

-30000 -20000 -10000 0 10000 20000 30000<br />

Field(G)<br />

Figure 4.32. Magnetic susceptibility of assynthesized<br />

Fe 2 O 3 NWs, measured at room<br />

temperature.<br />

This material was also characterized by XRD. Two XRD diffraction patterns are<br />

observed in Figure 4.42, corresponding to the reflections measured when this material is<br />

synthesized in oxidative atmosphere (flowing oxygen) at 600 ºC (Figure 4.33a) and<br />

700 ºC (Figure 4.33b). As can be seen there, some small differences can be observed<br />

as a function of the synthesis temperature, demonstrating possible changes in the<br />

structure (phase), crystallinity and density of the materials (Bonilla et al. 2011). Iron<br />

oxide as hematite phase (α-Fe 2 O 3 ) was not observed, due to the lack of peaks<br />

corresponding to reflections (210) and (211), that are always present in this<br />

crystallographic phase (Daou et al. 2006).


81<br />

110<br />

422<br />

320 440 300<br />

221<br />

220<br />

400<br />

511<br />

b<br />

a<br />

40 60<br />

2Theta (Degree)<br />

Figure 4.33. XRD diffraction patterns of<br />

Fe 2 O 3 NWs synthesized at 600 °C (a) and<br />

700 °C (b) at atmospheric pressure and in<br />

flowing oxygen.<br />

4.2.2. Capped Magnetite Nanoparticles (Fe 3 O 4 )<br />

The hematite, magnetite and maghemite are different crystallographic phases of<br />

iron oxides with different types of magnetic transitions (Wu et al, 2010). The magnetite<br />

is an iron oxide in an invert spinel , composed by Fe 2+ and Fe 3+ and is relevant because<br />

it can be used in a vast range of different applications, including biomedical uses,<br />

catalysis, fine chemistry, development of batteries, magnetic recorders, etc. (Daou et al.<br />

2006). Several important questions in the magnetite synthesis are the cationic<br />

distribution and vacancies in the structure, the stoichiometry variation during the reaction<br />

and spin canting (Daou et al. 2006). During a typical synthesis procedure the Fe 2+ and<br />

Fe 3+ ions are present in the solution, reacting with the base and the final intermediates<br />

and producing the magnetite phase, according to the following reactions (Nyiro-Kosa et<br />

al. 2009):<br />

Fe 2+ + 2OH - → Fe(OH) 2


82<br />

Fe 3+ + 3OH - → FeO(OH) + H 2 O<br />

Fe(OH) 2 + 2FeO(OH) → Fe 3 O 4 + 2H 2 O<br />

During the last few years, the synthesis of different materials based in magnetites<br />

has experienced an important increase. One example is the use of magnetite over<br />

graphene oxide to bind heavy metal pollutants as arsenic, with relevant applications in<br />

water remediation processes (Chandra et al. 2010). Many synthesis procedures, with<br />

different reagents as ferric chloride, ferrous chloride and ferric sulfate have been<br />

established for the synthesis of magnetites with different sizes (from nanoparticles with<br />

very low diameters to micro and macroparticles) under different pH conditions,<br />

temperature and reaction times (Nyiro-Kosa et al. 2009). Some of the most common<br />

synthesis methods for obtaining magnetites include the synthesis by coprecipitation<br />

using ferric and ferrous compounds and different hydrothermal approaches (Daou et al.<br />

2006; Daou et al. 2007; Wu et al. 2010). The magnetic materials are synthesized in<br />

many structural forms, as nanoclusters, nanoparticles, hollow nanoparticles, nanorings,<br />

nanocapsules, and nanowires (Wu et al. 2010). The magnetite (Fe 3 O 4 ) and maghemite<br />

(γ-Fe 2 O 3 ) have many technological applications; the hematite (α-Fe 2 O 3 ) is used as<br />

catalyst, pigment and gas sensor (Wu et al. 2010) and emerges as a relevant material in<br />

the nanotechnology.<br />

The capped magnetite nanoparticles (Fe 3 O 4 ) were synthesized according with<br />

the experimental procedure described previously (Chapter 3 - Material Synthesis). The<br />

specific surface area (S area ) of this material, determined using the BET method, was 97<br />

m 2 g -1 . FE-SEM images of this material are shown in Figure 4.34. At very low<br />

magnification (30x) the material is characterized by forming large aggregates (Figure<br />

4.34a). At larger magnifications (2 000x), it is possible to distinguish very small particles<br />

forming these aggregates (Figure 4.34b). The Figure 4.34c shows a TEM image of the<br />

capped magnetites previously disaggregated in ethanol by using a soft ultrasound


83<br />

treatment for 15 min. As can be seen there, these particles have sizes of no more than<br />

5 nm-diameter.<br />

Figure 4.34. FE-SEM images of the as-synthesized capped<br />

magnetite nanoparticles (Fe 3 O 4 ) at different magnification 30x<br />

(a) and 2000x (b).<br />

TEM image of the capped magnetite<br />

nanoparticles (Fe 3 O 4 ) previously disaggregated in ethanol (c).<br />

Capped magnetite nanoparticles (Fe 3 O 4 ) were also characterized by XPS. The<br />

Figure 4.35 shows the XPS spectrum of the Fe2p region. As can be seen there, two<br />

peaks arising from the Fe2p splitting are observed. Both peaks have been<br />

unambiguously assigned to Fe2p (3/2) (724.1 eV) and Fe2p (1/2) (710.9 eV), and their<br />

binding energies correspond to iron in magnetite phase. Figure 4.36 shows the XPS<br />

spectrum corresponding to the O1s region. As can be seen there, only a peak at ca.<br />

531.1 eV has been measured, corresponding to oxygen in the lattice of the magnetite<br />

phase.<br />

Raman spectrum of the as-synthesized capped magnetite nanoparticles is shown<br />

in Figure 4.37. Several peaks ranging from ca. 200 to 400 cm -1 , and an intense peak at


CPS<br />

CPS<br />

84<br />

ca. 980 cm -1 were measured and assigned to typical Fe-O vibrations corresponding to<br />

the magnetite phase (Marquez et al. 2012).<br />

1.20E+014<br />

Fe2p<br />

(3/2)<br />

Fe2p<br />

(1/2)<br />

O1sFe 3 O 3 O 4 4<br />

1.00E+014<br />

8.00E+013<br />

6.00E+013<br />

4.00E+013<br />

2.00E+013<br />

740 730 720 710 700 690<br />

Binding Energy (eV)<br />

Figure 4.35. XPS spectrum corresponding to the<br />

Fe2p region, of the as-synthesized capped<br />

magnetite nanoparticles (Fe 3 O 4 ).<br />

O1s<br />

540 538 536 534 532 530 528 526 524<br />

Binding Energy (eV)<br />

Figure 4.36. XPS spectrum corresponding to<br />

the O1s region, of the as-synthesized capped<br />

magnetite nanoparticles (Fe 3 O 4 ).


Moment/Mass(emu/g)<br />

Intensity (a.u.)<br />

85<br />

Fe 3<br />

O 4<br />

200 400 600 800 1000 1200<br />

Raman Shift (cm -1 )<br />

Figure 4.37.<br />

Raman spectrum of assynthesized<br />

capped magnetite<br />

nanoparticles.<br />

Figure 4.38 shows the magnetometry study of the synthesized magnetites as a<br />

function of the reaction temperature.<br />

The saturation magnetization (M s ) values<br />

experienced important variations depending on the synthesis temperature.<br />

45<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

-5<br />

-10<br />

-15<br />

-20<br />

-25<br />

-30<br />

-35<br />

-40<br />

-45<br />

70C<br />

80C<br />

90C<br />

100C<br />

110C<br />

0 -20000 -10000 0 10000 20000 0<br />

Field(G)<br />

Figure 4.38.<br />

Temperature effect on the<br />

magnetic properties of the magnetites at<br />

different temperatures.


Weight Loss (%)<br />

86<br />

The lack of coercivity demonstrated the paramagnetic properties of the magnetite<br />

particles (Marquez et al. 2012). The sample with higher paramagnetic properties is that<br />

synthesized at 100 °C.<br />

4.2.3. Ferrous Chloride (FeCl 2 )<br />

Ferrous chloride (FeCl 2 , Fisher Scientific, 99%) was used during this<br />

investigation as catalyst. During the TG analysis, the FeCl 2 loss approximately the<br />

16.14% of its weight (Figure 4.39). Only one step was observed in the TG curve and<br />

this weight loss could be attributed to the removal of water molecules adsorbed on the<br />

surface of this reagent (Niederberger et al. 2002). The specific surface area (S area ), as<br />

measured by using the BET method, was 55 m 2 g -1 .<br />

100<br />

98<br />

96<br />

FeCl 2<br />

94<br />

92<br />

90<br />

Weight Loss 16.14%<br />

88<br />

86<br />

84<br />

82<br />

0 100 200 300 400 500<br />

Temperature ( o C)<br />

Figure 4.39.<br />

TG curve of the ferrous<br />

chloride.<br />

This reagent was also characterized by XPS. Figure 4.40 shows the XPS region<br />

corresponding to Cl2p. As can be seen there, a peak at 198.18 eV, assigned to Cl - , was<br />

observed (Handbook of the Elements 1999). The Fe2p XPS region was also analyzed.<br />

As can be seen in Figure 4.41, two peaks at ca. 710.9 eV and 724.8 eV were measured,<br />

being ascribed to Fe2p (3/2) and Fe2p (1/2) transitions, respectively (Fundamental XPS Data


CPS<br />

CPS<br />

87<br />

1999). Figure 4.42 shows the XRD diffraction pattern of the FeCl 2 catalyst. All measured<br />

reflections are in agreement with those expected for this compound.<br />

198.18<br />

(2p3)<br />

Cl<br />

FeCl 2<br />

204 202 200 198 196 194 192<br />

Binding Energy (eV)<br />

Figure 4.40. XPS spectrum corresponding to<br />

the Cl2p region of FeCl 2 catalyst.<br />

724.66<br />

(2p 3/2<br />

)<br />

710.93<br />

(2p 1/2<br />

)<br />

Fe FeCl 2<br />

730 725 720 715 710<br />

Binding Energy (eV)<br />

Figure 4.41. XPS spectrum corresponding to<br />

the Fe2p region of FeCl 2 catalyst.


Intensity (cps)<br />

88<br />

4000<br />

FeCl 2<br />

3500<br />

3000<br />

2500<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

-500<br />

0 10 20 30 40 50 60 70 80 90<br />

2 Theta (Degree)<br />

Figure 4.42. XRD diffraction pattern of the FeCl 2<br />

catalyst.<br />

4.2.4. Copper Oxide (CuO)<br />

Cupric oxide (CuO) (JT Baker, Baker Analyzed Reagent) was used during the<br />

investigation as one of the Fenton catalysts. The FE-SEM images of this compound<br />

(Figure 4.43) revealed the presence of clusters or aggregations, showing particles with<br />

irregular forms and particle sizes ranging from lesser than 1 micrometer to more than 3-4<br />

micrometers. A weight loss of ca. 25.27% (ranging from RT to 575 o C) was observed<br />

during the TG analysis of the sample (Figure 4.44), that could be attributed to the<br />

removal of water molecules adsorbed on the material (Niederberger et al. 2002). The<br />

specific surface area (S area ), as determined using the BET method, was 32 m 2 g -1<br />

indicating that this compound does not have relevant porous structure (as was also<br />

observed by FE-SEM).


Weight Loss (%)<br />

89<br />

Figure 4.43. FE-SEM images of CuO at different magnification: 5000x<br />

(a) and 25 000x (b).<br />

100<br />

CuO<br />

95<br />

90<br />

85<br />

Weight Loss 25.27%<br />

80<br />

75<br />

70<br />

65<br />

60<br />

0 100 200 300 400 500 600<br />

Temperature ( o C)<br />

Figure 4.44.<br />

TG curve of the cupric oxide<br />

catalyst.<br />

The XRD diffraction pattern of the CuO catalyst is shown in Figure 4.45. The<br />

most intense peaks, ascribed to the characteristic reflections of crystalline CuO, were<br />

observed at 35.52°, 38.55°, 48.18° and 61.57° (Yang et al. 2010).


Intensity (cps)<br />

90<br />

CuO<br />

35.52 o 61.57 o<br />

12000<br />

38.55 o<br />

10000<br />

8000<br />

6000<br />

4000<br />

48.18 o<br />

2000<br />

0<br />

0 10 20 30 40 50 60 70 80 90<br />

2 Theta (Degree)<br />

Figure 4.45. XRD diffraction pattern of the<br />

CuO catalyst.


Chapter Five<br />

Results and Discussion<br />

5.1. Defining the Experimental Parameters<br />

At the beginning of the investigation it was necessary to determine the optimal<br />

parameters for the degradation processes.<br />

The parameters studied were the<br />

concentration of the catalyst, the pH and the temperature of the solution. The selected<br />

dye and catalyst that were used to establish the optimal parameters were Rhodamine B<br />

and titanium oxide in anatase phase, respectively.<br />

5.1.1. Effects of the Concentration<br />

For the study of the effects of the concentration of catalyst in the photochemical<br />

degradation process, different concentrations of the catalyst (anatase) were added to a<br />

reaction mixture containing Rhodamine B (RhB, 10 -5 M). The photocatalytic process<br />

was carried out using the procedure established previously. The Figure 5.01 shows the<br />

effects of the catalyst concentration on the degradation process.<br />

As expected, a<br />

maximum photodegradation was observed in a low concentration range of the catalyst<br />

(0.6 – 0.9 gL -1 ). Under these reaction conditions near the 100% of degradation was<br />

produced. At higher concentration of catalyst, the photocatalytic process was inefficient<br />

due possibly to the poor dispersion of the catalyst in the solution that increased the<br />

turbidity, reducing the contact between the catalyst and the reaction mixture. Previous<br />

studies (Velegraki and Mantvinos 2008) revealed that no significant changes in the<br />

photodegradation process were observed when the catalyst was increased from 0.6 gL -1<br />

to 0.8 gL -1 . Nevertheless, when the concentration of the catalyst increases to 1.0 gL -1 or<br />

more, a slight decrease in the photocatalytic conversion is observed.<br />

91


% Photo-degradation (at T= 60 min)<br />

92<br />

The results could be explained according the results of Lodha et al. (2008). The<br />

authors (Lodha et al. 2008) concluded that when the concentration of the dye<br />

increases, more molecules of the dye are present in the reaction system and this high<br />

concentration increases the opacity of the solution avoiding the pass of light through the<br />

water column and decreasing the degradation rate.<br />

100<br />

Rh-B + Anatase<br />

80<br />

60<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0<br />

Concentration of catalyst (gL -1 )<br />

Figure 5.01.<br />

Effects of the concentration of<br />

anatase on the photodegradation process of<br />

RhB.<br />

Another relevant test was carried out to determine the relevance of the catalyst<br />

and the hydrogen peroxide in the degradation process, because it was necessary to<br />

know if the degradation process can proceed without the presence of catalyst or the<br />

hydrogen peroxide.<br />

The Figure 5.02 shows the results of this test; both reagents<br />

(catalyst and hydrogen peroxide) were necessary for the process. Previous studies<br />

corroborated these results (Huang et al. 2010a). In the absence of the catalyst, no<br />

degradation process (photocatalysis, sonocatalysis and sonophotocatalysis) was<br />

observed (Minero et al. 2005; Vinu and Madras 2009).<br />

In the Fenton reactions,<br />

Massomboon et al. (2009) demonstrated that if an excess of the iron catalyst is added


C/Co<br />

93<br />

to the reaction mixture, a decrease in the degradation process was observed because<br />

the iron can react with the hydroxyl radicals formed during the process. A synergistic<br />

effect was also observed between the photocatalysis and sonocatalysis increasing the<br />

degradation effect because the sonocatalysis avoids any possible aggregation of the<br />

catalyst, increasing the surface area and the efficiency of the irradiation on the sample<br />

degradation (Vinu and Madras 2009). Su et al. (2010) studied the degradation of CV<br />

using Mn 2 O 3 as catalyst. The authors demonstrated the importance of the hydrogen<br />

peroxide in the degradation; a specific concentration is required for the reaction and<br />

higher concentrations do not seem to increase the dye degradation. The hydrogen<br />

peroxide is responsible for the generation of hydroxyl radicals in presence of the<br />

catalyst, being these radicals the species that initiate the degradation process<br />

(Masomboon et al. 2009).<br />

Without Catalyst<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

Without H 2<br />

O 2<br />

0.2<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure 5.02.<br />

Effects of the catalyst and<br />

hydrogen peroxide on the<br />

photodegradation process of RhB.<br />

An optimum concentration of the dye is also necessary for an effective process<br />

because this parameter can be related with the dispersion of the catalyst particles<br />

(Wang et al. 2010) (Figure 5.02). A small decrease in the concentration of the dye was


% Photo-degradation (at T= 60 min)<br />

94<br />

also observed when the reaction is carried out without the catalyst in presence of light<br />

(Asiri et al. 2011).<br />

5.1.2. Effects of the pH<br />

The study of the effects of the pH of the solution on the photochemical<br />

degradation process was also analyzed. The pH of the solution was changed from acid<br />

to basic without changing the concentration of the catalyst (anatase) and the dye (RhB,<br />

10 -5 M). The Figure 5.03 shows the effects of the pH on the degradation process. The<br />

optimal pH ranged from 7 – 8, obtaining approximately the 100% of degradation of the<br />

dye. At acid and basic pH ranges, the reaction is clearly less efficient, decreasing the<br />

percent of photodegradation. Some investigators (Devipriya and Yesodharan 2010),<br />

suggest that in acidic solutions the low reactivity observed in specific catalysts (i.e.<br />

ZnO) should be due to photocorrosion of the catalyst induced by the pH.<br />

100<br />

Rh-B + Anatase<br />

80<br />

60<br />

4 5 6 7 8 9 10<br />

pH<br />

Figure 5.03. The effects of the pH of the<br />

reaction mixture on the photodegradation<br />

process of RhB.<br />

The pH significantly affect the reaction because the concentration of the·OH<br />

groups changes (Lodha et al. 2008) the activity of the catalyst (Huang et al. 2010).<br />

According to Hong et al. (2009) the photodegradation process follows a kinetic of first


95<br />

order. The optimal pH is pH=7.0 and the photodegradation is favored at temperatures<br />

higher than room temperature, but if the temperature increases above 40-45 °C the<br />

degradation is partially stopped. Hong et al. (2009) observed that the concentration of<br />

the catalyst has a maximum value to be efficient during the degradation process. The<br />

optimal reaction conditions require a concentration of MO dye of 20ppm, a reaction time<br />

of 20 hours, 1.5g L -1 of catalyst, a reaction mixture at pH=7.0 and a temperature of 30<br />

°C.<br />

Hong et al. (2009) studied the photodegradation process of MO using ZnO<br />

nanoparticles and polystyrene-capped ZnO nanoparticles. The authors (Hong et al.<br />

2009) concluded that the ZnO was more efficient than the capped catalyst because<br />

their hydrophilic behavior permits the adsorption of more MO molecules from the<br />

solution and these molecules can easily be in contact with the air close to the surface of<br />

the catalyst and the flow of electron-holes from the catalyst to the surface.<br />

In the case of the Fenton reactions, the pH is also important for the degradation<br />

reaction rate. At lower pH the possible formation of (Fe(II)(H 2 O)) 2+ species and the low<br />

production of OH·are responsible for the decrease of the efficiency of this reaction<br />

(Massomboon et al. 2009).<br />

5.1.3. Effects of Temperature<br />

Effect of the temperature of the solution on the photochemical degradation was<br />

also analyzed.<br />

Concentrations of the catalyst and the RhB dye were previously<br />

determined.<br />

Figure 5.04 shows the effects of the temperature on the degradation<br />

process. Optimal degradation process was observed in a small range of temperatures<br />

(25-30 °C).


% Photo-degradation (at T= 60 min)<br />

96<br />

100<br />

Rh-B + Anatase<br />

80<br />

60<br />

40<br />

5 10 15 20 25 30 35 40 45 50<br />

Temperature (ºC)<br />

Figure 5.04. The effects of the temperature<br />

of the solution on the photodegradation<br />

process of RhB.<br />

5.2. Photochemical degradation<br />

Many substances commonly used are ecotoxic.<br />

The aromatic compounds<br />

constitute an important source of environmental pollution reaching the atmosphere and<br />

groundwaters because there are wi<strong>del</strong>y used as intermediates in the production of<br />

pesticides, synthetic polymers, dyes, etc. (Huang et al. 2010a). These substances in<br />

the environment are a concern because they possess carcinogenic, teratogenic and<br />

toxic characteristics (specially the azo dyes), decrease the light penetration through the<br />

water column, affect aesthetically and appreciably alter the gas solubility (Karadag et al.<br />

2006; Vanhulle et al. 2008; Huang et al. 2010) damaging the environment<br />

(Dafnopatidou et al. 2007).<br />

In the last decades, new applications for the use of nanoparticles in<br />

homogeneous and heterogeneous catalytic reactions were developed because these<br />

materials show a high efficiency and a high surface-to-volume ratio along with high<br />

surface energy (Pattapu et al. 2008) and will be part of the new green chemistry<br />

technologies (Cao et al. 2010). Similarly to the biological process of photosynthesis in<br />

which the chlorophyll (photosystem II) acts as a photoabsorbent substance, the


97<br />

photocatalyst is responsible for the generation of electron-hole pairs when the light has<br />

higher energy than the band gap of the photocatalyst, being a part of the chemical<br />

reaction (Hong et al. 2009).<br />

The photocatalyst (Hong et al. 2009) is one of the components of the<br />

photodegradation process, and the different reactions involved using these components<br />

are grouped in a variety of processes named Advanced Oxidation Processes or AOPs.<br />

As mentioned previously, the AOPs use different chemical methods to generate<br />

intermediate species, as the hydroxyl radicals for the oxidation of substances. The<br />

most common method for the generation of the OH radicals includes the use of<br />

hydrogen peroxide, ozone and UV irradiation (Hadj Salah et al. 2004).<br />

The oxidation processes are non-selective. In the case of the photochemical<br />

degradation, the energy source to drive the reaction is the UV or Visible light. During<br />

the photodegradation processes, a catalyst should be used to absorb the photons of the<br />

light. These catalysts are normally semiconductors, having a band gap lower or equal to<br />

the energy of the photons used during the reaction.<br />

The photochemical process<br />

generated by using these photocatalysts transforms the pollutants in CO 2 , H 2 O and<br />

inorganic acids without generation of secondary compounds that could be toxic (Asiri et<br />

al. 2011).<br />

The interest in these processes is increasing, because different studies have<br />

demonstrated that these processes are efficient in the degradation of organic<br />

compounds and generate very low concentration of by-products during the degradation<br />

reaction (Hernandez Enriquez et al. 2008).<br />

The heterogeneous photocatalysis is<br />

described by Hernandez Enrique et al. (2008) as the degradation of a contaminant<br />

using catalysts which normally are oxides of semiconductors, ultraviolet or solar<br />

irradiation to generate radicals as O 2·-,<br />

HO 2· and OH· that finally are the responsible for<br />

the oxidation of the pollutants. The possible functional groups on the surface of the


98<br />

titanium oxide in aqueous solution may be TiOH 2 + , TiOH and TiO -<br />

(Devipriya and<br />

Yesodharan 2010).<br />

equation:<br />

The band gap energy of a photocatalyst could be estimated by the following<br />

E <br />

g<br />

Where E g is the band gap energy, h is the Planck’s constant, c is the light velocity and<br />

λ<br />

0<br />

is the absorption wavelength (Lo et al. 2004; Yu et al. 2009). Yu et al. (2009)<br />

mentioned;<br />

hc<br />

λ<br />

0<br />

“Generally, the rate of the photocatalytic reaction is proportional<br />

to<br />

(I where Iα<br />

is the photon number absorbed by the<br />

n<br />

α Φ)<br />

photocatalyst per second and Φ is the efficiency of the band gap<br />

transition.”<br />

When the photocatalyst is exposed to low light intensity during the reaction, the<br />

exponential value of n=1 and if the catalyst is exposed to high light intensity n=1/2 (Yu<br />

et al. 2009). According with Ruan and Zhang (2009):<br />

“the UV –driven photocatalytic activity of the sample is much<br />

higher than the visible light –driven photocatalytic activity”<br />

because the shorter wavelength produces a higher increase of the quantum yield.<br />

In the photocatalytic process the generation of superoxide radicals and other<br />

oxygen radical species is caused by the transfer of an electron to an oxygen molecule<br />

when the dye is in the excited state (Yu et al. 2009). Other studies (Ma et al. 2007)<br />

demonstrated that the degradation reaction is mediated by a radical mechanism<br />

because during a comparative analysis between a control group and a radical<br />

scavenger-containing group a difference with statistical significance was observed. An


99<br />

example for the degradation of an organic compound is the photodegradation of 2-<br />

Mercaptobenzothiazole (Li et al. 2006).<br />

Another technique used for the degradation of organic pollutants is the<br />

photoelectrocatalytic process (Xu et al. 2009) in which capped electrodes are<br />

necessary to avoid the reduction of the cathode by the hydrogen peroxide formed<br />

during the reaction.<br />

O 2 + 2H + + 2e - → H 2 O 2<br />

The biological processes have some disadvantages. Biological decolorization process<br />

by K rosea is effective only under anaerobic conditions because the oxygen competes<br />

with the dye during the reaction and inhibits the process (Parshetti et al. 2006).<br />

According to Liu et al. (2008) the surface area is not the only factor that controls<br />

the process; the crystal structure is relevant for the catalytic process. Hadj Salah<br />

(2004) determined that the structure, diameter of the particle, size of the crystallite and<br />

the electronic properties are relevant to determine the catalytic activity. Another author<br />

(Liu et al. 2008) mentioned as important the size of the particle and the capability to<br />

remove the catalyst after the catalytic degradation process.<br />

As mentioned previously the relationship between the dye and the catalyst is<br />

relevant. Taking into account the adsorption equilibrium between the dye and the<br />

catalyst, the equilibrium is given by (Karadag et al. 2006; Mahanta et al. 2008):<br />

(C<br />

q e <br />

0 <br />

Where q e is the amount of dye adsorbed at the equilibrium; C 0 and C e are the initial<br />

concentration and concentration at equilibrium, respectively; V is the volume of the<br />

solution and W is the mass of the catalyst used during the reaction (Mahanta et al.<br />

2008).<br />

C<br />

W<br />

e<br />

)V


100<br />

Another important issue is the characteristic of the catalyst. According to Ma et<br />

al. (2007) the microstructure and morphology has a great influence in the selectivity of<br />

the catalyst to degrade a dye. Also, no synergistic or inhibited effect is observed when<br />

a mixture of TiO 2 and ZnO is used for photodegradation of phenols (Devipriya and<br />

Yesodharan 2010).<br />

5.2.1. Description of the Photocatalytic System<br />

The experimental setup used for the photocatalytic reaction during this<br />

investigation was adapted from a similar method described by Hernandez Enriquez et<br />

al. (2008). A cylindrical reactor (semi-batch type) with continuous stirring was located in<br />

the center of two double tubular lamps, which were the irradiation source.<br />

The<br />

experimental setup (Figure 5.05) was composed by two annular white bulb lights with a<br />

total irradiation power of 60 watts. A vessel of 1 L was used during the irradiation of the<br />

sample.<br />

The sample was mechanically stirred with a paddler to maintain a<br />

homogeneous mixture during the irradiation of the sample. Before the irradiation, the<br />

catalyst was suspended in the solution and kept in dark with stirring for at less 30 min<br />

(Hong et al. 2009) to reach the adsorption-desorption equilibrium (Zhou et al. 2010).<br />

All the system was covered to avoid any other irradiation on the sample; only<br />

the light of the bulbs could reach the sample. Every 10 minutes a sample of 10 mL was<br />

taken to obtain the UV and fluorescence spectra and to determine the TOC<br />

concentration. The concentration of the dye and the catalyst were 10 -5 M and 0.6 gL -1<br />

respectively (Velegraki and Mantvinos 2008; Asiri et al. 2011).


101<br />

Figure 5.05. Experimental setup used during<br />

this research, without irradiation (a) and<br />

during the irradiation (b)<br />

Different organic pollutants (dyes and organic compounds) with different<br />

structures were used during the investigation.<br />

The organic pollutants used were<br />

Methylene Blue (MB), Rhodamine B (RhB), Methyl Orange (MO), Crystal Violet (CV),<br />

Methyl Violet (MV) and p-aminobenzoic acid (pABA) (Figure 5.06).<br />

Some basic<br />

information is available in Table 5.1.<br />

Figure 5.06. Dye solutions used during the investigation.<br />

From left to right; Methylene Blue, Methyl Orange,<br />

Crystal Violet, Rhodamine B and Methyl Violet.


Absorbance<br />

102<br />

The sample solutions with the dyes and the organic compound (pABA) had a<br />

concentration of 10 -5 M (Velegraki and Mantvinos 2008; Asiri et al. 2011). The<br />

concentration of the catalysts was 0.6 gL -1 in 300 mL of the solution. At the beginning,<br />

the spectrum of the organic contaminants were obtained. An example is shown in<br />

Figure 5.07, corresponding to MB. After the filtration of the catalyst, it was necessary to<br />

determine the maximum wavelength (λ max ) of the contaminant.<br />

500 600 700<br />

Wavelenght (nm)<br />

Figure 5.07. Methylene blue Visible<br />

absorption spectrum.<br />

Fluorescence, UV-visible absorption and TOC were determined for each<br />

sample. A decrease in the intensity of the absorption and fluorescence spectra was<br />

observed for all compounds along the degradation process. Figure 5.08 shows the<br />

fluorescence and the absorption spectra of MB in presence of rutile at different reaction<br />

times, showing the degradation process. Figure 5.08c and Figure 5.08d clearly show<br />

how the area of the curves decrease during the reaction time. Additionally, a smooth<br />

displacement of the maximum absorption peak could be observed.


103<br />

Table 5.1. Basic information of the studied organic compounds (adapted from Ma et al.<br />

2007).<br />

Dye Chemical Structure Molecular Weight<br />

(g mol -1 )<br />

A max (nm)<br />

(Observed)<br />

Methylene<br />

Blue<br />

C<br />

H 3 C<br />

N<br />

S +<br />

N<br />

CH 3<br />

H 3<br />

Cl - 373.88 g mole-1 658<br />

CH3<br />

N<br />

Rhodamine<br />

H 3 C<br />

C<br />

N<br />

O N + CH 3<br />

479.02 g mole -1 553<br />

B<br />

H 3<br />

CH 3<br />

COOH<br />

Methyl<br />

Orange<br />

Na 3<br />

327.34 g mole -1 465 Ma et<br />

O S<br />

N N N CH 3<br />

CH 3<br />

al. (2007)<br />

C<br />

Crystal Violet N + CH 3<br />

H 3 C<br />

CH 3<br />

H 3<br />

408.00 g mole -1 583<br />

CH 3<br />

CH 3<br />

Methyl Violet<br />

N + NH<br />

CH 3<br />

Cl 393.96 g mole -1 580<br />

-<br />

CH 3<br />

H 3 C N CH 3<br />

CH 3<br />

p-ABA<br />

O OH<br />

137.14 g mole -1 280 nm<br />

(Schmidt<br />

et al. 1997<br />

NH 2


Absorbance<br />

Absorbance<br />

Intensity (a.u.)<br />

Fluorescence<br />

104<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

Absorbance<br />

1: t=0<br />

2: t=10m<br />

3: t=20m<br />

4: t=30m<br />

5: t=45m<br />

6: t=60m<br />

1<br />

a) 1: t=0 b)<br />

2<br />

3<br />

4<br />

5<br />

6<br />

500<br />

450<br />

400<br />

350<br />

300<br />

250<br />

200<br />

150<br />

100<br />

Fluorescence<br />

1<br />

2<br />

3<br />

2: t=10m<br />

3: t=20m<br />

4: t=30m<br />

5: t=45m<br />

6: t=60m<br />

4<br />

5<br />

6<br />

0.0<br />

50<br />

500 600 700<br />

Wavelenght (nm)<br />

0<br />

700<br />

Wavelenght (nm)<br />

0.6<br />

c) d)<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure 5.08. Visible absorbance (a) and fluorescence (b) spectra<br />

of MB in presence of rutile, irradiated with white light (60W) at<br />

different reaction times and the corresponding representation of the<br />

areas (c and d). The inset of c corresponds to the original (left) and<br />

degraded (right) solutions.<br />

These behaviors could be observed due the formation of intermediates during<br />

the degradation process (Sun et al. 2009). The possible intermediates have absorption<br />

peaks at different wavelength than the original organic compounds; for this reason<br />

different absorption peaks were observed during the degradation process.<br />

Some<br />

possible intermediates were detected by Sun et al. (2009) during the photodegradation<br />

process of RhB with the CaSb 2 O 5 (OH) 2 catalyst (Figure 5.09).


105<br />

CH 3<br />

H 3 C N<br />

O N + CH 3<br />

COOH<br />

Deethylation<br />

CH 3<br />

Carboxylation<br />

COOH<br />

HOOC<br />

C H 3<br />

H<br />

N<br />

O N + CH 3<br />

CH 3<br />

N<br />

CH 3<br />

O N + CH 3<br />

CH 3<br />

Carboxylation<br />

Deethylation<br />

Carboxylation<br />

Deethylation<br />

COOH<br />

.<br />

HOOC<br />

H<br />

N<br />

O N + COOH<br />

Hydroxylation<br />

H<br />

HOOC<br />

H<br />

N<br />

H<br />

O N + CH 3<br />

COOH<br />

HOOC<br />

H<br />

N<br />

.OH<br />

O N + COOH<br />

H<br />

.OH<br />

CO 2 , H 2 O, Low Molecular Weight Byproducts<br />

Figure 5.09. Possible degradation intermediates of RhB during the photocatalytic<br />

processs (Adapted from Sun et al. 2009).


106<br />

<strong>To</strong> determine the photocatalytic degradation percent, the following equation was<br />

used (Parshetti et al. 2006; Dafnopatidou et al. 2007; Ma et al. 2007; Shimizu et al.<br />

2007; Mahanta et al. 2008; Hong et al. 2009):<br />

A0<br />

A<br />

PDP% 100%<br />

A<br />

Where A 0 is the absorbance at t=0 min and A is the absorbance at t=60 min. Table 5.2<br />

shows the percent of degradation obtained after the photocatalytic reaction between<br />

each mo<strong>del</strong> organic compound with the different catalysts used in this research.<br />

According with the data presented in Table 5.2 the titanium oxide nanowires has the<br />

highest degradation rate, reaching values between 93.13% and 98.51%. Wang et al.<br />

(2010) observed that pure TiO 2 NWs reflect near the 95% of the visible light irradiated to<br />

the catalyst; most of the light absorbed is UV-light (Wang et al. 2010). Using the<br />

nanowires to degrade the pABA compound, the 90.20% of degradation was reached.<br />

For the degradation of pABA the most efficient catalyst was the rutile catalyst (94.40%<br />

of degradation).<br />

The low efficiency of the ZnO catalyst could be due to the possible<br />

0<br />

photodecomposition of the catalyst in the solution during the photoreaction.<br />

The<br />

photocatalytic activity has an inverse correlation between the photolysis of the catalyst<br />

and the photodegradation of the dye (Kislov et al. 2009). The lowest degradation rate<br />

was obtained using the TiO 2 @MWNT (from 72.88% to 84.87%) (Figure 5.10). The<br />

degradation of aromatic pollutants by ·OH species is accomplished by an electrophilic<br />

mechanism (Huang et al. 2010).


Degradation (%)<br />

107<br />

Table 5.2. Degradation percent of dye solutions during the Photocatalytic Process.<br />

Organic<br />

Contaminant<br />

Catalyst<br />

Anatase Rutile TiO 2 @MWNT TiO 2 NWs ZnO<br />

MB 90.24% 88.17% 72.88% 93.13% 79.49%<br />

RhB 92.72% 94.37% 74.54% 96.44% 84.45%<br />

MO 93.55% 89.83% 74.12% 93.55% 88.59%<br />

CV 88.17% 92.72% 84.45% 96.85% 79.49%<br />

MV 91.07% 94.37% 84.87% 98.51% 90.24%<br />

p-ABA 94.00% 94.40% 77.40% 90.20% 82.40%<br />

100<br />

95<br />

90<br />

Anatase<br />

Rutile<br />

TiO2MWCNTs<br />

TiO2NWs<br />

ZnO<br />

85<br />

80<br />

75<br />

70<br />

MB RhB MO CV MV p-ABA<br />

Pollutant<br />

Figure 5.10. Graphic of the percent of degradation<br />

of the different organic compounds by the<br />

photocatalytic process.


Ln (C/C 0<br />

)<br />

108<br />

The synthesis of different reaction intermediates during the degradation process<br />

occurs. It is possible to think that many degradation reactions occur simultaneously in<br />

the same reaction mixture, and for this reason to define a reaction rate for all the<br />

different processes is extremely difficult. Therefore, the degradation process is defined<br />

as a pseudo-kinetic reaction (Pey 2008). For the photocatalytic process carried out<br />

during this research, the best type of kinetic reaction adapted is the pseudo-first order<br />

reaction (Figure 5.11). The equation used to determine the reaction rate is based on<br />

the definition of the mo<strong>del</strong>. The first kinetic mo<strong>del</strong> is defined as (Wang et al. 2006a;<br />

Asiri et al. 2011):<br />

ln( C0 C) kt<br />

where C 0 and C are the initial concentration and the concentration at any time,<br />

respectively. The semilogarithmic plots of the concentrations vs time give straight lines<br />

in which the slope represent the value of k (rate reaction) (Figure 5.11).<br />

0.0<br />

MB/Rutile/Photocatalysis<br />

-0.5<br />

-1.0<br />

y = -0.0365x<br />

R² = 0.997<br />

-1.5<br />

-2.0<br />

-2.5<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure 5.11.<br />

Regression curve of the<br />

Methylene Blue (MB) with rutile under<br />

photochemical process.


109<br />

The Table 5.3 shows the data of the kinetic reaction rate for the mo<strong>del</strong> organic<br />

compounds and the catalyst used during the degradation processes.<br />

The mean<br />

velocity for the reaction is approximately 10 -2 min. The values of R 2 were between 0.99<br />

and 0.71.<br />

The difference between these values can be justified as due to the<br />

adsorption-desorption process of the dye by the catalyst during the degradation<br />

process.<br />

Table 5.3. Kinetic reaction rates and R 2 values for the degradation reaction of the<br />

organic compounds during the photocatalytic process.<br />

Organic<br />

Contaminant<br />

Catalyst<br />

Anatase Rutile TiO 2 @MWNT TiO 2 NWs ZnO<br />

MB 4.24 x 10 -2<br />

3.65 x 10 -2<br />

2.75 x 10 -2<br />

4.95 x 10 -2<br />

3.19 x 10 -2<br />

R 2 = 0.9724<br />

R 2 =0.9970<br />

R 2 = 0.8004<br />

R 2 = 0.9625<br />

R 2 = 0.8872<br />

RhB 4.88 x 10 -2<br />

5.47 x 10 -2<br />

2.76 x 10 -2<br />

5.78 x 10 -2<br />

3.33 x 10 -2<br />

R 2 = 0.9589<br />

R 2 = 0.9429<br />

R 2 = 0.7102<br />

R 2 = 0.9939<br />

R 2 = 0.9633<br />

MO 4.53 x 10 -2<br />

3.87 x 10 -2<br />

2.75 x 10 -2<br />

5.23 x 10 -2<br />

3.45 x 10 -2<br />

R 2 = 0.9984<br />

R 2 = 0.9981<br />

R 2 = 0.8472<br />

R 2 = 0.9446<br />

R 2 = 0.9825<br />

CV 3.55 x 10 -2<br />

4.52 x 10 -2<br />

3.95 x 10 -2<br />

6.19 x 10 -2<br />

3.19 x 10 -2<br />

R 2 = 0.9926<br />

R 2 = 0.9797<br />

R 2 = 0.8313<br />

R 2 = 0.9824<br />

R 2 = 0.8728<br />

MV 4.70 x 10 -2<br />

5.16 x 10 -2<br />

3.53 x 10 -2<br />

7.31 x 10 -2<br />

4.17 x 10 -2<br />

R 2 = 0.9246<br />

R 2 = 0.9806<br />

R 2 = 0.9362<br />

R 2 = 0.9807<br />

R 2 = 0.9674<br />

p-ABA 4.26 x 10 -2<br />

4.52 x 10 -2<br />

2.90 x 10 -2<br />

4.37 x 10 -2<br />

2.95 x 10 -2<br />

R 2 = 0.9725<br />

R 2 = 0.9842<br />

R 2 = 0.8802<br />

R 2 = 0.9424<br />

R 2 = 0.9812


110<br />

Liu et al. (2008) suggest that the degradation rate of the titanium oxide is related<br />

to the band gap energy: “at higher band gap energy, the higher ultraviolet energy that<br />

can be absorbed to active the photocatalyst” enhancing the oxidation process. The<br />

difference in reaction time between the commercial and the synthesized material could<br />

be explained by the difference in absorption capacity, wavelength and the energy of the<br />

prohibited bands determined by both materials (Hernandez Enriquez et al. 2008).<br />

The ZnO has a lower degradation rate when this catalyst is compared with the<br />

TiO 2 catalysts (rutile, anatase and nanowires).<br />

Common semiconductors for<br />

degradation of contaminants are TiO 2 , ZnO, CdS, etc. (Lo et al. 2004), but some of<br />

them (as ZnO and CdS) has poor stability (Hernandez-Alonso et al. 2009).<br />

The titanium oxide has an amphoteric property, and positive or negative charges<br />

can be generated on the surface (Velegraki and Mantzavinos 2008). Changes in the<br />

pH during the degradation process could be observed and this phenomenon could be<br />

caused by a change in the charge on the surface of the catalyst. The charge on the<br />

TiO 2 surface is positive when the pH is 1 and the charge is negative at pH>9 (Asiri et al.<br />

2011). Also it is important to know that some by-products formed during the reaction<br />

have an acid pH that can alter the surface charge (Velegraki and Mantzavinos 2008).<br />

The specific surface area affects the reaction activity (Lo et al. 2004). Velegraki<br />

and Mantzavinos (2008) suggest that the reduction in the reaction rates could be due to<br />

the decrease of the active sites on the surface of the catalyst (titanium oxide) and the<br />

possible development of multilayers formed by the organic compound on the surface of<br />

the catalyst, avoiding the direct contact between the molecules of the compound and<br />

the catalyst.<br />

On the surface of the catalyst the semiconductor is excited by a photon of light<br />

and an electron-hole pair is generated.


111<br />

TiO<br />

<br />

<br />

<br />

<br />

<br />

<br />

2 hν TiO2<br />

eCB<br />

hVB<br />

<br />

The valence band hole has a high oxidative potential producing the oxidation of<br />

the dye and hydroxyl radical from the water molecule. Consecutive reactions allow the<br />

oxidation of the dye and the complete photodegradation.<br />

<br />

<br />

h VB dye dye Oxidation of the dye<br />

<br />

<br />

hVB<br />

2<br />

h<br />

H O H<br />

<br />

VB<br />

<br />

<br />

<br />

OH<br />

OH<br />

OH<br />

OH dye photodegradation of the dye<br />

The conduction band electron liberated from the surface produces radicals of the<br />

.<br />

oxygen molecule in the solution<br />

e<br />

<br />

<br />

CB O2<br />

O2<br />

. The oxygen radicals react with the<br />

hydrogen peroxide producing hydroxyl radical and ions. At the same time regenerate<br />

the O 2 to continue with the reaction O2 H2O2<br />

OH<br />

OH<br />

O2<br />

<br />

<br />

<br />

(Velegraki and<br />

Mantzavinos 2008; Asiri et al. 2011). Other authors mention the presence of four<br />

processes during the heterogeneous photocatalysis using TiO 2 (Figure 5.12) (Wang et<br />

al. 2006a).<br />

According to Asiri et al. (2011), the presence of “anchor” groups on the surface<br />

of the catalyst facilitates the anchorage of groups available in the dye, increasing the<br />

degradation processes.<br />

In a typical heterogeneous catalytic reaction, the decay<br />

observed in the RhB dye concentration was part of the adsorption-desorption process<br />

prior to reach the equilibrium (Figure 5.13) (Yu et al. 2009).


112<br />

Charge-carrier generation<br />

TiO 2<br />

hv <br />

<br />

h VB<br />

eCB<br />

Charge-carrier traping<br />

<br />

Ti( VI ) OH ( Ti(<br />

VI ) OH)<br />

<br />

h VB<br />

<br />

Ti( VI )) OH ( Ti(<br />

III)<br />

OH)<br />

e CB<br />

<br />

e CB<br />

Ti( VI ) Ti(<br />

III)<br />

Charge-carrier recombination<br />

h<br />

<br />

VB<br />

e<br />

<br />

CB<br />

heat<br />

<br />

<br />

e CB<br />

( Ti(<br />

VI ) OH)<br />

Ti(<br />

VI ) OH<br />

<br />

h VB<br />

( Ti(<br />

III)<br />

OH)<br />

Ti(<br />

VI ) OH<br />

Interfacial charge transfer<br />

( d<br />

<br />

<br />

Ti(<br />

VI ) OH)<br />

Red<br />

Ti(<br />

VI ) OH (Re )<br />

( Ti(<br />

III)<br />

OH)<br />

Ox Ti(<br />

VI ) OH ( Ox)<br />

Figure 5.12. Possible processes involved in the degradation reaction<br />

using TiO 2 as catalyst.


Absorbance<br />

113<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0 10 20 30 40 50 60<br />

Wavelength (nm)<br />

Figure 5.13. Spectrum corresponding to<br />

the degradation of Rhodamine B by<br />

TiO 2 @MWCNTs under photochemical<br />

proces demonstrating the adsorptiondesorption<br />

equilibrium.<br />

According to Wang et al. (2010a) the initial high concentration of the MB dye is<br />

caused by the adsorption of the dye by the nanowires. In the Pt@TiO 2 NWs a decrease<br />

in the adsorption is observed when it is compared with the pure TiO 2 NWs and the<br />

authors suggest that the Pt particles occupy the adsorption sites of the MB dye on the<br />

surface of the catalyst (Wang et al. 2010a).<br />

During the photocatalytic degradation (under visible light) of these types of<br />

compounds, two photooxidation mechanisms are common: the N-deethylation and the<br />

cleavage of the chromophore structure.<br />

The cleavage of the chromophore<br />

predominates over the other mechanisms (Yu et al. 2009). According to Yu et al.<br />

(2009), the active species or the photogenerated hole attack the central carbon to<br />

decolorize the dye.<br />

After that, the degradation continues with any N-deethylation<br />

intermediates and other smaller molecules until the mineralization process is finished


114<br />

with the formation of CO 2 and H 2 O. Yu et al. (2009) determined that 97% of an RhB<br />

solution was completely bleached in three hours using NaBiO 3 as catalyst.<br />

According to Vinu and Madras (2009), N-demethylation and N-dealkylation are<br />

the mechanisms involved in the synthesis of intermediates during the degradation of the<br />

triphenyl methane dyes (i.e. RhB). For RhB, the characteristic absorption peak at 554<br />

nm decreases during the photocatalytic degradation and a concomitant hypsochromic<br />

peak appears at 534 nm. These shifts are part of the formation and transformation of<br />

the N-deethylated intermediates and imply that the chromophores of the RhB molecules<br />

are cleavaged (Yu et al. 2009). Yu et al. (2009) identify four different N-deethylated<br />

intermediates; N,N-diethyl-N’-ethylrhodamine (DER), N,N-diethylrhodamine (DR), N-<br />

ethylrhodamine (ER), rhodamine (R) and other small molecular intermediates (18<br />

compounds) as ethane-1,2-diol, benzoic acid, glutaric acid and dibutyl phthalate.<br />

During the degradation process, a competition between the DR and EER (two peaks at<br />

m/z=387 appear) occurs, but the DR domain over EER (Yu et al. 2009). The change of<br />

550 nm to 508 nm was correlated with the hypsochromic shift; the intermediates of<br />

degradation of the N-ethyl occurred at shorter wavelength due to their auxochromic<br />

properties (Yu et al. 2009).<br />

An oxidative cleavage in the carbons near the azo bond forms the primary<br />

products of degradation of a dye, which has an azo bond (Vinus and Madras 2009) as<br />

in MO. The four principal by-products are depicted in Figure 5.14. Four possible byproducts<br />

could be principally generated during the photodegradation process, along<br />

with other low molecular weight compounds (He et al. 2009; Hong et al. 2009).<br />

During the degradation of the MO using CaSb 2 O 5 (OH) 2 as photocatalyst, Sun et<br />

al. (2009) determined other possible intermediates (Figure 5.15).<br />

Possibilities of<br />

different intermediates can occur due to the influence of the catalyst, specially the<br />

active sites of the catalyst, used during the reaction.


115<br />

Wang et al. (2010) observed a shift from 655nm to 613nm (to the blue region of<br />

the spectrum) during the degradation process, suggesting the N-demethylation of the<br />

MB dye. According to the authors (Wang et al. 2010) the methyl groups of the dyes are<br />

removed one by one of the chromophore, modifying gradually the wavelength of the<br />

peak. The complete mineralization of the MB is described by the following reaction<br />

(Panizza et al. 2006):<br />

51OH<br />

C16H18ClN<br />

3S<br />

<br />

16CO<br />

2<br />

6H2O<br />

H2SO<br />

4<br />

3HNO<br />

3<br />

HCl<br />

But when an electrochemical process for the degradation of the MB dye is used the<br />

chlorine atom mediates the oxidation reaction (Panizza et al. 2006).<br />

Figure 5.14. Four principal by-products of the MO degradation process.


116<br />

CH 3<br />

O 3<br />

- S<br />

N N N<br />

CH 3<br />

Demethylation<br />

Demethylation<br />

Hydroxylation<br />

OH<br />

CH 3<br />

CH 3<br />

O 3<br />

- S<br />

N N N<br />

O 3<br />

- S<br />

N N N<br />

H<br />

H<br />

Opening-Ring<br />

Hydroxylation<br />

Carboxylation<br />

Openning-Ring<br />

O - OC N N N<br />

CH 3<br />

CH 3<br />

CH 3<br />

O - OC<br />

(C 2 H 5 ) 5<br />

N<br />

H<br />

.OH<br />

CO 2 , H 2 O, Low Molecular Weight Byproducts<br />

Figure 5.15.<br />

Possible intermediates of degradation of MO during the<br />

photocatalytic degradation (Adapted from Sun et al. 2009).


117<br />

5.3. Sono-Fenton Process<br />

5.3.1. Description of the Sono-Fenton System<br />

A similar method described by Hernández Enriquez et al. (2008) was used<br />

during this research. A reactor (semi-batch type) was incorporated in the center of an<br />

ultrasound bath. Similarly to the photochemical process, the homogeneous sample<br />

(catalyst and dye solution) was kept in the dark under stirring for at least 30 min (Hong<br />

et al. 2009) to reach the adsorption-desorption equilibrium (Zhou et al. 2010).<br />

All the system was covered to avoid any other irradiation source on the sample;<br />

only the energy of the sound waves could reach the sample. Every 10 minutes an<br />

aliquot sample of 10 mL was taken to determine the UV and fluorescence spectrum and<br />

to measure the TOC concentration. The concentration of the dye solutions was 10 -5 M<br />

and the concentration of the catalyst was 0.6 gL -1 (Velegraki and Mantvinos 2008; Asiri<br />

et al. 2011). Different organic pollutants (dyes and organic compounds) with different<br />

structures were used during the investigation.<br />

The organic pollutants used were<br />

Methylene Blue (MB), Rhodamine B (RhB), Methyl Orange (MO), Crystal Violet (CV),<br />

Methyl Violet (MV) and p-aminobenzoic acid (pABA).<br />

The sonochemical process is similar to the photochemical process because<br />

different radicals are produced and, after that, the radicals react in a cascade of<br />

reactions to degrade the organic compounds. Vinu and Madras (2009) observed an<br />

order in the degradation processes with a synergistic effect as follows: UV +US > UV<br />

only > US only. According to Seymour and Gupta (1997) the process occurs when:<br />

“the heat from cavity implosion decompose water into extremely reactive<br />

hydrogen atoms (H·) and hydroxyl radicals (OH·).<br />

During the quick<br />

cooling phase, hydrogen atoms and hydroxyl radicals recombine to form<br />

hydrogen peroxide (H 2 O 2 ) and molecular hydrogen (H 2 )…”.


118<br />

The process is summarized in the following diagram (Figure 5.16).<br />

That reaction<br />

encourages the decomposition of the organic pollutants and the reduction or oxidation<br />

of the inorganic pollutants.<br />

Development of supercritical areas is relevant for the<br />

reaction rate. Fluorescence, absorption and TOC were determined for each sample. A<br />

decrease in the intensity of the absorption and fluorescence was observed for all the<br />

analyzed compounds along the reaction time. Figure 5.17 shows the fluorescence and<br />

absorption spectra.<br />

The absorbance and fluorescence curves clearly show the<br />

degradation process observed when RhB is treated in a sono-Fenton degradation<br />

process (Figure 5.17a and Figure 5.17b). Figure 5.17c and Figure 5.17d clearly show<br />

how the area of the curve decreases along the reaction time.<br />

Figure 5.16. Schematic diagram of the sonochemical<br />

generation of the degradation radicals (Adapted from<br />

Minero et al. 2005; Dafnopatidou et al. 2007).


Absorbance<br />

Fluorescence<br />

119<br />

0.7<br />

0.6<br />

1.0<br />

a) b)<br />

0.8<br />

0.5<br />

0.4<br />

C/C 0<br />

0.6<br />

0.4<br />

0.3<br />

0.2<br />

0.2<br />

0 10 20 30 40 50 60<br />

0.0<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

c) d)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure 5.17. Degradation curves of RhB; UV-vis absorbance (a), TOC<br />

(b), fluorescence (c) and dye solutions before (left) and after (right) the<br />

sono-Fenton degradation process (d).<br />

The Table 5.4 shows the percent of degradation (based on the decrease in TOC<br />

concentration).<br />

The determination of the degradation percent was similar to the<br />

process used during the photocatalytic process.<br />

According with the data presented in Table 5.4, the FeCl 2 has the highest<br />

degradation rate, reaching values between 70.82% to 96.85%. For the degradation of<br />

p-ABA, the most efficient catalyst was the FeCl 2 catalyst (95.20% of degradation). The<br />

high efficiency of the FeCl 2 as catalyst could be due to the high solubility of this catalyst<br />

with respect to the other catalysts used in this study. The catalyst with the lowest<br />

degradation rate was CuO, with a degradation percent ranging from 48.08% to 70.80%.<br />

The order of efficiency was: FeCl 2 > Fe 2 O 3 NWs > Fe 3 O 4 Comp > CuO (Figure 5.18).


Degradation (%)<br />

120<br />

Table 5.4. Degradation percent of dye solutions during the Sono-Fenton Process.<br />

Mo<strong>del</strong> Organic<br />

Contaminant<br />

Catalyst (Sono-Fenton)<br />

CuO Fe 2 O 3 NWs Fe 3 O 4 Mag FeCl 2<br />

MB 67.51% 84.87% 81.97% 94.37%<br />

RhB 65.86% 79.49% 71.23% 90.24%<br />

MO 65.03% 83.21% 73.71% 96.85%<br />

CV 48.08% 64.20% 58.00% 70.82%<br />

MV 65.44% 84.87% 74.12% 94.37%<br />

p-ABA 70.80% 77.80% 68.70% 95.20%<br />

100<br />

95<br />

90<br />

85<br />

80<br />

75<br />

70<br />

65<br />

60<br />

55<br />

50<br />

45<br />

MB RhB MO CV MV p-ABA<br />

Pollutants<br />

CuO<br />

FeNWs<br />

FeComp<br />

FeCl2<br />

Figure 5.18. Graphic of percent of degradation of<br />

the organic compounds by the Sono-Fenton<br />

process.


Ln (C/C 0<br />

)<br />

121<br />

According to the authors (Vinus and Madras 2009) the mo<strong>del</strong> of<br />

sonophotocatalytic degradation of a sulfonated azo dye is a pseudo-first order reaction.<br />

This degradation is also considered as a Dual-Pathway Mo<strong>del</strong> in which the<br />

sonocatalytic and photocatalytic processes are included in four different pathways as:<br />

the absorption-desorption equilibrium, the generation of charge-carriers, generation of<br />

electron-hole pairs and the radical formation (Vinu and Madras 2009). Equation used to<br />

determine the reaction rate is based on the definition of the mo<strong>del</strong>. First kinetic mo<strong>del</strong><br />

is defined as (Wang et al. 2006; Asiri et al. 2011):<br />

ln( C0 C) kt<br />

where C 0 and C are initial concentration and concentration at different times,<br />

respectively. Semilogarithmic plots of the concentrations vs time gave straight lines in<br />

which the slopes represent the value of k (reaction rate) (Figure 5.19).<br />

0.2<br />

0.0<br />

MB/Fe 3 O 4 /Sono-Fenton<br />

-0.2<br />

-0.4<br />

-0.6<br />

-0.8<br />

y = -0.0293x<br />

R² = 0.9812<br />

-1.0<br />

-1.2<br />

-1.4<br />

-1.6<br />

-1.8<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure 5.19. Regression curve of the Methylene<br />

Blue (MB) with Fe 3 O 4 under sono-Fenton<br />

process.


122<br />

The Table 5.5 shows the data corresponding to the kinetic reaction rates of the<br />

mo<strong>del</strong> organic compounds and the different catalysts used during the degradation<br />

processes. The mean velocity for the reaction is approximately 10 -2 min. The values of<br />

R 2 range from 0.98 and 0.77. The difference between these values should be caused<br />

by the adsorption-desorption process of the dye by the catalyst during the degradation<br />

process.<br />

Table 5.5 Kinetic reaction rates and R 2 values for the degradation reaction of the<br />

organic compounds during the sono-Fenton process.<br />

Mo<strong>del</strong><br />

Organic<br />

Catalyst (Sono-Fenton)<br />

CuO Fe 2 O 3 NWs Fe 3 O 4 Mag FeCl 2<br />

Contaminant<br />

MB 2.20 x 10 -2<br />

3.35 x 10 -2<br />

2.93 x 10 -2<br />

5.25 x 10 -2<br />

R 2 = 0.8621<br />

R 2 =0.9650<br />

R 2 = 0.9812<br />

R 2 = 0.9455<br />

RhB 2.05 x 10 -2<br />

2.97 x 10 -2<br />

2.41 x 10 -2<br />

4.48 x 10 -2<br />

R 2 = 0.8350<br />

R 2 =0.9241<br />

R 2 = 0.7979<br />

R 2 = 0.9068<br />

MO 2.09 x 10 -2<br />

3.19 x 10 -2<br />

2.45 x 10 -2<br />

5.32 x 10 -2<br />

R 2 = 0.8622<br />

R 2 =0.9510<br />

R 2 = 0.9214<br />

R 2 = 0.9784<br />

CV 1.30 x 10 -2<br />

2.03 x 10 -2<br />

1.71 x 10 -2<br />

2.53 x 10 -2<br />

R 2 = 0.7797<br />

R 2 =0.8348<br />

R 2 = 0.8172<br />

R 2 = 0.8399<br />

MV 1.89 x 10 -2<br />

3.21 x 10 -2<br />

2.46 x 10 -2<br />

5.05 x 10 -2<br />

R 2 = 0.8764<br />

R 2 =0.9771<br />

R 2 = 0.9445<br />

R 2 = 0.9880<br />

p-ABA 2.24 x 10 -2<br />

2.81 x 10 -2<br />

2.12 x 10 -2<br />

5.37 x 10 -2<br />

R 2 = 0.9757<br />

R 2 =0.8745<br />

R 2 = 0.9387<br />

R 2 = 0.9794


123<br />

According to Vinu and Madras (2009), the degradation reaction of a dye by a<br />

hydroxyl radical generated by UV irradiation of ultrasonic is as follows:<br />

TiO 2 (OH·) ads – D ads + TiO 2 – D ads (orD) → intermediates (P) → CO 2 + H 2 O<br />

Dafnopatidou et al. (2007) describe the molecular environment during the ultrasound<br />

degradation.<br />

H 2 O + ultrasound → ·OH + ·H<br />

2·OH → H 2 O 2<br />

Dyestuff + ·OH → products<br />

According to Dafnopatidou et al. (2007), after the decolorization process by sonolysis, a<br />

water effluent could be reused because it complies with the environmental regulations.<br />

Authors (Wang et al. 2003) studied the exponential decrease of the methyl violet<br />

with the sonication time, showing that the reaction process had a first order degradation<br />

reaction with a reaction rate coefficient of 1.35 x 10 -2 min -1 at 20 + 1 °C. Besides, they<br />

showed that the degradation process decreased, when the temperature of the solution<br />

increased to 80 °C because the cavitation bubbles decrease in the solution.<br />

According to Wang et al. (2003), during the aqueous sonochemical process<br />

three regions could be observed: the first one is the gas phase (formation of small<br />

bubbles) in which high temperature and pressure are produced; the second one is an<br />

interfacial zone between the cavitation bubble and the aqueous phase, in which the<br />

temperature is lower than in the gas phase; and the third one is the bulk solution in<br />

which the reaction takes place (Figure 5.20). The pH of the solution influences the<br />

degradation of the dyes; for instance, at lower pH (pH 2 to 4) increases the degradation<br />

rate (Wang et al. 2003).


124<br />

Cavity<br />

Interface<br />

Bulk (Liquid Media)<br />

Figure 5.20. Scheme of the different areas<br />

of interest during the sonochemical<br />

process (Adapted from Seymour and<br />

Gupta 1997).<br />

5.4. Photo-Fenton Process<br />

5.4.1. Description of the Photo-Fenton System<br />

A similar method described by Hernández Enríquez et al. (2008) was used<br />

during the photo-Fenton process. The photo-Fenton process used during this research<br />

is quite similar to the photochemical process; the difference between them is the use of<br />

an iron catalyst (with the exception of the CuO). A cylindrical reactor (semi-batch type)<br />

with continuous stirring was located in the center of two double tubular lamps which are<br />

the irradiation source. The system (Figure 5.01) was composed by two annular white<br />

bulb lights, with a total power of 60 watts. A vessel of 1 L was used during the<br />

irradiation of the sample.<br />

The sample was mechanically stirred with a paddler to<br />

maintain a homogeneous mixture during the irradiation of the sample.<br />

Before the<br />

irradiation, the particles (catalyst and dye) were suspended in the solution and kept in


125<br />

the dark under stirring for at least 30 min (Hong et al. 2009), to reach the adsorptiondesorption<br />

equilibrium (Zhou et al. 2010).<br />

The experimental system was covered to avoid any other irradiation source in<br />

the sample; only the light of the bulbs could reach the sample. Every 10 minutes an<br />

aliquot sample of 10 mL was taken to determine the UV and fluorescence spectrum and<br />

to measure the TOC concentration.<br />

As in the other catalytic reactions, the<br />

concentration of the dye was 10 -5 M and the concentration of the catalyst was 0.6 gL -1<br />

(Velegraki and Mantvinos 2008; Asiri et al. 2011).<br />

Different organic pollutants (dyes and organic compounds) with different<br />

structures were used during the investigation.<br />

The organic pollutants used were<br />

Methylene Blue (MB), Rhodamine B (RhB), Methyl Orange (MO), Crystal Violet (CV),<br />

Methyl Violet (MV) and p-aminobenzoic acid (pABA)<br />

According to Lodha et al. (2008), the photo-Fenton process is a new method for<br />

the degradation of contaminants as dyes. This process is described as a classical<br />

photochemical reaction, which involves the presence of the iron ion, hydrogen peroxide<br />

and the visible or UV radiation.<br />

One of the disadvantages of the Fenton process is the cease of the reaction<br />

when the Fe 2+ is consumed but if the process is carried out in the presence of light the<br />

Fenton process is cyclic, and the reaction continues, because the Fe 2+ is regenerated<br />

from Fe 3+ in the presence of light (Lodha et al. 2008). The Fe 2+ reacts with the H 2 O 2 ,<br />

decomposing the peroxide in ·OH radical OH - and oxidize the iron ion forming Fe 3+ .<br />

The ferric ion decomposes the water molecule, forming ·OH radical and the iron ion is<br />

reduced to Fe 2+ (Lodha et al. 2008). Some authors (Garrriga I Cabo 2007; Lodha et al.<br />

2008) indicate that in the Fenton reactions some ferryl complex and hydrocomplexes of<br />

iron could be involved resulting in the formation of Fe 2+ and ·OH radicals.


Fluorescence<br />

Absorbance<br />

126<br />

Fluorescence, absorption and TOC were determined for each sample.<br />

A<br />

decrease in the intensity of the absorption and fluorescence signals was observed for<br />

all the systems along the reaction time. Figure 5.21 shows the fluorescence and UV-vis<br />

absorption spectra of MO.<br />

Absorbance and fluorescence curves clearly show the<br />

degradation process observed when MO is treated with FeCl 2 (Figure 5.21a and Figure<br />

5.21b). Figure 5.21c and Figure 5.21d show how the areas of the curves decrease<br />

along the reaction time.<br />

0.8<br />

0.7<br />

0.6<br />

1.0<br />

a) b)<br />

0.8<br />

0.5<br />

0.6<br />

0.4<br />

0.3<br />

C/C 0<br />

0.4<br />

0.2<br />

0.2<br />

0.1<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

c) d)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure 5.21. Degradation curves of MO; UV-vis absorbance (a), TOC<br />

(b), fluorescence (c) and dye solution before (left) and after (right) the<br />

photo-Fenton process (d).<br />

The Table 5.6 shows the percent of degradation (based on the decrease in TOC<br />

concentration) of the organic pollutants. The determination of the degradation percent<br />

was similar to the process used during the sonochemical process;


127<br />

According with the results on the Table 5.6, the most efficient catalyst was the<br />

FeCl 2 , which was a commercial material. The degradation percent using FeCl 2 ranged<br />

from 79.49% to 98.10%. The less efficient catalyst was the CuO with degradation<br />

percents from 50.15 to 58.00 %. The order of efficiency was: FeCl 2 > Fe 2 O 3 NWs ><br />

Fe 3 O 4 Comp > CuO, similar to the sono-Fenton process (Figure 5.22).<br />

Table 5.6. Degradation percent of dye solution during the Photo-Fenton Process<br />

Mo<strong>del</strong> Organic<br />

Contaminant<br />

Catalyst (Photo-Fenton)<br />

CuO Fe 2 O 3 NWs Fe 3 O 4 Mag FeCl 2<br />

MB 57.18% 86.11% 73.30% 92.31%<br />

RhB 53.87% 79.49% 65.44% 86.93%<br />

MO 57.59% 84.45% 65.44% 93.94%<br />

CV 50.15% 58.00% 50.15% 79.49%<br />

MV 55.55% 78.25% 65.03% 89.41%<br />

p-ABA 58.00% 86.10% 69.20% 98.10%<br />

Similarly to the photochemical and sono-Fenton process, the first kinetic mo<strong>del</strong><br />

was used to determine the reaction rate for the photo-Fenton process. The equation is<br />

defined as (Wang et al. 2006; Asiri et al. 2011):<br />

ln( C0 C) kt<br />

The semilogarithmic plots of the concentrations vs time were used to determine the<br />

reaction rate (Figure 5.23). The Table 5.7 shows the k values for the degradation<br />

reaction of the dyes during the photo-Fenton process.


Ln (C/C 0<br />

)<br />

Degradation (%)<br />

128<br />

100<br />

95<br />

90<br />

85<br />

80<br />

75<br />

70<br />

65<br />

60<br />

55<br />

50<br />

MB RhB MO CV MV p-ABA<br />

Pollutants<br />

CuO<br />

FeNWs<br />

FeComp<br />

FeCl2<br />

Figure 5.22. Graphic of degradation percent of the<br />

organic compounds by the Photo-Fenton process.<br />

0.0<br />

MB/Fe 3 O 4 /Photo-Fenton<br />

-0.5<br />

-1.0<br />

y = -0.0365x<br />

R² = 0.9582<br />

-1.5<br />

-2.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure 5.23. Regression curve of the Methylene Blue<br />

(MB) with Fe 3 O 4 during the photo-Fenton degradation<br />

process.<br />

The Table 5.7 shows the data of the kinetic reaction rates for the mo<strong>del</strong> organic<br />

compounds and the catalysts used during the degradation processes.<br />

The mean<br />

velocity for the reaction is approximately 10 -2 min. The values of R 2 ranged from 0.99 to


129<br />

0.51. The difference between these values could be due to the adsorption-desorption<br />

processes of the dye by the catalyst during the degradation process and the<br />

agglomeration of the catalyst during the reaction.<br />

Table 5.7. Kinetic reaction rates and R 2 values for the degradation reaction of the<br />

organic compounds during the photo-Fenton process.<br />

Mo<strong>del</strong> Organic<br />

Contaminant<br />

Catalyst (Photo-Fenton)<br />

CuO Fe 2 O 3 NWs Fe 3 O 4 Mag FeCl 2<br />

MB 1.66 x 10 -2<br />

3.65 x 10 -2<br />

2.37 x 10 -2<br />

4.67 x 10 -2<br />

R 2 = 0.7264<br />

R 2 =0.9582<br />

R 2 = 0.9150<br />

R 2 = 0.9241<br />

RhB 1.52 x 10 -2<br />

3.43 x 10 -2<br />

2.02 x 10 -2<br />

4.17 x 10 -2<br />

R 2 = 0.5138<br />

R 2 =0.7527<br />

R 2 = 0.7936<br />

R 2 = 0.6288<br />

MO 1.53 x 10 -2<br />

3.78 x 10 -2<br />

2.01 x 10 -2<br />

5.28 x 10 -2<br />

R 2 = 0.8749<br />

R 2 =0.8905<br />

R 2 = 0.8983<br />

R 2 = 0.9428<br />

CV 1.16 x 10 -2<br />

1.69 x 10 -2<br />

1.34 x 10 -2<br />

3.09 x 10 -2<br />

R 2 = 0.9655<br />

R 2 =0.7964<br />

R 2 = 0.8548<br />

R 2 = 0.8473<br />

MV 1.46 x 10 -2<br />

2.64 x 10 -2<br />

1.86 x 10 -2<br />

3.96 x 10 -2<br />

R 2 = 0.9181<br />

R 2 =0.9712<br />

R 2 = 0.8993<br />

R 2 = 0.9854<br />

p-ABA 1.72 x 10 -2<br />

3.45 x 10 -2<br />

2.09 x 10 -2<br />

6.63 x 10 -2<br />

R 2 = 0.8045<br />

R 2 =0.9784<br />

R 2 = 0.9808<br />

R 2 = 0.9941<br />

5.5. Statistical analysis<br />

A Multiple Factorial Design was used for the statistical analysis of the results of<br />

the three degradation mechanisms: Photocatalysis, Sono-Fenton and Photo-Fenton


% Degradation<br />

130<br />

processes. The photochemical process was analyzed independently from the Sono-<br />

Fenton and Photo-Fenton, because different catalysts were used during this process. A<br />

complex matrix 2x4x6 was generated using the Minitab 14 program. The 48 responses<br />

were analyzed. According with the results obtained, no significant differences were<br />

observed between the sono-Fenton and photo-Fenton processes.<br />

A comparison between the results of the degradation processes between the<br />

photocatalytic and photo-Fenton was carried out. A few differences were observed<br />

when both processes were compared (Figure 5.24). TiO 2 NWs was the most effective<br />

catalyst for the photocatalytic process and the FeCl 2 was the catalyst with higher<br />

degradation activity for the sono and photo-Fenton processes.<br />

100<br />

80<br />

Anatase<br />

Rutile<br />

TiO2MWCNTs<br />

TiO2NWs<br />

ZnO<br />

CuO<br />

FeNWs<br />

FeComp<br />

FeCl2<br />

60<br />

40<br />

20<br />

0<br />

MB RhB MO CV MV p-ABA<br />

Dye<br />

Figure 5.24.<br />

Graphic of comparison between the<br />

Photocatalytic process and the Photo-Fenton process.<br />

Figure 5.25 represents the sono-Fenton and photocatalytic process for MB, RhB<br />

and MO. A similar pattern between photocatalysis and sono-Fenton was observed.<br />

The percentual difference between the three studied processes (photocatalysis, photo-


Degradation (%)<br />

131<br />

Fenton and Sono-Fenton) was minimal. In CV, some differences between the sono-<br />

Fenton and photo-Fenton processes were observed.<br />

100<br />

80<br />

Anatase<br />

Rutile<br />

TiO 2<br />

@MWCNTs<br />

TiO 2<br />

NWs<br />

ZnO<br />

CuO<br />

Fe 2<br />

O 3<br />

NWs<br />

Fe 3<br />

O 4<br />

Mag<br />

FeCl 2<br />

60<br />

40<br />

20<br />

0<br />

MB RhB MO<br />

Dyes<br />

Figure 5.25.<br />

Graphic of comparison between the<br />

Photocatalytic process and the Sono-Fenton process for<br />

MB, RhB and MO.


Chapter Six<br />

Conclusion<br />

After the analysis of the data obtained during this investigation we can conclude<br />

that the goal of this investigation was achieved. During the present research, different<br />

catalysts (TiO 2 nanowires, TiO 2 @MWNTs, ZnO nanoparticles, Fe 2 O 3 nanowires and<br />

magnetite nanoparticles) were synthesized and fully characterized by different<br />

techniques as FE-SEM, TGA, specific surface area (BET), XRD, Raman spectroscopy,<br />

XPS and magnetic susceptibility. Commercial and synthesized catalysts were used in<br />

different processes with the aim to reduce the amount of mo<strong>del</strong> compounds (organic<br />

dyes) in water, by using different heterogeneous catalytic processes (photocatalysis,<br />

sono-Fenton and photo-Fenton). As mo<strong>del</strong> pollutants, we selected different dyes or<br />

organic compounds that are considered as hazardous contaminants, normally used by<br />

the chemical industry (Methylene Blue, Rhodamine B, Methyl Orange, Gential Violet and<br />

Methyl Violet and p-aminobenzoic acid).<br />

In all cases, the catalysts used in the present research were able to degrade the<br />

pollutants. For the photocatalytic process, the most effective catalyst was the TiO 2 NWs<br />

(approximately 94.78% of degradation) and the less effective was the TiO 2 @MWCNTs<br />

(with approximately 78.04% of degradation). During the photo-Fenton and sono-Fenton<br />

processes the same catalysts were used, to demonstrate if any of the processes was<br />

more effective than the other.<br />

However, no significant differences were observed<br />

between photo-Fenton and sono-Fenton processes when the same catalysts were<br />

studied and compared. A slightly decrease in the degradation percent was observed for<br />

CV pollutant. For the sono-Fenton and photo-Fenton processes, the more efficient<br />

catalyst was, in both cases, FeCl 2 (with approximately 90.31% and 90.03% of<br />

132


133<br />

degradation, respectively) and the less effective was CuO (approx. 63.79% and 55.39%<br />

of degradation, respectively).<br />

Hence, it is deduced that the catalytic reactions studied in this research can be<br />

efficiently used for the degradation and decolorization of organic pollutants. The catalytic<br />

processes can be suitably and cost effectively employed for the removal of pollutants<br />

from wastewaters in a short period of time. We can predict that, with high probability,<br />

these catalytic processes can be implemented as appropriate chemical procedures for<br />

pollutant removal from water or even from soil.


Literature Cited<br />

Aarthi T, Madras G. 2007. Photocatalytic Degradation of Rhodamine Dyes with Nano-<br />

TiO 2 . Ind Eng Chem Res. 46(1):7-14.<br />

Ab<strong>del</strong>malek F, Ghezzar MR, Belhadj M, Addou A, Brisset JL. 2006. Bleaching and<br />

Degradation of Textile Dyes by Nonthermal Plasma Process at Atmospheric<br />

Pressure. Ind Eng Chem Res. 45(1):23-29.<br />

Ai Z, Lu L, Li J, Zhang L, Qiu J, Wu M. 2007a. Fe@Fe 2 O 3 Core-Shell Nanowires as Iron<br />

Reagent. 1. Efficient Degadation of Rhodamine B by a Novel Sono-Fenton<br />

Process. J Phys Chem C. 111(11):4087-4093.<br />

Ai Z, Lu L, Li J, Zhang L, Qiu J, Wu M. 2007b. Fe@Fe 2 O 3 Core-Shell Nanowires as Iron<br />

Reagent. 2. An Efficient and Reusable Sono-Fenton System Working at Neutral<br />

pH. J Phys Chem C. 111(20):7430-7436.<br />

Albers WM, Annila A, Barcelo D, Butler L, Crighton JS, Cullum B, Davies AN, Emsley L,<br />

Fetzer JC, Fligge TA, Fujii T, Goddard NJ, Green J, Hadden CE, Hagaman EW,<br />

Hassell C, Hof M, Janssens K, Lane DA, Mackova A, Martin GE, Morton S,<br />

Panne U, Patonay G, Przybylski M, Robien W, Rosenberg E, Russell DJ,<br />

Sablinskas V, Sablier M, Salzer R, Schroeder HF, Soini E, Steiner G, Thiele S,<br />

Varmuza K, Vo-Dinh T, Volka K, Walker CGH, Weinmann W, Weisenberger LA.<br />

Handbook of Spectroscopy. [Gauglitz G, Vo-Dinh T, editors. 2003]. Weinheim<br />

(FRG): Wiley-VCH. 2003.<br />

Anson Casaos A. 2005. Nanotubos de Carbono: Estructura porosa y sus implicaciones<br />

en el campo de la energía [Carbon nanotubes: Porous structure and their<br />

implications in the energy field]. [dissertation]. [Zaragoza (Z)]: <strong>Universidad</strong> de<br />

Zaragoza; 2005. Spanish.<br />

134


135<br />

Arribas MA, Marquez F, Martinez A. 1999. Activity, Selectivity, and Sulfur Resistance of<br />

Pt/WO x -ZrO 2 and Pt/Beta Catalyst for the Simultaneous Hydroisomerization of n-<br />

Heptane and Hydrogenation of Benzene. J. Catal. 190:309-319.<br />

Asiri AM, Al-Amoudi MS, Al-Talhi TA, Al-Talhi AD. 2011. Photodegradation of<br />

Rhodamine 6G and phenol red by nanosized TiO 2 under solar irradiation. J Saudi<br />

Chem Soc. 15:121-128.<br />

Bae JY, Yun TK, Han SS, Min BG, Bako I, Suh SH. 2009. Fabrication and<br />

Characteristics of Variable Crystalline TiO 2 for Photocatalytic Activities. Revue<br />

Roumaine de Chimie. 54(9):773-778.<br />

Bal S. 2010. Dispersion and reinforcing mechanism of carbon nanotubes in epoxy<br />

nanocomposites. Bull Mater Sci. 33(1):27-31.<br />

Barnes KK, Kolpin DW (United State Geological Survey). 2002. Phamaceuticals,<br />

Hormones, and Other Organic Wastewater Contaminants in U.S. Streams.<br />

[Internet]. Department of the Interior (US). USGS Fact Sheet FS-027-02.<br />

Available from: http://toxic.usgs.gov/pub/FS-027-02/pdf/FS-027-02.pdf.<br />

Barnes KK, Kolpin DW, Meyer MT, Thurman EM, Furlong ET, Zaugg SD, Barber B<br />

(United State Geological Survey). 2002. Water-Quality data for Phamaceuticals,<br />

Hormones, and Other Organic Wastewater Contaminants in U.S. Streams, 1999-<br />

2000. [Internet] Iowa City (IA): Department of the Interior (US). Open-File Report<br />

02-94. Available from: http://toxics.usgs.gov/pubs/OFR-02-94/.<br />

Barrera-Diaz C, Linares-Hernández I, Roa-Morales G, Bilyeu B, Balderas-Hernández P.<br />

2009. Removal of Biorefractory Compounds in Industrial Wastewater by<br />

Chemical and Electrochemical Pretreatments. Ind Eng Chem Res. 48(3):1253-<br />

1258.


136<br />

Barrett EP, Joyner LG, Halenda PP. 1951. The Determination of Pore Volume and Area<br />

Distribution in Porous Substances. I. Computations from Nitrogen Isotherms. J.<br />

Am. Chem. Soc. 73:373-380.<br />

Behnajady M, Modirshahla N, Ghazalian E. 2011. Synthesis of ZnO Nanoparticles at<br />

Different Conditions: A Comparison of Photocatalytic Activity. [Internet] Digest J<br />

Nanomat Biostructure. 6(1):467-474. Available from:<br />

http://www.chalcogen.infim.ro/467_Behnajadi.pdf.<br />

Bonilla C, Garcia A, Duconge J, Campo T, Elizalde E, Morant C, Marquez F. 2011.<br />

Synthesis and Characterization of Crystalline Fe2O3 Nanowires by Oxidation of<br />

Iron Substrates. Poster session presented at: 43rd IUPAC World Chemistry<br />

Congress, San Juan, PR.<br />

Buser HR, Balmer ME, Schmid P, Kohler M. 2006. Ocurrence of UV Filters 4-<br />

Methylbenzylidene Camphor and Octocrylene in Fish from Various Swiss Rivers<br />

with Inputs from Wastewater Treatment Plants. Environ Sci Technol. 40(5):1427-<br />

1431.<br />

Cao G. 2004. Growthn of Oxide Nanorod Arrays through Sol Electrophoretic Deposition.<br />

J Phys Chem B. 108(52):19921-19931.<br />

Cao SW, Zhu YJ, Cheng GF, Huang YH. 2010. Preparation and photocatalytic property<br />

of α-Fe 2 O 3 Hollow core/shell hierarchical nanostructures. J Phys Chem Solids.<br />

71:1680-1683.<br />

Chandra V, Park J, Chun Y, Lee JW, Hwang IC, Kim KS. 2010. Water-Dispersible<br />

Magnetite-Reduced Graphene Oxide Composite for Arsenic Removal. ACS<br />

Nano. 4(7):3979-3986.<br />

Chowdhury D, Paul A, Chattopadhyay A. 2005. Photocatalytic Polypyrrole-TiO 2 -<br />

Nanoparticles Composite Thin Film Generated at the Air-Water Interface.<br />

Langmuir. 21(9):4123-4128.


137<br />

Confortin D, Neevel H, van Bommel MR, Reissland B. Study of the degradation of an<br />

early synthetic dye, Crystal Violet [internet].Padua (ITA): Padua University. The<br />

Netherlands Institute for Cultural Heritage (ICN), Amsterdam. Available from:<br />

http://www.create.uwe.ac.uk/norway_paperlist/confortin.pdf.<br />

Corma A, Palomares A, Márquez F. 1997. Determining the Nature of the Active Sites of<br />

Cu-Beta Zeolites for the Selective Catalytic Reduction (SCR) of NO x by Using a<br />

Coupled Reaction-XAES/XPS Study. J. Catal. 170:132-139.<br />

Corma A, Palomares A, Rey F, Marquez F. 1997. Simultaneous Catalytic Removal of<br />

SO x and NO x with Hydrotalcite-Derived Mixed Oxides Containing Copper, and<br />

Their Possibilities to be Used in FCC Units. J. Catal. 170:140-149.<br />

Cotto M, Campo T, Elizalde E, Morant C, Marquez F. 2011. Hydrothermal Synthesis and<br />

Photocatalytic Activity of Titanium Oxide Nanowires. Poster session presented<br />

at: 43rd IUPAC World Chemistry Congress 2011, San Juan, PR.<br />

Dafnopatidou EK, Gallios GP, Tsatsaroni EG, Lazaridos NK. 2007. Reactive Dyestuffs<br />

Removal from Aqueous Solution by Flotation, Possibility of Water Reuses, and<br />

Dyestuff Degradation. Ind Eng Chem Res. 46(7):2125-2132.<br />

Daou TJ, Begin-Colin S, Grenèche JM, Thomas F, Derory A, Bernhardt P, Legaré P,<br />

Pourroy G. 2007. Phosphate Adsorption Properties of Magnetite-Based<br />

Nanoparticles. Chem Mater. 19(18):4494-4505.<br />

Daou TJ, Pourroy G, Begin-Colin, Greneche JM, Ulhaq-Bouillet C, Legare P, Bernhardt<br />

P, Leuvrey C, Rogez G. 2006. Hydrothermal Synthesis of Monodisperse<br />

Magnetite Nanoparticles. Chem Mater. 18(18):4399-4404.<br />

Devipriya SP, Yesodharan S. 2010. Photocatalytic degradation of phenol in water using<br />

TiO 2 and ZnO. J Environ Biol. 31(3): 247-249.


138<br />

Dixit, A. Investigation on the structural dielectric and ferroelectric properties of sol-gel<br />

derived zirconium substituted barium titanate thin films [dissertation]. Rio Piedras<br />

(PR): University of Puerto Rico. 2003.<br />

Doi T, Miwa Y, Iriyama Y, Abe T, Ogumi Z. 2009. Single-Step Synthesis of Nanosized<br />

Titanium-Based Oxide/Carbon Nanotube Composites by Electrospray Deposition<br />

and Their Electrochemical Properties. J Phys Chem C. 113(18):7719-7722.<br />

Eichenseher T. March 1, 2006. The cloudy side of sunscreens. Environmental News.<br />

Environ Sci Technol. 1377-1378.<br />

Flewitt PEJ, Wild RK. Physical Methods for Materials Characterization. 2nd ed.<br />

Phila<strong>del</strong>phia (PA): 2003. Inst Phys Publishing. p. 43, 89-90, 274.<br />

FR: Federal Ragister. Notice of Final 2010 Effluent Gui<strong>del</strong>ines Program Plan.<br />

FR76(207): 66286-66304.<br />

Fundamental XPS Data from Pure Elements, Pure Oxides, and Chemical Compounds.<br />

[Internet] 1999. Available from: http://home.postech.ac.kr/~chey/paper/fundxps.<br />

pdf.<br />

Garriga I Cabo, C. Estrategias de optimizacion de procesos de descontaminacion de<br />

efluentes acuosos y gaseosos mediante fotocatalisis heterogenea. [Strategies for<br />

the optimization of the decontamination of aqueous and gaseous effluents using<br />

heterogeneous photocatalysis]. [dissertation].[Canarias (GC)]. Jun 2007.<br />

Spanish.<br />

Gaulke LS, Strand SE, Kalhorn TF, Stensel HD. 2009. Estrogen Biodegradation Kinetics<br />

and Estrogenic Activity Reduction for Two Biological Wastewater Treatment<br />

Methods. Environ Sci Technol. 43(18):7111-7116.<br />

Gersten JI, Smith FW. 2001. The Physics and Chemistry of Materials. [Internet] USA:<br />

John Wiley & Sons. Chapter W22; p. 413-419, 433-437, 443-445, 457-458.<br />

Available from: ftp://ftp.wiley.com/public/sci_tech_med/materials.


139<br />

Gupta VK, Suhas, Ali I, Saini VK. 2004. Removal of Rhodamine B, Fast Green, and<br />

Methylene Blue from Wastewater Using Red Mud, an Aluminum Industry Waste.<br />

Ind Eng Chem Res. 43(7):1740-1747.<br />

Hadj Salah N, Bouhelassa M, Bekkouche S, Boultif A. 2004. Study of photocatalytic<br />

degradation of phenol. Desalination. 166:347-354.<br />

Handbook of the Elements and Native Oxides. BE Lookup Table for Signals from<br />

Elements and Common Chemical Species [Internet]. 1999. Available from:<br />

http://www.xpsdata.com/XI_BE_Lookup_table.pdf<br />

He Y, Li D, Xiao G, Chen W, Chen Y, Sun M, Huang H, Fu X. 2009. A New Application<br />

of Nanocrystal In 2 S 3 in Efficient Degradation of Organic Pollutants under Visible<br />

Light Irradiation. J Phys Chem C. 113(13):5254-5262.<br />

Hernandez Enrique JM, Garcia Serrano LA, Zeifert Soares BH. 2008. Sintesis y<br />

Caracterizacion de Nanoparticulas de N-TiO 2 – Anatasa [Synthesis and<br />

Characterization of N-TiO 2 – Anatase Nanoparticles]. Superficies y Vacio.<br />

21(4):1-5. Spanish.<br />

Hernandez Rueda JJ. Control de la nanoestructura como procedimiento para optimizar<br />

las propiedades fisicas de materiales compuestos de matriz polimerica y<br />

nanotubos de carbon [Nanostructure control as a procedure for the optimization<br />

of the physical properties of materials composed of a polymeric matrix and<br />

carbon nanotubes] [dissertation]. [Madrid (M)]: <strong>Universidad</strong> Autonoma de Madrid;<br />

Feb 2010. Spanish.<br />

Hernandez-Alonso MD, Fresno F, Suarez S. Coronado JM. 2009. Development of<br />

alternative photocatalysts to TiO 2 : Challenges and opportunities. Energy Environ<br />

Sci. 2:1231-1257.


140<br />

Hong RY, Li JH, Chen LL, Liu DQ, Li HZ, Zheng Y, Ding J. 2009. Synthesis, surface<br />

modification and photocatalytic property of ZnO nanoparticles. Powder Technol.<br />

189:426-432.<br />

Horstman Fenton HJ, Jackson HJ. 1899. The Oxidation of Polyhydric Alcohols in<br />

Presence of Iron. J Chem Soc Trans. 75:1.<br />

Hotta Y, Hosoda A, Sano F, Wakayama M, Niwa K, Yoshikawa H, Tamura H. 2010.<br />

Ecotoxicity by the Biodegradation of Alkylphenol Polyethoxylates Depends on the<br />

Effect of Trace Elements. J Agric Food Chem. 58(2):1062-1067.<br />

Hu X, Li G, Yu JC. 2010. Design, Fabrication, and Modification of Nanostrutured<br />

Semiconductor Materials for Environmental and Energy Applications. Langmuir<br />

Invited Feature Article. 26(5):3031-3039.<br />

Huang J, Cui Y, Wang X. 2010a. Visible Light-Sensitive ZnGe Oxynitride Catalyst for the<br />

Decomposition of Organic Pollutants in Water. Environ Sci Technol. 44(9):3500-<br />

3504.<br />

Huang W, Ji Y, Yang Z, Feng X, Liu C, Zhu Y, Lu X. 2010b. Mineralization of Trace<br />

Nitro/Chloro/Methyl/Amino-Aromatic Contaminants in Wastewaters by Advanced<br />

Oxidation Processes. Ind Eng Chem Res. 49(13):6243-6249.<br />

Huh HH, Hoa NTQ, Nang LV, Kim ET. Thermal Growth and Visble-Light Photocatalysis<br />

of TiO 2 Nanowires. [Internet]. 2010. Department of Materials Science &<br />

Engineering; Chungnam National University.; Daejeon, (KOR). Available from:<br />

http://wjoe.hebeu.edu.cn/sup.2.2010/HIJ/Huh,%20Hoon-<br />

Hoe%20(Chungnam%20N.U.,%20Daejeon,%20S.%20Korea)%20%20271.pdf<br />

Hussein Sharif Zein S, Boccaccini AR. 2008. Synthesis and Characterization of TiO 2<br />

Coated Multiwalled Cabon Nanotubes Using a Sol Gel Method. Material and<br />

Interfaces. Ind Eng Chem Res. 47(17):6598-6606.


141<br />

Jackson KDO. A Guide to Identify Common Inorganic Fillers and Activators Using<br />

Vibrational Spectroscopy. Internet J Vib Spectosc. [Internet]. 2004.<br />

Brickendonbury, Hertford (UK): Wiley and Sons. Chapter 7. [cited 2011 Aug<br />

29]. Available from: http:www.ijvs.com/volumen2/edition3/section3.html.<br />

Karadag D, <strong>To</strong>k S, Akgul E, Ulucan K, Evden H, Kaya MA. 2006. Combining Adsorption<br />

and Coagulation for the Treatment of Azo and Anthraquinone Dyes from<br />

Aqueous Solution. Ind Eng Chem Res. 45(11):3969-3973.<br />

Kavitha SK, Palanisamy PN. 2011. Photocatalytic and Sonophotocatalytic Degradation<br />

of Reactive Red 120 using Dye Sensitized TiO 2 under Visible Light. Inter J Civil<br />

and Environ Eng. 3(1):1-6.<br />

Kislov N, Lahiri J, Verma H, Goswami DY, Stefanakos E, Batzill. 2009. Photocatalytic<br />

Degradation of Methyl Orange over Single Crystalline ZnO: Orientation<br />

Dependance of Photoactivity and Photostability of ZnO. Langmuir. 25(5):3310-<br />

3315.<br />

Leco Corp, Organic Application Note, Carbon, Hydrogen, Nitrogen and Sulfur<br />

Determination in Fuel Oils. 2010. [cited 2012 Jan 26]. Available from:<br />

http://www.leco.com/resources/application_notes/pdf/CHNS_FUEL_OILS_203-<br />

821-006.PDF<br />

Li FB, Li XZ, Ng KH. 2006. Photocatalytic Degradation of an Odorous Pollutant: 2-<br />

Mercaptobenzothiazole in Aqueous Suspension Using Nd 3+ -TiO 2 Catalysts. Ind<br />

Eng Chem Res. 45(1):1-7.<br />

Li Q, Shang JK. 2010. Composite Photocatalyst of Nitrogen and Fluorine Codoped<br />

Titanium Oxide Nanotubes Arrays with Dispersed Palladium Oxide Nanoparticles<br />

for Enhanced Visible Light Photocatalytic Performance. Environ Sci Technol.<br />

44(9):3493-3499.


142<br />

Liu L, Liu H, Zhao YP, Wang Y, Duan Y, Gao G, Ge M, Chen W. 2008. Directed<br />

Synthesis of Hierarchical Nanostructured TiO 2 Catalyst and their Morphology-<br />

Dependent Photocatalysis for Phenol Degradation. Environ Sci Technol.<br />

42(7):2342-2348.<br />

Lo SC, Lin CF, Wu CH, Hsieh PH. 2004. Capability of coupled CdSe/TiO 2 for<br />

photocatalytic degradation of 4-chlorophenol. J Haz Mat. (B114):183-190.<br />

Lodha S, Vaya D, Ameta R, Punjabi PB. 2008. Photocatalytic degradation of Phenol Red<br />

using complexes of some transition metals and hydrogen peroxide. J Serb Chem<br />

Soc. 73(6):631-639.<br />

Lopez-Fernandez V. Nanomateriales basados en carbono [Nanomaterials based in<br />

carbon] [dissertation]. [Madrid (M)]: <strong>Universidad</strong> Autónoma de Madrid; 2009.<br />

Spanish.<br />

Ma DL, Kwan MHT, Chan DSH, Lee P, Yang H, Ma VPY, Bai LP, Jiang ZH, Leung CH.<br />

2011. Crystal Violet as a Fluorescent Switch-On Probe for I-Motif: Label-Free<br />

DNA-Based Logic Gate. Royal Soc Chem. Analyst. 2011 Jul 7;136(13):2692-6.<br />

Ma H, Zhuo Q, Wang B. 2007. Chracteristic of CuO-MoO 3 -P 2 O 5 Catalyst and Its<br />

Catalytic Wet Oxidation (CWO) of Dye Wastewater under Extremely Mild<br />

Conditions. Environ Sci Technol. 41(21):7491-7496.<br />

Mahanta D, Madras G, Radhakrishnan S, Patil S. 2008. Adsorption of Sulfonated Dyes<br />

by Polyaniline Emeraldine Salt and Its Kinetic. J Phys Chem B. 112(33):10153-<br />

10157.<br />

Marbois B, Xie LX, Choi S, Hirano K, Hyman K, Clarke CF. Sep 3, 2010. para-<br />

Aminobenzoic Acid Is a Precursor in Coenzyme Q 6 Biosynthesis in<br />

Saccharomyces cerevisiae. J Biol Chem. 285(36):27827-27838.


143<br />

Marin JM, Montoya J, Monsalve E, Granda CF, Rios LA, Restrepo G. Mayo 2007.<br />

Degradación de naranja de metilo de un nuevo fotorreactor solar de placa plana<br />

con superficie corrugada [Degradation of Methyl Orange by a new solar<br />

photoreactor . Scientia Et Technica. 13(34): 435-440.<br />

Marquez F, Campo T, Cotto M, Polanco R, Roque R, Fierro P, Sanz J, Elizalde E and<br />

Morant C. 2011. Synthesis and Characterization of Monodisperse Magnetite<br />

Hollow Microspheres. Soft Nanoscience Letters. 1(2):25-32.<br />

Marquez FM, Herrera GM, Campo T, Cotto MC, Duconge J, Sanz JM, Elizalde E,<br />

Perales O, Morant C. 2012. Preparation of hollow magnetite microspheres and<br />

their applications as drugs carriers. Nanoscale Res Lett. 7(210).<br />

Masomboon N, Ratanatamskul C, Lu MC. 2009. Chemical Oxidation of 2,6-<br />

Dimethylaniline in the Fenton Process. Environ Sci Technol. 43(22):8629-8634.<br />

Mezohegyi G, Fabregat A, Font J, Bengoa C, Stuber F. 2009. Advance Bioreduction of<br />

Commercially Important Azo Dyes: Mo<strong>del</strong>ing and Correlation with<br />

Electrochemical Characteristics. Ind Eng Chem Res. 48(15):7054-7059.<br />

Mezohegyi G, Kolodkin A, Castro UI, Bengoa C, Stuber F, Font J, Fabregat A. 2007.<br />

Effective Anaerobic Decolorization of Azo Dye Acid Orange 7 in Continuos<br />

Upflow Packed-Bed Reactor Using Biological Activated Carbon System. Ind Eng<br />

Chem Res. 46(21):6788-6792.<br />

Minero C, Lucchiari M, Vione D, Maurino V. 2005. Fe(III)-Enhanced Sonochemical<br />

Degradation of Methylene Blue In Aqueous Solution. Environ Sci Technol.<br />

39(22):8936-8942.<br />

Mosteo R, Sarasa J, Ormad MP, Ovelleiro JL. 2008. Sequential Solar Photo-Fenton-<br />

Biological System for the Treatment of Winery Wastewaters. J Agric Food Chem.<br />

56(16):7333-7338.


144<br />

Neafsey K, Zeng X, Lemley AT. 2010. Degradation of Sulfonamides in Aqueous Solution<br />

by Membrane Anodic Fenton Treatment. J Agric Food Chem. 58(2):1068-1076.<br />

Niederberger M, Bartl MH, Stucky GD. 2002. Benzyl Alcohol and Titanium Tetrachloride-<br />

A Versatile Reaction System for the Nonaqueous and Low Temperature<br />

Preparation of Crystalline and Luminescent Titania Nanoparticles. Chem Mater.<br />

14(10):4364-4370.<br />

NMT: New Mexico Tech; Department Materials & Metallurgical Engineering; Material<br />

Characterization Lab. FE-SEM Principle. [Internet]. Available from:<br />

http://infohost.nmt.edu/~mtls/instruments/Fesem/FESEM%20principle.htm<br />

Nyiro-Kosa I, Csakberenyi Nagy D, Posfai M. 2009. Size and shape control of<br />

precipitated magnetite nanoparticles. Eur J Mineral. 21:293-302.<br />

OAS:Organization of American States [Internet]. 2010. Evaluación de la implementación<br />

de las iniciativas 47 a 58 <strong>del</strong> plan de acción para el desarrollo sostenibles de las<br />

Américas; Programas, leyes y políticas especificas para asegurar que el agua<br />

este libre de contaminantes. [cited 2011 Dic 22]. Available from:<br />

http://www.oas.org/dsd/publications/Unit/oea46s/ch04.htm.<br />

Osterloh FE. 2008. Inorganic Materials as Catalyst for Photochemical Splitting of Water.<br />

Reviews. Chem Mater. 20(1):35-54.<br />

Ozen AS, Aviyente V, Tezcanli-Guyer G, Ince NH. 2005. Experimental and Mo<strong>del</strong>ing<br />

Approch to Decolorization of Azo Dyes by Ultrasound: Degradation of the<br />

Hydrazone Tautomer. J Phys Chem A. 109(15):3506-3516.<br />

Panizza M, Barbucci A, Ricotti R, Cerisola G. 2006. Electrochemical degradation of<br />

methylene blue. Separation and Purification Technology. 54(3):382–387.<br />

Parshetti G, Kalme S, Saratale G, Govindwar S. 2006. Biodegradation of Malchite Green<br />

by Kocuria rosea MTCC 1532. Acta Chim Slov. 53:492-498.


145<br />

Parthasarathi V, Thilagavathi G. 2009. Synthesis and Characterization of titanium<br />

dioxide nanoparticles and their applications to textiles for microbe resistance. J<br />

Textile Apparel, Technol Manag. 6(2):1-8.<br />

Pattapu S, Saha S, Jana S, Pal A. 2008. Novel Resin Bound MnO 2 Nanocomposite for<br />

the Degradation of Crystal Violet Dye in Aqueous Medium. World J Eng.<br />

International Conference on Composite/Nano Engineering (ICCE-16). Kumming<br />

(CHN):Sun Light Publishing. [cited 2012 Apr]. Available from:<br />

http://wjoe.hebeu.edu.cn/mulu.sup.2.2010.html<br />

Pey Clemente, J. Aplicaciones de Procesos de Oxidacion Avanzada (Fotocatalisis<br />

Solar) para Tratamiento y Reutilizacion de Efluentes Textiles [Applications of the<br />

Advance Oxidation Processes (Solar Photocatalysis) for treatment and reuse of<br />

textile industry effluents] [dissertation].[Valencia (V)]: <strong>Universidad</strong> Politecnica de<br />

Valencia. 2008. Spanish<br />

Prakash Naik HR, Bhojya Naik HS, Aravinda T. 2009. Nano-Titanium dioxide (TiO 2 )<br />

mediated simple and efficient modification to Biginelli reaction. African J Pure<br />

Appl Chem. 3(9):202-207.<br />

Priya MH, Madras G. 2006. Kinetic of TiO 2 -Catalyzed Ultrasonic Degradation of<br />

Rhodamine Dyes. Ind Eng Chem Res. 45(3):913-921.<br />

Rahmani AR, Samadi MT, Enayati Moafagh A. 2008. Investigations of Photocatalytic<br />

Degradation of Phenol by UV/TiO 2 Process in Aquatic Solution. J Res Health Sci.<br />

8(2):55-60.<br />

Randorn C, Wongnawa S, Boonsin P. 2004. Bleaching of Methylene Blue by Hydrated<br />

Titanium Dioxide. ScienceAsia. 149-156.<br />

Remnev GE, Pushkarev AI, Ponomarev DV, Sazonov RV, Martinez YS. Morphology of<br />

particles of nanodispersed titanium oxide and compositional powder


146<br />

(SiO 2 ) x (TiO 2 ) 1-x . [Internet]. 2009. [cited 2012 Mar 1]; p13-16. Available from:<br />

http://www.icpig2009.unam.mx/pdf/PA13-7.pdf<br />

Ruan QJ, Zhang WD. 2009. Tunable Morphology of Bi 2 Fe 4 O 9 Crystals for Photocatalytic<br />

Oxidation. J Phys Chem C. 113(10):4168-4173.<br />

Schlumpf M, Kypke K, Vokt, CC, Birchler M, Durrer S, Faass O, Ehnes C, Fuetsch M,<br />

Gaille C, Henseler M, Hofkamp L, Maerkel K, Reolon S, Zenker A, TimmB,<br />

Tresguerres JAF, Lichtensteiger W. 2008. Endocrine Active UV Filters:<br />

Developmental <strong>To</strong>xicity and Exposure Through Breast Milk. Chimia. 62(5):345-<br />

351.<br />

Schmidt TC, Petersmann M, Kaminski L, von Löw E. Stork G. 1997. Analysis of<br />

aminobenzoic acids in waste water from a former ammunition plant with HPLC<br />

and combined diode array and fluorescence detection. Fresenius J Anal Chem.<br />

357: 121-126.<br />

Schwartz DT.. Raman Spectroscopy: Introductory Tutorial [Internet]. Seattle (WA):<br />

University of Washington. Available from:<br />

http://depts.washington.edu/ntuf/facility/docs/NTUF-Raman-Tutorial.pdf.<br />

Seymour JD, Gupta RB. 1997. Oxidation of Aqueous Pollutants Using Ultrasound:<br />

Salt:Induced Enhancement. Ind Eng Chem Res. 36(9):3453-3457.<br />

Shimizu N, Ogino C, Dadjour MF, Murata T. 2007. Sonocatalytic degradation of<br />

methylene blue with TiO 2 pellets in water. Ultrason Sonochem. 14:184-190.<br />

Sokmen M, Allen DW, Akkas F, Kartal N, Acar F. 2000. Photo-degradation of some dyes<br />

using Ag-loaded titaniumdioxide. Water,Air, and Soil Poll. 132:153-163.<br />

Sridevi D, Rajendran KV. 2009. Synthesis and optical characteristics of ZnO<br />

nanocrystals. Bull Mater Sci. 32(2):165-168.


147<br />

Su P, Chu D, Wang L. 2010. Studies on Catalytic Activity of Nanostructure Mn 2 O 3<br />

Prepared by Solvent-thermal Method on Degrading Crystal Violet.<br />

Modern<br />

<strong>Applied</strong> Sci. 4(5):125-129.<br />

Sun M, Li D, Chen Y, Chen W, Li W, He Y, Fu X. 2009. Synthesis and Photocatalytic<br />

Activity of Calcium Antimony Oxide Hydroxide for the Degradation of Dyes in<br />

Water. J Phys Chem C. 113(31):13825-13831.<br />

Takahashi M, Sendoh M, Kobayashi K, Tajima K. 2011. Preparation of TiO 2 Thin Films<br />

Using Octadecylamine Langmuir-Blodgett Films and Evaluation of Their<br />

Photocatalytic Activity. J Oleo Sci. 60(1):25-29.<br />

TA: Thermogravimetric Analysis. Available from: http://www1.chm.colostate.edu/Files/<br />

CIFDSC/TGA-Ms.pdf.<br />

Thomas JM, Thomas WJ. 1997. Principles and Practice of <strong>Heterogeneous</strong> <strong>Catalysis</strong>.<br />

Wiley-VCH [New York (NY)]. P171-175, 239, 257-259.<br />

<strong>To</strong>rres Martinez LM, Juarez Ramirez I, Garcia Montelongo XL, Cruz Lopez A. 2010.<br />

Desarrollo de semiconductores con estructuras tipo perovskitas para purificar el<br />

agua mediante oxidaciones avanzadas [Development of perosvskite structured<br />

semiconductors for wáter purification by advance oxidations]. [Internet] Ciencia<br />

UANL 23(4):376-388. Available from:<br />

http://redalyc.uaemex.mx/redalyc/pdf/402/40215505008.pdf. Spanish.<br />

UNEP, UNICEF and WHO: United Nations Environment Programme, United Nations<br />

Childrens Fund and World Health Organization (2002a). Children in the New<br />

Millenium: Environmental Threats to Children [Internet]. Available from:<br />

http://www.unep.org/ceh/main01.html.<br />

UNEP: United Nations Environment Programme. Division of Environmental Law and<br />

Conventions.Water resources; The issue of water resources. [Internet]. Available<br />

from: www.unep.org/gc/gcss-viii/Chile-e.pdf.


148<br />

UNEP: United Nations Environment Programme; Division of Environmental Law and<br />

Conventions. The Greening Water Law in Africa: Managing Freshwater<br />

Resources for People and the Environment; 2010 Nov 10-12;. Kampala,<br />

(Uganda). Available from:http://www.unep.org/DEC/support/Freshwater/Greening<br />

ofWaterLawin Africa.asp..<br />

USEPA:United State Environmental Protection Agency [Internet]. An Introduction to the<br />

Water Elements of EPA’s Strategic Plan. [date unknown] [cited 2011 Dic 31].<br />

Available from:http://water.epa.gov/aboutow/goals_objectives/goals.cfm.<br />

USGS: United State Geological Survey. <strong>To</strong>xic Substances Hydrology Program;<br />

Research Project. Emerging Contaminants in the Environment. 2011a. [Internet]<br />

Available from: http://toxics.usgs.gov/regional/emc/.<br />

USGS: United State Geological Survey. <strong>To</strong>xic Substances Hydrology Program;<br />

Research Project; Emerging Contaminants in the Environment. Potential<br />

Removal by Wastewater and Drinking Water Treatment. 2008. [Internet]<br />

Available from:http://toxics.usgs.gov/regional/emc/water_treatment.html.<br />

USGS: United State Geological Survey. USGS Multimedia Gallery. Pharmaceuticals in<br />

the Nation’s Water. 2011. [Internet] Available from:<br />

http://gallery.usgs.gov/audios/40.<br />

Vanhulle S, Trosvaslet M, Enaud E, Lucas M, Taghavi S, van der Lelie D, van Aken B,<br />

Foret M, Onderwater RCA, Wesenberg D, Agathos SN, Schneider YJ, Corbisier<br />

AM. 2008. Decolorization, Cytotoxicity, and Genotoxicity Reduction During a<br />

Combined Ozonation/Fungal Treatment of Dye-Contaminated Wastewater.<br />

Environ Sci Technol. 42(2):584-589.<br />

Velegraki T, Mantzavinos D. 2008. Conversion of benzoic acid during TiO 2 -mediated<br />

photocatalytic degradation in water. Chem Eng J. 140:15-21.


149<br />

Vinu R, Madras G. 2009. Kinetic of Sonophotocatalytic Degradation of Anionic Dyes with<br />

Nano-TiO 2 . Environ Sci Technol. 43(2):473-479.<br />

Vinu R, Madras G. 2010. Environmental remediation by photocatalysis. J Indian Inst Sci.<br />

90(21):189-230.<br />

Wang C, Yin L, Zhang L, Liu N, Lun N, Qi Y. 2010. Platinum-Nanoparticle-Modified TiO 2<br />

Nanowires with Enhanced Photocatalytic Property. Appl Mat Inter. 2(11):3373-<br />

3377.<br />

Wang XH, Li JG, Kamiyama H, Moriyoshi Y, Ishigaki T. 2006a. Wavelength-Sensitive<br />

Photocatalytic Degradation of Methyl Orange in Aqueous Suspension over Iron<br />

(III)-doped TiO 2 Nanopowders under UV and Visible Light Irradiation. J Phys<br />

Chem B. 110(13):6804-6809.<br />

Wang XK, Chen GH, Guo WL. 2003. Sonochemical Degradation Kinetic of Methyl Violet<br />

in Aqueous Solution. Molecules. 8:40-44.<br />

Wang Y, Iqbal Z, Mitra S. 2006b. Rapidly Functionalized, Water-Dispersed Carbon<br />

Nanotubes at High Concentration. J Am Chem Soc. 128(1):95-99.<br />

Wu W, Xiao X, Zhang S, Zhou J, Fan L, Ren F, Jiang C. 2010. Large-Scale and<br />

Controlled Synthesis of Iron Oxide Magnetite Short Nanotubes: Shape Evolution,<br />

Growth Mechanism, and Magnetic Properties. J Phys Chem C. 114(39):16092-<br />

16103.<br />

Xu Y, He Y, Jia J, Zhong D, Wang Y. 2009. Cu-TiO 2 /Ti Dual Rotating Disk Photocatalytic<br />

(PC) Reactor: Dual Electrode Degradation Facilitated by Spontaneous Electron<br />

Transfer. Environ Sci Technol. 43(16):6289-6294.<br />

Yang L, Luo S, Li Y, Xiao Y, Kang Q, Cai Q. 2010. High Efficient Photocatalytic<br />

Degradation of p-Nitrophenol on a Unique Cu 2 O/TiO 2 p-n Heterojunction Network<br />

Catalyst. Environ Sci Technol. 44(19):7641-7646.


150<br />

Yin M, Li Z, Kou J, Zou Z. 2009. Mechanism Investigation of Visible Light-Induced<br />

Degradation in a <strong>Heterogeneous</strong> TiO 2 /Eosin Y/Rhodamine B System. Environ Sci<br />

Technol. 43(21):8361-8366.<br />

Yu K, Yang S, He H, Sun C, Gu C, Ju Y. 2009. Visible Light-Driven Photocatalytic<br />

Degradation of Rhodamine B over NaBiO 3 : Pathways and Mechanisms. J Phys<br />

Chem A. 113(37):10024-10032.<br />

Zhang S, Chen J, Qiao X, Ge L, Cai X, Na G. 2010. Quantum Chemical Investigation<br />

and Experimental Verification on the Aquatic Photochemistry of the Sunscreen 2-<br />

Phenylbenzimidazole-5-Sulfonic Acid. Environ Sci Technol. 44(19):7484-7490.<br />

Zhou J, Zhang Y, Zhao XS, Ray AK. 2006. Photodegradation of Benzoic Acid over<br />

Metal-Doped TiO 2 . Ind Eng Chem Res. 45(10):3503-3511.<br />

Zhou W, Liu H, Wang J, Liu D, Du G, Cui J. 2010. Ag 2 O/TiO 2 Nanobelts Heterostructure<br />

with Enhanced Ultraviolet and Visible Photocatalytic Activity. Appl Mat Interf.<br />

2(8):2385-2390.


Appendix One<br />

Dyes Solutions<br />

Figure A1.01 shows the different dye solutions (10 -5 M) before being exposed to<br />

the photo-Fenton process using FeCl 2 as Fenton catalyst and the same solution after the<br />

photo-Fenton reaction.<br />

Solutions of the dyes used during the investigation. From<br />

left to right: methylene blue, methyl orange, crystal violet,<br />

rhodamine B and methyl violet, before and after the<br />

catalytic process, respectively.<br />

151


Intensity (a.u.)<br />

Absorbance<br />

C/Co<br />

Ln (C/C 0<br />

)<br />

Intensity (a.u.)<br />

Fluorescence<br />

Appendix Two<br />

Photocatalytic Process<br />

In the Appendix Two, the absorbance and fluorescence spectra, and graphics of the<br />

TOC data obtained during the photocatalytic process are shown for each pair of organic<br />

pollutant – photocatalyst.<br />

0.75<br />

0.70<br />

0.65<br />

0.60<br />

0.55<br />

0.50<br />

0.45<br />

0.40<br />

0.35<br />

0.30<br />

0.25<br />

0.20<br />

0.15<br />

0.10<br />

Absorption<br />

1: t=0<br />

2: t=10m<br />

3: t=20m<br />

4: t=30m<br />

5: t=45m<br />

6: t=60m<br />

1<br />

2<br />

3<br />

a) 1: t=0 b)<br />

4<br />

5<br />

6<br />

500<br />

450<br />

400<br />

350<br />

300<br />

250<br />

200<br />

150<br />

100<br />

Fluorescence<br />

1<br />

2<br />

3<br />

2: t=10m<br />

3: t=20m<br />

4: t=30m<br />

5: t=45m<br />

6: t=60m<br />

4<br />

5<br />

6<br />

0.05<br />

50<br />

0.00<br />

500 600 700<br />

0<br />

700<br />

Wavelenght (nm)<br />

Wavelenght (nm)<br />

0.6<br />

c) d)<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

1.0<br />

TOC: C/Co vs Irradiation time<br />

e) MB/Rutile/Photocatalysisf)<br />

0.0<br />

0.8<br />

Without H 2<br />

O 2<br />

-0.5<br />

0.6<br />

Without catalyst<br />

-1.0<br />

y = -0.0365x<br />

R² = 0.997<br />

0.4<br />

-1.5<br />

0.2<br />

-2.0<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

-2.5<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure A2.01. UV-vis absorption (a and c), fluorescence (b and d), TOC (e) and<br />

kinetic reaction rate (f) of the photocatalytic degradation process of Methylene Blue<br />

with TiO 2 (Rutile phase) as catalyst.<br />

152


C/Co<br />

Ln(C/C 0<br />

)<br />

Absorbance<br />

Fluorescence<br />

153<br />

0.6<br />

a) b)<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

1.0<br />

0.8<br />

TOC: C/Co vs Irradiation time<br />

c) d)<br />

0.0<br />

Without Catalyst<br />

-0.5<br />

MB\Anatase\Photocatalysis<br />

-1.0<br />

0.6<br />

Without H 2<br />

O 2<br />

-1.5<br />

y = -0.0424x<br />

R² = 0.9724<br />

0.4<br />

-2.0<br />

0.2<br />

-2.5<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time<br />

Figure A2.02. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the photocatalytic degradation process of Methylene Blue<br />

with TiO 2 (Anatase phase) as catalyst.


LnC/Co<br />

Ln(C\C 0<br />

)<br />

Absorbance<br />

Fluorescence<br />

154<br />

0.5<br />

a)<br />

b)<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0 10 20 30 40 50 60<br />

Wavelength (nm)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

1.0<br />

c)<br />

0.0<br />

MB\TiO 2<br />

NWs\Photocatalysis<br />

d)<br />

0.8<br />

-0.5<br />

-1.0<br />

0.6<br />

0.4<br />

Without H 2<br />

O 2<br />

Without Catalyst<br />

-1.5<br />

-2.0<br />

y = -0.0495x<br />

R² = 0.9625<br />

0.2<br />

-2.5<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

-3.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure A2.03. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the photocatalytic degradation process of Methylene Blue<br />

with TiO 2 NWs as catalyst.


C/Co<br />

Ln(C/C 0 )<br />

Absorbance<br />

Fluorescence<br />

155<br />

0.5<br />

a) b)<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

1.0<br />

TOC: C/Co vs Irradiation time<br />

c) d)<br />

0.0<br />

Without H 2<br />

O 2<br />

-0.2<br />

MB\TiO 2<br />

MWCNTs\Photocatalysis<br />

0.8<br />

-0.4<br />

-0.6<br />

0.6<br />

0.4<br />

Without Catalyst<br />

-0.8<br />

-1.0<br />

y = -0.0275x<br />

R² = 0.8004<br />

-1.2<br />

0.2<br />

-1.4<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

-1.6<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure A2.04. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the photocatalytic degradation process of Methylene Blue<br />

with TiO 2 @MWCNTs as catalyst.


C/Co<br />

Ln(C/C 0<br />

)<br />

Absorbance<br />

Fluorescence<br />

156<br />

0.6<br />

0.5<br />

a)<br />

b)<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0 10 20 30 40 50 60<br />

Wavelength (nm)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

1.0<br />

TOC: C/Co vs Irradiation time<br />

c) d)<br />

0.0<br />

-0.2<br />

MB\ZnO\Photocatalyst<br />

0.8<br />

-0.4<br />

-0.6<br />

0.6<br />

0.4<br />

Without H 2<br />

O 2<br />

Without Catalyst<br />

-0.8<br />

-1.0<br />

-1.2<br />

y = -0.0319x<br />

R² = 0.8872<br />

0.2<br />

-1.4<br />

-1.6<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

-1.8<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure A2.05. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the photocatalytic degradation process of Methylene<br />

Blue with ZnO as catalyst.


157<br />

Figure A2.06. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the photocatalytic degradation process of Rhodamine B<br />

with TiO 2 (Rutile phase) as catalyst.


158<br />

Figure A2.07. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the photocatalytic degradation process of Rhodamine B<br />

with TiO 2 (Anatase phase) as catalyst.


159<br />

Figure A2.08.<br />

UV-vis absorption (a), fluorescence (b), TOC (c) and<br />

kinetic reaction rate (d) of the photocatalytic degradation process of<br />

Rhodamine B with TiO 2 NWs as catalyst.


160<br />

Figure A2.09.<br />

UV-vis absorption (a), fluorescence (b), TOC (c) and<br />

kinetic reaction rate (d) of the photocatalytic degradation process of<br />

Rhodamine B with TiO 2 @MWCNTs as catalyst.


161<br />

Figure A2.10. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the photocatalytic degradation process of Rhodamine B<br />

with ZnO as catalyst.


162<br />

Figure A2.11. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the photocatalytic degradation process of Methyl Orange<br />

with TiO 2 (Rutile phase) as catalyst.


163<br />

Figure A2.12.<br />

UV-vis absorption (a), fluorescence (b), TOC (c) and<br />

kinetic reaction rate (d) of the photocatalytic degradation process of<br />

Methyl Orange with TiO 2 (Anatase phase) as catalyst.


164<br />

Figure A2.13. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the photocatalytic degradation process of Methyl Orange<br />

with TiO 2 NWs as catalyst.


165<br />

Figure A2.14. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the photocatalytic degradation process of Methyl<br />

Orange with TiO 2 MWCNTs as catalyst.


166<br />

Figure A2.15. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic reaction<br />

rate (d) of the photocatalytic degradation process of Methyl Orange with ZnO as<br />

catalyst.


167<br />

Figure A2.16.<br />

UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the photocatalytic degradation process of Crystal Violet with<br />

TiO 2 (Rutile phase) as catalyst.


C/Co<br />

Ln(C/C 0<br />

)<br />

Absorbance<br />

Fluorescence<br />

168<br />

0.8<br />

0.7<br />

a)<br />

b)<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

1.0<br />

TOC: C/Co vs Irradiation time<br />

c) CV\Anatase\Photocatalysis d)<br />

0.0<br />

0.8<br />

-0.5<br />

0.6<br />

Without Catalyst<br />

Without H 2<br />

O 2<br />

-1.0<br />

y = -0.0355x<br />

R² = 0.9926<br />

0.4<br />

-1.5<br />

0.2<br />

-2.0<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

-2.5<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure A2.17. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic reaction<br />

rate (d) of the photocatalytic degradation process of Crystal Violet with TiO 2 (Anatase<br />

phase) as catalyst.


169<br />

Figure A2.18. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic reaction<br />

rate (d) of the photocatalytic degradation process of Crystal Violet with TiO 2 NWs as<br />

catalyst.


170<br />

Figure A2.19. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic reaction<br />

rate (d) of the photocatalytic degradation process of Crystal Violet with<br />

TiO 2 @MWCNTs as catalyst.


171<br />

Figure A2.20. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic reaction<br />

rate (d) of the photocatalytic degradation process of Crystal Violet with ZnO as<br />

catalyst.


172<br />

Figure A2.21.<br />

UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the photocatalytic degradation process of Methyl Violet with<br />

TiO 2 (Rutile phase) as catalyst.


173<br />

Figure A2.22. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic reaction<br />

rate (d) of the photocatalytic degradation process of Methyl Violet with TiO 2<br />

(Anatase phase) as catalyst.


174<br />

Figure A2.23.<br />

UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the photocatalytic degradation process of Methyl Violet with<br />

TiO 2 NWs as catalyst.


175<br />

Figure A2.24. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic reaction<br />

rate (d) of the photocatalytic degradation process of Methyl Violet with<br />

TiO 2 MWCNTs as catalyst.


176<br />

Figure A2.25. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic reaction<br />

rate (d) of the photocatalytic degradation process of Methyl Violet with ZnO as<br />

catalyst.


177<br />

Figure A2.26.<br />

Curves of the TOC (a) and kinetic<br />

reaction rate (b) of the photocatalytic degradation<br />

process of the p-ABA using TiO 2 (Rutile phase) as<br />

catalyst.


178<br />

Figure A2.27.<br />

Curves of the TOC (a) and kinetic<br />

reaction rate (b) of the photocatalytic degradation<br />

process of the p-ABA using TiO 2 (Anatase phase) as<br />

catalyst as catalyst.


179<br />

Figure A2.28. Curves of the TOC (a) and kinetic<br />

reaction rate (b) of the photocatalytic degradation<br />

process of the p-ABA using TiO 2 NWs as catalyst.


180<br />

Figure A2.29. Curves of the TOC (a) and kinetic<br />

reaction rate (b) of the photocatalytic degradation<br />

process of the p-ABA using TiO 2 @MWCNTs as<br />

catalyst.


181<br />

Figure A2.30. Curves of the TOC (a) and kinetic<br />

reaction rate (b) of the photocatalytic degradation<br />

process of the p-ABA using ZnO as catalyst.


C/Co<br />

Ln(C/C 0<br />

)<br />

Absorbance<br />

Fluorescence<br />

Appendix Three<br />

Sono-Fenton Process<br />

In the Appendix Three, the absorbance and fluorescence spectra, and graphics of the<br />

TOC data obtained during the sono-Fenton process are shown for each pair of organic<br />

pollutant – photocatalyst.<br />

0.5<br />

a)<br />

b)<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

1.0<br />

TOC: C/Co vs Irradiation time<br />

Without Catalyst<br />

c)<br />

0.0 MB\CuO\Sono-Fenton d)<br />

-0.2<br />

0.8<br />

-0.4<br />

0.6<br />

0.4<br />

Without H 2<br />

O 2<br />

-0.6<br />

-0.8<br />

y = -0.022x<br />

R² = 0.8621<br />

-1.0<br />

0.2<br />

-1.2<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

-1.4<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure A3.01. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic rate<br />

reaction (d) of the Sono-Fenton degradation process of Methylene blue with CuO<br />

as catalyst.<br />

182


C/Co<br />

Ln(C/C 0<br />

)<br />

Absorbance<br />

Fluorescence<br />

183<br />

0.5<br />

a) b)<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0 10 20 30 40 50 60<br />

Wavelength (nm)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

1.0<br />

Without Catalyst<br />

MB\Fe 2 O 3 NWs\Sono-Fenton<br />

c) d)<br />

0.0<br />

0.8<br />

-0.5<br />

0.6<br />

0.4<br />

Without H 2<br />

O 2<br />

-1.0<br />

-1.5<br />

y = -0.0335x<br />

R² = 0.965<br />

0.2<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

-2.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure A3.02. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic reaction<br />

rate (d) of the Sono-Fenton degradation process of Methylene blue with Fe 2 O 3 NWs<br />

as catalyst.


C/Co<br />

Ln(C/C 0<br />

)<br />

Absorbance<br />

Fluorescence<br />

184<br />

0.5<br />

a)<br />

b)<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

1.0<br />

TOC: C/Co<br />

Without Catalyst<br />

c)<br />

MB\Fe<br />

0.0<br />

3<br />

O 4<br />

Magnetite\Sono-Fenton d)<br />

-0.2<br />

-0.4<br />

0.8<br />

-0.6<br />

0.6<br />

Without H 2<br />

O 2<br />

-0.8<br />

-1.0<br />

y = -0.0293x<br />

R² = 0.9812<br />

0.4<br />

-1.2<br />

-1.4<br />

0.2<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

-1.6<br />

-1.8<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure A3.03.<br />

UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the Sono-Fenton degradation process of Methylene blue with<br />

Fe 3 O 4 (Magnetite) as catalyst.


C/Co<br />

Ln(C/C 0<br />

)<br />

Absorbance<br />

Fluorescence<br />

185<br />

0.5<br />

a)<br />

b)<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

TOC: C/Co<br />

Without Catalyst<br />

c) d)<br />

0.0<br />

MB\FeCl 2<br />

\Sono-Fenton<br />

1.0<br />

0.8<br />

-0.5<br />

-1.0<br />

0.6<br />

Without H 2<br />

O 2<br />

-1.5<br />

y = -0.0525x<br />

R² = 0.9455<br />

0.4<br />

-2.0<br />

0.2<br />

-2.5<br />

-3.0<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure A3.04.<br />

UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the Sono-Fenton degradation process of Methylene blue with<br />

FeCl 2 as catalyst.


186<br />

Figure A3.05. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the Sono-Fenton degradation process of Rhodamine B with<br />

CuO as catalyst.


187<br />

Figure A3.06. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the Sono-Fenton degradation process of Rhodamine B with<br />

Fe 2 O 3 NWs as catalyst.


188<br />

Figure A3.07. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the Sono-Fenton degradation process of Rhodamine B with<br />

Fe 3 O 4 (Magnetite) as catalyst.


189<br />

Figure A3.08. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the Sono-Fenton degradation process of Rhodamine B with<br />

FeCl 2 as catalyst.


190<br />

Figure A3.09. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the Sono-Fenton degradation process of Methyl Orange<br />

with CuO as catalyst.


191<br />

Figure A3.10.<br />

UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the Sono-Fenton degradation process of Methyl Orange with<br />

Fe 2 O 3 NWs as catalyst.


192<br />

Figure A3.11. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic reaction<br />

rate (d) of the Sono-Fenton degradation process of Methyl Orange with Fe 3 O 4<br />

(Magnetite) as catalyst.


193<br />

Figure A3.12. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the Sono-Fenton degradation process of Methyl Orange<br />

with FeCl 2 as catalyst.


C/Co<br />

Ln(C/C 0<br />

)<br />

Absorbance<br />

Fluorescence<br />

194<br />

0.9<br />

0.8<br />

a)<br />

b)<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

1.0<br />

TOC: C/Co<br />

c) 0.0<br />

CV\CuO\Sono-Fenton<br />

d)<br />

-0.1<br />

0.8<br />

-0.2<br />

0.6<br />

Without Catalyst<br />

Without H 2<br />

O 2<br />

-0.3<br />

-0.4<br />

y = -0.013x<br />

R² = 0.7797<br />

0.4<br />

-0.5<br />

-0.6<br />

0.2<br />

-0.7<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

-0.8<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure A3.13. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the Sono-Fenton degradation process of Crystal Violet with<br />

CuO as catalyst.


C/Co<br />

Ln(C/C 0<br />

)<br />

Absorbance<br />

Fluorescence<br />

195<br />

0.9<br />

0.8<br />

a)<br />

b)<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

TOC: C/Co<br />

c)<br />

0.0<br />

CV\Fe 2<br />

O 3<br />

NWs\Sono-Fenton d)<br />

1.0<br />

-0.2<br />

0.8<br />

-0.4<br />

Without H 2<br />

O 2<br />

0.6<br />

Without Catalyst<br />

-0.6<br />

y = -0.0203x<br />

R² = 0.8348<br />

0.4<br />

-0.8<br />

0.2<br />

-1.0<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

-1.2<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure A3.14. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic reaction<br />

rate (d) of the Sono-Fenton degradation process of Crystal Violet with Fe 2 O 3 NWS<br />

as catalyst.


C/Co<br />

Ln(C/C 0<br />

)<br />

Absorbance<br />

Fluorescence<br />

196<br />

0.9<br />

0.8<br />

a)<br />

b)<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

TOC: C/Co<br />

c) d)<br />

0.0<br />

CV\Fe 3<br />

O 4<br />

Magnetite\Sono-Fenton<br />

1.0<br />

-0.2<br />

0.8<br />

0.6<br />

0.4<br />

Without Catalyst<br />

Without H 2<br />

O 2<br />

-0.4<br />

-0.6<br />

y = -0.0171x<br />

R² = 0.8172<br />

0.2<br />

-0.8<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

-1.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure A3.15. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the Sono-Fenton degradation process of Crystal Violet with<br />

Fe 3 O 4 (Magnetite) as catalyst.


C/Co<br />

Ln(C/C 0<br />

)<br />

Absorbance<br />

Fluorescence<br />

197<br />

0.9<br />

0.8<br />

a)<br />

b)<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

1.0<br />

TOC: C/Co<br />

c)<br />

0.0<br />

d)<br />

CV\FeCl 2<br />

\Sono-Fenton<br />

-0.2<br />

0.8<br />

-0.4<br />

0.6<br />

0.4<br />

Without Catalyst<br />

Without H 2<br />

O 2<br />

-0.6<br />

-0.8<br />

-1.0<br />

y = -0.0253x<br />

R² = 0.8399<br />

0.2<br />

-1.2<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

-1.4<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure A3.16. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic reaction<br />

rate (d) of the Sono-Fenton degradation process of Crystal Violet with FeCl 2 as<br />

catalyst.


C/Co<br />

Ln(C/C 0<br />

)<br />

Absorbance<br />

Fluorescence<br />

198<br />

1.0<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

a)<br />

b)<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

1.0<br />

TOC: C/Co<br />

Without H 2<br />

O 2<br />

c) MV\CuO\Sono-Fenton<br />

d)<br />

0.0<br />

-0.2<br />

0.8<br />

0.6<br />

Without catalyst<br />

-0.4<br />

-0.6<br />

y = -0.0189x<br />

R² = 0.8764<br />

0.4<br />

-0.8<br />

0.2<br />

-1.0<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

-1.2<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure A3.17.<br />

UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the Sono-Fenton degradation process of Methyl Violet with<br />

CuO as catalyst.


C/Co<br />

Ln(C/C 0<br />

)<br />

Absorbance<br />

Fluorescence<br />

199<br />

1.0<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

a)<br />

b)<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

TOC: C/Co<br />

Without H 2 O 2<br />

c)<br />

0.0<br />

MV\Fe 2<br />

O 3<br />

NWs\Sono-Fenton<br />

d)<br />

1.0<br />

0.8<br />

-0.5<br />

0.6<br />

Without Catalyst<br />

-1.0<br />

y = -0.0321x<br />

R² = 0.9771<br />

0.4<br />

-1.5<br />

0.2<br />

-2.0<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure A3.18. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic reaction<br />

rate (d) of the Sono-Fenton degradation process of Methyl Violet with Fe 2 O 3 NWS<br />

as catalyst.


C/Co<br />

Ln(C/C 0<br />

)<br />

Absorbance<br />

Fluorescence<br />

200<br />

1.0<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

a)<br />

b)<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

TOC: C/Co<br />

c)<br />

MV\Fe<br />

0.0<br />

3<br />

O 4<br />

Magnetite\Sono-Fenton<br />

d)<br />

1.0<br />

Without H 2<br />

O 2<br />

-0.2<br />

0.8<br />

-0.4<br />

0.6<br />

Without Catalyst<br />

-0.6<br />

-0.8<br />

y = -0.0246x<br />

R² = 0.9445<br />

0.4<br />

-1.0<br />

0.2<br />

-1.2<br />

-1.4<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure A3.19. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic reaction<br />

rate (d) of the Sono-Fenton degradation process of Methyl Violet with Fe 3 O 4<br />

(Magnetite) as catalyst.


C/Co<br />

Ln(C/C 0<br />

)<br />

Fluorescence<br />

Absorbance<br />

201<br />

a)<br />

1.0<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

b)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

1.0<br />

TOC: C/Co<br />

c) MV\FeCl 2<br />

\Sono-Fenton d)<br />

0.0<br />

0.8<br />

-0.5<br />

-1.0<br />

0.6<br />

0.4<br />

Without Catalyst<br />

Without H 2<br />

O 2<br />

-1.5<br />

-2.0<br />

y = -0.0505x<br />

R² = 0.988<br />

0.2<br />

-2.5<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

-3.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure A3.20. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic reaction<br />

rate (d) of the Sono-Fenton degradation process of Methyl Violet with FeCl 2 as<br />

catalyst.


202<br />

Figure A3.21. Curves of the TOC (a) and kinetic<br />

reaction rate (b) of the Sono-Fenton degradation<br />

process of the p-ABA with CuO as catalyst.


203<br />

Figure A3.22. Curves of the TOC (a) and kinetic<br />

reaction rate (b) of the Sono-Fenton degradation<br />

process of the p-ABA with Fe 2 O 3 NWs as catalyst.


204<br />

Figure A3.23. Curves of the TOC (a) and kinetic<br />

reaction rate (b) of the Sono-Fenton degradation<br />

process of the p-ABA with Fe 3 O 4 (Magnetite) as<br />

catalyst.


205<br />

Figure A3.24. Curves of the TOC (a) and kinetic<br />

reaction rate (b) of the Sono-Fenton degradation<br />

process of the p-ABA with FeCl 2 as catalyst.


C/Co<br />

Ln(C/C 0<br />

)<br />

Absorbance<br />

Fluorescence<br />

Appendix Four<br />

Photo-Fenton Process<br />

In the Appendix Four, the absorbance and fluorescence spectra, and graphics of the<br />

TOC data obtained during the photo-Fenton process are shown for each pair of organic<br />

pollutant – photocatalyst.<br />

0.5<br />

a)<br />

b)<br />

0.4<br />

0.3<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

1.0<br />

TOC: C/Co vs Irradiation time<br />

Without Catalyst<br />

c) MB\CuO\Photo-Fenton d)<br />

0.0<br />

0.8<br />

-0.2<br />

0.6<br />

0.4<br />

Without H 2<br />

O 2<br />

-0.4<br />

-0.6<br />

y = -0.0166x<br />

R² = 0.7264<br />

0.2<br />

-0.8<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

-1.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure A4.01.<br />

UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the Photo-Fenton degradation process of Methylene Blue with<br />

CuO as catalyst.<br />

206


C/Co<br />

Ln(C/C 0<br />

)<br />

Absorbance<br />

Fluorescence<br />

207<br />

0.5<br />

a)<br />

b)<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0 10 20 30 40 50 60<br />

Wavelength (nm)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

1.0<br />

TOC: C/Co vs Irradiation time<br />

Without Catalyst<br />

c)<br />

0.0<br />

MB\Fe 2<br />

O 3<br />

NWs\Photo-Fenton<br />

d)<br />

0.8<br />

-0.5<br />

0.6<br />

0.4<br />

Without H 2<br />

O 2<br />

-1.0<br />

-1.5<br />

y = -0.0365x<br />

R² = 0.9582<br />

0.2<br />

-2.0<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure A4.02.<br />

UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the Photo-Fenton degradation process of Methylene Blue with<br />

Fe 2 O 3 NWS as catalyst.


C/Co<br />

Ln(C/C 0<br />

)<br />

Absorbance<br />

Fluorescence<br />

208<br />

0.5<br />

a)<br />

b)<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

1.0<br />

TOC: C/Co<br />

Without Catalyst<br />

c)<br />

MB\Fe 3<br />

O 4<br />

Magnetite\Photo-Fenton<br />

0.0<br />

d)<br />

-0.2<br />

0.8<br />

-0.4<br />

0.6<br />

Without H 2<br />

O 2<br />

-0.6<br />

-0.8<br />

y = -0.0237x<br />

R² = 0.915<br />

0.4<br />

-1.0<br />

0.2<br />

-1.2<br />

-1.4<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

-1.6<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure A4.03.<br />

UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the Photo-Fenton degradation process of Methylene Blue with<br />

Fe 3 O 4 (Magnetite) as catalyst.


C/Co<br />

Ln(C/C 0<br />

)<br />

Absorbance<br />

Fluorescence<br />

209<br />

0.5<br />

a)<br />

b)<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

1.0<br />

TOC: C/Co<br />

Without Catalyst<br />

c)<br />

MB\FeCl 2<br />

\Photo-Fenton<br />

d)<br />

0.0<br />

-0.5<br />

0.8<br />

-1.0<br />

0.6<br />

Without H 2<br />

O 2<br />

-1.5<br />

y = -0.0467x<br />

R² = 0.9241<br />

0.4<br />

-2.0<br />

0.2<br />

-2.5<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

-3.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure A4.04.<br />

UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the Photo-Fenton degradation process of Methylene Blue with<br />

FeCl 2 as catalyst.


210<br />

Figure A4.05. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the Photo-Fenton degradation process of Rhodamine B with<br />

CuO as catalyst.


211<br />

Figure A4.06.<br />

UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the Photo-Fenton degradation process of Rhodamine B with<br />

Fe 2 O 3 NWs as catalyst.


212<br />

Figure A4.07.<br />

UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the Photo-Fenton degradation process of Rhodamine B with<br />

Fe 3 O 4 (Magnetite) as catalyst.


213<br />

Figure A4.08. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the Photo-Fenton degradation process of Rhodamine B with<br />

FeCl 2 as catalyst.


214<br />

Figure A4.09.<br />

UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the Photo-Fenton degradation process of Methyl Orange with<br />

CuO as catalyst.


215<br />

Figure A4.10.<br />

UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the Photo-Fenton degradation process of Methyl Orange with<br />

Fe 2 O 3 NWs as catalyst.


216<br />

Figure A4.11.<br />

UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the Photo-Fenton degradation process of Methyl Orange with<br />

Fe 3 O 4 (Magnetite) as catalyst.


217<br />

Figure A4.12.<br />

UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the Photo-Fenton degradation process of Methyl Orange with<br />

FeCl 2 as catalyst.


C/Co<br />

Ln(C/C 0<br />

)<br />

Absorbance<br />

Fluorescence<br />

218<br />

0.9<br />

0.8<br />

a)<br />

b)<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

1.0<br />

0.8<br />

TOC: C/Co<br />

c)<br />

Without H 2<br />

O 2 0.0<br />

Without Catalyst<br />

CV\CuO\Photo-Fenton<br />

d)<br />

-0.1<br />

-0.2<br />

0.6<br />

-0.3<br />

-0.4<br />

y = -0.0116x<br />

R² = 0.9655<br />

0.4<br />

-0.5<br />

0.2<br />

-0.6<br />

-0.7<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

-0.8<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure A4.13. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic reaction<br />

rate (d) of the Photo-Fenton degradation process of Crystal Violet with CuO as<br />

catalyst.


C/Co<br />

Ln(C/C 0<br />

)<br />

Absorbance<br />

Fluorescence<br />

219<br />

0.9<br />

0.8<br />

a)<br />

b)<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

TOC: C/Co<br />

Without H 2<br />

O 2<br />

c) CV\Fe 2<br />

O 3<br />

NWs\Photo-Fenton d)<br />

0.0<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

Without Catalyst<br />

-0.2<br />

-0.4<br />

-0.6<br />

y = -0.0169x<br />

R² = 0.7964<br />

0.2<br />

-0.8<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

-1.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure A4.14.<br />

UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the Photo-Fenton degradation process of Crystal Violet with<br />

Fe 2 O 3 NWs as catalyst.


C/Co<br />

Ln(C/C 0<br />

)<br />

Absorbance<br />

Fluorescence<br />

220<br />

0.9<br />

0.8<br />

a)<br />

b)<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

1.0<br />

TOC: C/Co<br />

Without H 2<br />

O 2<br />

c)<br />

0.0<br />

CV\Fe 3<br />

O 4<br />

Magnetite\Photo-Fenton d)<br />

Without Catalyst<br />

-0.1<br />

0.8<br />

-0.2<br />

-0.3<br />

0.6<br />

0.4<br />

0.2<br />

-0.4<br />

-0.5<br />

-0.6<br />

-0.7<br />

y = -0.0134x<br />

R² = 0.8548<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

-0.8<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure A4.15.<br />

UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the Photo-Fenton degradation process of Crystal Violet with<br />

Fe 3 O 4 (Magnetite) as catalyst.


C/Co<br />

Ln(C/C 0<br />

)<br />

Absorbance<br />

Fluorescence<br />

221<br />

0.9<br />

0.8<br />

a)<br />

b)<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

TOC: C/Co<br />

c)<br />

0.0<br />

CV\FeCl 2<br />

\Photo-Fenton<br />

d)<br />

Without H 2<br />

O 2<br />

Without Catalyst<br />

1.0<br />

-0.2<br />

-0.4<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

-0.6<br />

-0.8<br />

-1.0<br />

-1.2<br />

-1.4<br />

-1.6<br />

y = -0.0309x<br />

R² = 0.8473<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

-1.8<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure A4.16.<br />

UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the Photo-Fenton degradation process of Crystal Violet with<br />

FeCl 2 as catalyst.


C/Co<br />

Ln(C/C 0<br />

)<br />

Absorbance<br />

Fluorescence<br />

222<br />

1.0<br />

0.9<br />

a)<br />

b)<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

1.0<br />

TOC: C/Co Without H 2<br />

O 2<br />

c)<br />

Without Catalyst<br />

MV\CuO\Photo-Fenton d)<br />

0.0<br />

0.8<br />

-0.2<br />

0.6<br />

-0.4<br />

y = -0.0146x<br />

R² = 0.9181<br />

0.4<br />

-0.6<br />

0.2<br />

-0.8<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure A4.17. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic reaction<br />

rate (d) of the Photo-Fenton degradation process of Methyl Violet with CuO as<br />

catalyst.


C/Co<br />

Ln(C/C 0<br />

)<br />

Absorbance<br />

Fluorescence<br />

223<br />

0.9<br />

0.8<br />

a)<br />

b)<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

TOC: C/Co<br />

0.2<br />

c) d)<br />

MV\Fe 2<br />

O 3<br />

NWs\Photo-Fenton<br />

0.0<br />

1.0<br />

Without H 2<br />

O 2<br />

-0.2<br />

0.8<br />

-0.4<br />

0.6<br />

0.4<br />

Without Catalyst<br />

-0.6<br />

-0.8<br />

-1.0<br />

-1.2<br />

y = -0.0264x<br />

R² = 0.9712<br />

0.2<br />

-1.4<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

-1.6<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure A4.18.<br />

UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the Photo-Fenton degradation process of Methyl Violet with<br />

Fe 2 O 3 NWs as catalyst.


C/Co<br />

Ln(C/C 0<br />

)<br />

Absorbance<br />

Fluorescence<br />

224<br />

1.0<br />

0.9<br />

a)<br />

b)<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

TOC: C/Co<br />

c)<br />

MV\Fe 3<br />

O 4<br />

Magnetite\Photo-Fenton<br />

d)<br />

1.0<br />

Without H 2<br />

O 2<br />

0.0<br />

-0.2<br />

0.8<br />

0.6<br />

Without Catalyst<br />

-0.4<br />

-0.6<br />

y = -0.0186x<br />

R² = 0.8993<br />

0.4<br />

-0.8<br />

0.2<br />

-1.0<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

-1.2<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure A4.19. UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic reaction<br />

rate (d) of the Photo-Fenton degradation process of Methyl Violet with Fe 3 O 4<br />

(Magnetite) as catalyst.


C/Co<br />

Ln(C/C 0<br />

)<br />

Absorbance<br />

Fluorescence<br />

225<br />

1.0<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

a)<br />

b)<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

1.0<br />

TOC: C/Co<br />

c)<br />

MV\FeCl 2<br />

\Photo-Fenton<br />

d)<br />

0.0<br />

Without H 2<br />

O 2<br />

Without Catalyst<br />

0.8<br />

-0.5<br />

0.6<br />

0.4<br />

-1.0<br />

-1.5<br />

y = -0.0396x<br />

R² = 0.9854<br />

0.2<br />

-2.0<br />

0.0<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

-2.5<br />

0 10 20 30 40 50 60<br />

Time (min)<br />

Figure A4.20.<br />

UV-vis absorption (a), fluorescence (b), TOC (c) and kinetic<br />

reaction rate (d) of the Photo-Fenton degradation process of Methyl Violet with<br />

FeCl 2 as catalyst.


226<br />

Figure A4.21. Curves of the TOC (a) and kinetic<br />

reaction rate (b) of the Photo-Fenton degradation<br />

process of the p-ABA with CuO as catalyst.


227<br />

Figure A4.22. Curves of the TOC (a) and kinetic<br />

reaction rate (b) of the Photo-Fenton degradation<br />

process of the p-ABA with Fe 2 O 3 NWs as catalyst.


228<br />

Figure A4.23. Curves of the TOC (a) and kinetic<br />

reaction rate (b) of the Photo-Fenton degradation<br />

process of the p-ABA with Fe 3 O 4 (Magnetite) as<br />

catalyst.


229<br />

Figure A4.24.<br />

Curves of the TOC (a) and kinetic<br />

reaction rate (b) of the Photo-Fenton degradation<br />

process of the p-ABA with FeCl 2 as catalyst.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!