MODIFIED MANGANESE OXIDE OCTAHEDRAL MOLECULAR SIEVES FOR OXIDATION AND CONSECUTIVE OXIDATION-ACID REACTIONS FITRI HAYATI DARMALIS UNIVERSITI TEKNOLOGI MALAYSIA MODIFIED MANGANESE OXIDE OCTAHEDRAL MOLECULAR SIEVES FOR OXIDATION AND CONSECUTIVE OXIDATION-ACID REACTIONS FITRI HAYATI DARMALIS A thesis submitted in fulfilment of the requirements for the award of the degree of Doctor of Philosophy (Chemistry) Faculty of Science Universiti Teknologi Malaysia DECEMBER 2009 iii To my husband, Dr. Hendriyawan To my mother, Raidas To my father, Darmalis To my sons: Faizan Munawwar Alfindri and Muhammad Husein Murtaza For my Brothers: Yon Elfi, Joni Indra, Yalmasri and Khairul Arif iv ACKNOWLEDGEMENTS Bismillahirrahmanirrahim In the name of Allah, the Most Gracious, the Most Merciful All praise to be Allah, the supreme Lord of the world. May peace and blessings to Rasulullah Muhammad Shollallahu’ Alaihi Wassalam, all the prophets, his families, his close friends and all Muslims. Firstly, I wish to express my deep sincere appreciation to my supervisors Prof Dr. Halimaton Hamdan and Assoc. Prof. Dr. Hadi Nur, for their gratitude, encouragement, friendly advice, earnest guidance, and motivation. Without continuous support and curiosity from my supervisors, this thesis would not have been the same as presented here. I am also indebted to UTM and research grant from the Ministry of Science Technology and Innovation Malaysia (MOSTI) for funding my Ph.D. study under VOT number 74506. This support is gratefully acknowledged. I also would like to express my gratitude to all lecturers and researchers in Department of Chemistry for their support and Ibnu Sina Institute for Fundamental Science Studies (IIS) for catalysts characterization. Special gratitude is addressed to Dr. Lee and my colleagues at Zeolite and Porous Materials Group (ZPMG), namely Mr. Izan, Mrs. Rozana, Mrs. Suryani and Miss. Hidayah and others for useful discussion and help. My sincere appreciation also extends to all my friends who have provided supports at various occasions especially Mrs. Rosyida Permatasari and Mrs. Suryani Alifah. Last but not least, my gratitude goes to my caring family, my husband, my mother and father, my mother and father (Allah Yarham) -in law, my son and my brothers. To Bapak Nur Anas Djamil, Ibu Sofiah Djamaris and Dr Hadi’s family. I thank you for continuous support, prayer, love, understanding and encouragement. v ABSTRACT The research was focused on improvement of the catalytic activity of octahedral manganese oxide molecular sieve (OMS-2) in oxidation and consecutive oxidation-acid reactions. For oxidation reaction, OMS-2 was modified through incorporation of metals and ion-exchange. Sulphation was applied in order to create the acid sites on Ti-OMS-2 sample which was proven to have the highest oxidative properties and tested for consecutive oxidation-acidic reaction. Ti, Fe, Co and Cu were incorporated into the OMS-2 framework by isomorphous substitution to form Ti-OMS-2, Fe-OMS-2, Co-OMS-2 and Cu-OMS-2, respectively. A new method to synthesize Ti-OMS-2 with high Ti/Mn ratio was applied. Titanium incorporated OMS-2 was successfully synthesized without addition of manganese (II) solution which was normally necessary to synthesize metal substituted OMS-2. Ion-exchange was carried out in order to replace K+ ion in the tunnel structure of OMS-2 framework by H+ ions using concentrated HNO3. Sulphation was done by impregnation of certain amount of H2SO4 in different solvent. The characterization results show that Ti-OMS-2 exhibited a significantly higher Lewis acidity compared to the un-incorporated one. The physicochemical properties-catalytic activity of the modified OMS-2 catalyst was studied in the oxidation of cyclohexane, cylohexene and styrene, and also in consecutive transformation of 1-octene to 1,2-octanediol. Oxidation of cyclohexane with TBHP as oxidant on transition metal substituted OMS-2 showed that Ti-OMS-2 with high titanium content gave the highest conversion, which may be due to the presence Ti sites in the framework and nonframework. For metal incorporated in the framework, there is a correlation of ionic radii of metal substituted with conversion of cyclohexane. An increase in ionic radii of metal substituted OMS-2 increased the conversion of cyclohexane. This correlation may be due to the increase in the Lewis acidity in the metal incorporated OMS-2. The study on the catalytic activity of H-exchanged catalysts in oxidation of cyclohexane showed an increase in conversion of cyclohexane after ion-exchange. For further investigation, Ti-OMS-2 with high Ti/Mn ratio was used in oxidation of cyclohexene and styrene using TBHP as oxidant. The results showed that both titanium sites in framework and non-framework increased the activity of OMS-2 in the oxidation of cyclohexene. However, it was observed that only non-framework titanium species induced a synergetic effect that enhanced the oxidation of styrene. There is a correlation between Ti site location in Ti-OMS-2 catalyst with activation of C-H and C=C bonds. Ti sites in the framework only played role in C-H bond activation whereas Ti site non-framework enhanced the catalytic activity for both types of bond activation. SO42-/Ti-OMS-2 was proven to be active for consecutive transformation of 1-octene to 1,2-octanediol. However, it was confirmed that Brönsted acid sites did not exist in the sample. The success of the consecutive reactions may be due to the generation of Brönsted acid from hydrolysis of water on the Lewis acid sites of SO42-/Ti-OMS-2 sample. vi ABSTRAK Penyelidikan difokuskan pada peningkatan aktiviti pemangkinan daripada penapis molekul oksida mangan oktahedral (OMS-2) sebagai mangkin dalam tindak balas pengoksidaan dan juga tindak balas berturutan pengoksidaan dan keasidan. Untuk tindak balas pengoksidaan, OMS-2 diubah suai melalui pemasukan logam dan pertukaran ion. Pensulfatan telah dilakukan untuk membina tapak asid pada mangkin Ti-OMS-2 yang terbukti mempunyai aktiviti pemangkinan paling tinggi dalam tindak balas pengoksidaan dan seterusnya diuji pada tindakbalas berturutan pengoksidaan dan keasidan. Ti, Fe, Co dan Cu dimasukkan ke dalam bingkaian OMS-2 melalui penukargantian isomorfus untuk membentuk Ti-OMS-2, Fe-OMS-2, Co-OMS-2 dan Cu-OMS-2. Satu kaedah baru telah digunakan untuk mensintesis Ti-OMS-2 pada nisbah Ti:Mn yang tinggi. Titanium yang digabungkan dengan OMS-2 telah berjaya disintesis tanpa penambahan larutan mangan(II) yang biasanya diperlukan untuk mensintesis penukargantian logam pada OMS-2 seperti yang dilaporkan sebelum ini. Penukaran ion telah dilakukan untuk menggantikan ion K+ didalam struktur terowong dengan ion H+ menggunakan HNO3 pekat. Pensulfatan telah dilakukan dengan pengisitepuan jumlah tertentu H2SO4 dalam pelarut yang berbeza. Hasil kaedah pencirian didapati bahawa penggabungan titanium kepada OMS-2 telah meningkatkan keasidan Lewis jika dibandingkan dengan OMS-2 sahaja. Pengoksidaan terhadap sikloheksana dengan TBHP sebagai agen pengoksidaan pada OMS-2 yang telah ditukarkan dengan logam peralihan memperlihatkan bahawa TiOMS-2 yang mempunyai kandungan titanium yang tinggi menghasilkan peratus penukaran yang paling tinggi, mungkin disebabkan oleh wujudnya tapak titanium dalam dan luar bingkaian. Untuk penggabungan logam di dalam bingkai, ada hubungkait antara jejari ion logam yang menukarganti dengan peratus penukaran sikloheksana. Peningkatan jejari ionik daripada logam yang menukarganti meningkatkan peratus penukaran sikloheksana. Hubungkait ini mungkin disebabkan oleh peningkatan tapak asid Lewis pada logam yang digabungkan dengan OMS-2. Kajian terhadap aktiviti pemangkinan ke atas mangkin tertukarganti H di dalam pengoksidaan menunjukkan peningkatan peratusan penukaran daripada sikloheksana setelah penukaran ion. Untuk kajian selanjutnya, Ti-OMS-2 dengan nisbah Ti:Mn yang lebih tinggi telah digunakan dalam pengoksidaan sikloheksena dan stirena dengan TBHP sebagai agen pengoksida. Didapati bahawa kedua-dua tapak titanium dalam dan luar bingkai meningkatkan aktiviti OMS-2 dalam pengoksidaan sikloheksena. Walau bagaimanapun, telah dibuktikan bahawa hanya spesies titanium bukan-bingkai sahaja menghasilkan kesan sinergi yang meningkatkan pengoksidaan stirena. Didapati bahawa wujud hubungkait antara lokasi tapak titanium pada mangkin Ti-OMS-2 dengan pengaktifan ikatan C-H dan C=C. Tapak titanium dalam bingkai memainkan peranan hanya untuk pengaktifan ikatan C-H sahaja sedangkan tapak titanium luar bingkai meningkatkan aktiviti pemangkinan pada kedua-dua jenis pengaktifan ikatan. Sampel SO42-/Ti-OMS-2 aktif untuk pertukaran berterusan 1-oktena kepada 1,2 oktanadiol. Bagaimanapun, telah dipastikan bahawa tidak ada tapak asid Brönsted pada sampel SO42-/Ti-OMS-2. Kejayaan daripada tindak balas berterusan tersebut mungkin disebabkan oleh pembentukan asid Brönsted daripada hidrolisis air yang terjadi pada tapak asid Lewis yang terdapat pada sampel SO42-/Ti-OMS-2. vii TABLE OF CONTENTS CHAPTER TITLE DECLARATION ii DEDICATION iii ACKNOWLEDGEMENTS iv ABSTRACT v ABSTRAK vi TABLE OF CONTENTS vii LIST OF TABLES xii LIST OF FIGURES xv LIST OF SYMBOLS/ABBREVIATIONS xxi LIST OF PUBLICATIONS 1 xxiv INTRODUCTION 1 1.1 Research Background 5 1.1.1 Porous Manganese Oxide Materials 5 1.1.2 Manganese Oxide Octahedral Molecular Sieves (OMS-2) 7 1.1.3 Modification of OMS-2 Materials 9 1.1.4 Designing of Sulfated Ti-OMS-2 as Bifunctional Oxidative and Acidic Catalyst 2 PAGE 12 1.2 Research Questions and Scope of the Research 13 1.3 Research Objectives 14 LITERATURE REVIEW 18 2.1 Introduction 18 2.2 Classification of Porous Manganese Oxide 18 viii 2.3 2.4 2.5 Material Synthesis 24 2.3.1 Synthesis of Porous Manganese Oxide 24 2.3.2 Synthesis of OMS-2 Materials 26 Modification of Manganese oxide 29 2.4.1 29 Overview of Modification of OMS-2 materials 2.4.2 Ion Exchange 37 2.4.3 Impregnation 38 2.4.4 Alkylsilylation 39 2.4.5 Sulphation 41 Catalytic Activity 42 2.5.1 Catalytic Activity of OMS-2 Materials 42 2.5.2 Catalytic Oxidation 45 2.5.3 Titanium Incorporated Materials 49 2.5.4 52 Sulphated Metal Oxides as a Solid Acid Catalyst 2.5.5 Synthesis of Diols 3 53 EXPERIMENTAL 55 3.1 Synthesis of OMS-2 Materials 55 3.1.1 Synthesis of OMS-2 without Buffer (OMS-2a) 57 3.1.2 Synthesis of OMS-2 with buffer (OMS-2b) 57 Modification of OMS-2 materials 57 3.2.1 Synthesis of Metal Substituted OMS-2 (M-OMS-2) 57 3.2.2 Ion Exchange of OMS-2 and M-OMS-2 Samples 58 3.2.3 Synthesis of Ti Incorporated OMS-2 (Ti-OMS-2) 59 3.2.4 Synthesis of Ti Impregnated OMS-2 [Ti-OMS-2 (imp)] 59 3.2 3.2.5 Preparation of TiO2-OMS-2 (mix) 3.2.6 Synthesis of Sulphated Ti-OMS-2 3.3 59 (SO42-/Ti-OMS-2) 60 3.2.7 Surface Modification by Alkylsilylation 60 Characterization Techniques 60 3.3.1 X-Ray Diffraction (XRD) Spectroscopy 61 3.3.2 Atomic Absorption Spectroscopy (AAS) 62 3.3.3 Fourier Transform Infrared (FTIR) Spectroscopy 63 3.3.4 Total Specific Surface Area (BET) and Pore Volume . ... .....Analysis 65 ix 3.3.5 Thermal Gravimetry and Differential Thermal Analysis ......(TG-DTA) 67 3.3.6 Field Emission Electron Scanning Microscopy (FESEM) 68 3.3.7 Photoluminescence 68 3.3.8 X- Ray Photoelectron Spectroscopy (XPS) 70 3.3.9 71 Pyridine Adsorption 3.3.10 Adsorption Capacity of Adsorbed Water 74 3.3.11 Gas Chromatography (GC) Analysis 74 3.3.12 Gas Chromatography-Mass Spectrometry (GC-MS) 3.4 4 ......Analysis 75 Catalytic Testing 75 3.4.1 Oxidation of Benzyl Alcohol 77 3.4.2 Oxidation of Cyclohexane 78 3.4.3 Oxidation of Cyclohexene 79 3.4.4 Oxidation of Styrene 80 3.4.5 Transformation of 1-octene to 1,2-octanediol 81 PHYSICOCHEMICAL PROPERTIES OF OMS-2 AND MODIFIED OMS-2 CATALYSTS 83 4.1 Introduction 83 4.2 Physicochemical Properties of Prepared OMS-2 by Different Methods 84 4.3 Physical Properties of Metal Substituted OMS-2 Material 92 4.3 Physical Properties of H-OMS-2 and H-M-OMS-2 Materials 102 4.4 Physicochemical Properties of Ti-OMS-2 Materials 106 4.4.1 Structural Properties of Ti Substituted OMS-2 Catalyst 106 4.4.2 Acidity Properties 113 4.4.3 114 Morphology, Surface Area and Textural Properties 4.4.4 Thermal Stability 4.4.5 119 More Evidence of the Location of Titanium on ............... .....Ti-OMS-2 Materials 121 4.5 Alkylsilylated of OMS-2 and Ti-OMS-2(0.67) 127 4.6 Sulphated Ti-OMS-2 129 x 5 CATALYTIC ACTIVITY OF OMS-2 AND MODIFIED OMS-2 SAMPLE IN OXIDATION AND ACID REACTIONS 135 5.1 Introduction 135 5.2 Catalytic Activity and Selectivity of OMS-2 and Modified OMS-2 Samples in Oxidation Reactions 5.2.1 135 Oxidation of Benzyl Alcohol over OMS-2 Prepared by .....Different Method 5.3.2 135 Oxidation of Cyclohexane over Metals Substituted and .....Ion.Exchanged OMS-2 5.4.3 Oxidation of Cyclohexene over Ti-OMS-2 Catalyst 140 150 5.4.2 Oxidation of Styrene over Different Location of .....Titanium sites on Ti-OMS-2 Catalysts 5.3 The Effect of Lewis Acidity in Catalytic Oxidations 5.4 The Role of Different Location of Ti Sites in Ti-OMS-2 in Oxidation Reactions 5.5 156 157 Catalytic Study on Consecutive Reaction of 1-octene to 1,2-octanediol 6 152 158 SUMMARY AND CONCLUSION 161 6.1 Summary 161 6.2 Conclusion 166 REFERENCES 171 APPENDIXES 189-199 xi LIST OF TABLES TABLE NO 1.1 TITLE PAGE The synthetic manganese oxides, their natural counterpart and structures as reported by Suib [12, 20-22]. 1.2 2.1 The cations were doped OMS-2 and their location in OMS-2 material. 2.3 2.4 templates. 26 Review of Metal Doped into OMS-2; its synthesis routes, location, properties, and catalytic application. 30 Some catalytic applications of OMS-2 materials reported in the last 42 Characteristic features of Shell’s epoxidation catalyst compared with Titanium Silicalite 1. 2.6 19 Synthesis of tunnel and layered manganese oxides with various decade. 2.5 10 Classification of tunnel and layered manganese oxides and their crystallographic data [13]. 2.2 7 49 The comparisons of the common mild oxidant are used in oxidation process. 50 3.1 Position of bands and classification for linkages of pyridine. 72 4.1 Effect of potassium concentration in reflux method on the synthesis of OMS-2. 4.2 The relative intensity and ratio of I(110)/I(200) plane of OMS-2b and calcined OMS-2b samples calculated by XRD. 4.3 85 88 The ionic radii of metals ion and average crystallite size of OMS-2 and M-OMS-2 samples. 94 xii 4.4 Lattice parameter (a and c) and cell volume (V) of OMS-2 and MOMS-2 samples. 4.5 95 The relative intensity and ratio of I(110)/I(200) plane of OMS-2 and M-OMS-2 samples. 96 4.6 The physical properties of metal ions. 97 4.7 Source of metal, its charge and their effect to cryptomelane structure. 98 4.8 AAS data of OMS-2 and M-OMS-2 samples. 99 4.9 The relative intensity and ratio of I(200)/I(211) plane of OMS-2 and M-OMS-2 samples calculated by XRD. 102 4.10 The percentage of potassium substituted by H+. 104 4.11 Chemical composition and physicochemical properties of OMS-2, Ti-OMS-2 and TiO2–OMS-2. 4.12 107 The lattice parameters (a and c) and cell volume (V) of OMS-2 and Ti-OMS-2 samples. 109 4.13 Vibrational spectroscopy feature of samples. 111 4.14 The relative intensity and ratio of I(111)/I(211) plane of samples calculated by XRD. 4.15 116 Binding Energies (eV) of Mn 2p, Ti 2p, and its line separation (BE) and difference of line separation from selected samples. 122 4.16 Assignments of as-observed IR bands on sulfated samples [154]. 131 5.1 Conversion of benzyl alcohol and selectivity to benzaldehyde by different catalysts. 5.2 136 The relation of amount of potassium exchanged by H+ with enhancement of % conversion of cyclohexane on H-M-OMS-2 catalyst. 5.3 144 The correlation of Lewis acidity of samples to conversion of cyclohexane, cyclohexene and styrene. 156 xiii 5.4 The possible role of Lewis acids of catalyst in oxidation of cyclohexane, cyclohexene and styrene. 157 5.5 The role of Ti sites location in oxidation reaction. 158 6.1 The physicochemical properties-catalytic activity relationship of the catalysts. 163 xiv LIST OF FIGURES FIGURE NO 1.1 TITLE PAGE Steps Potential energy diagram for a catalytic reaction (solid line), i.e. reaction of A and B to form AB, compared with the noncatalytic reaction (dashed line). The presence of a catalyst lowers the activation energy (Ea) considerably [7]. 1.2 Process options in catalytic oxidation and some of their disadvantages. 1.3 2 3 Crystal structure of cryptomelane-type OMS-2: potassium atoms are shown as green spheres; MnO6 octahedra are shown in brown. 8 1.4 Mars van Krevelen mechanism. 9 1.5 Partial periodic table showing transition metals that have been introduced into OMS-2 materials. 10 1.6 Transformation of alkenes to diols via two step reaction. 12 1.7 The schematic of the research approach and research questions. 15 1.8 The significant of the use Ti-OMS-2 in heterogeneous oxidation and acid reactions. 2.1 Schematic structures of one-dimensional tunnel and layered manganese oxides [13]. 2.2 2.4 22 Intergrowth tunnels of (a) (1×1) and (1×2), and (b) (2×2) and (2×3) in the tunnel manganese oxides. 2.3 17 23 Transformation reaction from birnessite to hollandite under hydrothermal conditions. 28 The ion exchange reaction. 37 xv 2.5 Impregnation of porous catalyst. 38 2.6 Chemical structure of OTS. 39 2.7 Mechanism of complete reaction of OTS on the support surface to form a well ordered layer. 40 2.8 Mechanism of incomplete reaction of OTS on the support surface. 40 2.9 Epoxidation of alkenes using (a) organic peracids, (b) chlorohyrins route and (c) H2O2 or TBHP as the oxidant. 2.10 47 Epoxidation of propene on Shell catalyst using an organic peroxide. 50 2.11 Oxidation reactions catalyzed by TS-1. 51 3.1 Materials preparation and their labelling. 56 3.2 The IUPAC classification for adsorption isotherms, where nad = amount of adsorbed and P/Po= relative pressure. 3.3 The physical process following absorption of a photon by a molecule. 3.4 69 The mechanism of interaction between pyridine molecules with Lewis. 3.5 66 73 The interaction between pyridine molecules with Brönsted acid sites.Catalytic reactions over various modified OMS-2. 73 3.6 Catalytic reactions over various modified OMS-2. 76 4.1 XRD patterns of OMS-2 materials and reference pattern of Crypromelane, Q JCPDS 29, 1020. 4.2 86 Effect of calcination on XRD patterns of OMS-2b materials, (a) before calcination, (b) calcination at 400oC, (c) at 500oC, and (d) at 600oC. 88 4.3 FTIR spectra of (a) OMS-2a and (b) OMS-2b. 89 4.4 Defect on OMS-2 structure. 90 xvi 4.5 Amount of adsorbed water on the surface of OMS-2a and OMS-2b samples. 90 4.6 TGA plots for OMS-2 materials in N2 atmosphere. 91 4.7 X-ray diffractograms of (a) OMS-2; (b) Ti-OMS-2(0.05); (c) Fe-OMS-2(0.09); (d) Co-OMS-2(0.02); and (e) Cu-OMS-2(0.04). 93 4.8 Schematic incorporation of metals in M-OMS-2 materials. 4.9 XRD pattern of (a) cryptomelane and (b) Ti-, (c) Cu-, (d) Co- and 93 (e) Fe-OMS-2 was prepared by oxidation of its metal ions source by potassium permanganate in acidic condition without the addition of Mn2+ solution. 97 4.10 Morphology of OMS-2 and M-OMS-2 samples. 101 4.11 Schematic synthesis of H-OMS-2 and H-M-OMS-2. 102 4.12 XRD pattern of (a) OMS-2 and (b) H-OMS-2. 103 4.13 Morphology of H-OMS-2 and H-M-OMS-2 samples. 105 4.14 X-ray diffractograms of (a) cryptomelane (JCPDS 29, 102), (b) OMS-2, (c) Ti-OMS-2 (0.18), (d) Ti-OMS-2 (0.43), (e) Ti-OMS-2 (0.67), (f) TiO2-OMS-2 (imp) and (g) Ti-OMS-2 (mix). 4.15 IR spectra at lower wavelength region of (a) OMS-2, (b) Ti-OMS-2 (0.18), (c) Ti-OMS-2 (0.67). 4.16 111 Photoluminescence spectra of OMS-2, Ti-OMS-2 (0.43) and TiO2-OMS-2 (mix). The excitation wavelength is 430 nm. 4.18 110 IR spectra at higher wavelength region of (a) OMS-2, (b) Ti-OMS-2 (0.18), (c) Ti-OMS-2 (0.67). 4.17 108 113 FTIR spectra of (a) Ti-OMS-2 (0.67) and (b) OMS-2 after evacuation under vacuum at 400 oC for 4 h followed by pyridine adsorption at room temperature and evacuation at 150 oC for an hour. 4.19 114 Morphology of (a) Ti-OMS-2 (0.18), (b) Ti-OMS-2 (0.18) and (c) TiO2-OMS-2 (imp). 115 xvii 4.20 N2 adsorption isotherm for OMS-2 at 77 K. 4.21 N2 adsorption isotherm for (a) Ti-OMS-2(0.43) and (b) Ti-OMS-2(0.67) at 77 K. 4.22 118 TGA profile (a) in original and (b) differential forms of OMS-2, Ti-OMS-2 (0.18) and TiO2-OMS-2 (imp). 4.23 116 120 Detailed XPS spectra for the Mn 2p transition for (a) OMS-2, (b) Ti-OMS-2(0.18), (c) Ti-OMS-2(0.67) and (d) TiO2-OMS-2(imp). 123 4.24 Bond strength on bridging oxygen atom. 4.25 Detailed XPS spectra for the Ti 2p transition for (a) Ti-OMS-2 (0.18), (b) Ti-OMS-2(0.67) and (c) TiO2-OMS-2(imp). 4.26 126 FTIR spectra of modified OTS samples (a) OTS/OMS-2 and (b) OTS/Ti-OMS-2. 4.27 125 128 Percentage of adsorbed water on the sample (a) OMS-2, (b) Ti-OMS-2(0.67), (c) OTS/Ti-OMS-2(0.67) and (d) OTS/OMS-2. 128 4.28 Proposed polymeric octadecylsiloxane on the surface of OMS-2 and Ti-OMS-2 samples. 4.29 129 XRD pattern of Ti-OMS-2(0.04) and sulphated Ti-OMS-2 (0.04). # = TiOSO4.H2O,* = MnSO4 7H2O ¤ = MnSO4. 130 4.30 FTIR spectra of Ti-OMS-2 and sulphated Ti-OMS-2. 131 4.31 The bridging of bidentated structure of sulphated Ti-OMS-2. 132 4.32 Thermograms (TGA) of samples. 133 4.33 First derivative curves (DTGA) of samples. 133 4.34 FESEM micrograph of SW150-Ti-OMS-2(0.67) sample in different magnitude. 134 5.1 The schematic reaction of benzyl alcohol to benzaldehyde. 136 5.2 Overall alcohol oxidation mechanism [24]. 139 5.3 Resonance model of Mn-O-Mn bond structure: Resonance structure in crystalline OMS-2 (Structure A); and non-resonance structure in amorphous materials (Structure B). 140 xviii 5.4 Schematic reaction of cyclohexane. 5.5 The conversion and product selectivity of oxidation of 141 cyclohexane with tert-butyl hydroperoxide (TBHP) using OMS-2, M-OMS-2 and H-M-OMS-2. All reactions were carried out at 60 ºC for 24 h with cyclohexane (26 mmol), 70% aqueous TBHP (10 mmol), and catalyst (50 mg) under reflux condition. 5.6 142 The relationship of ionic radii of metals substituted OMS-2 to conversion of cyclohexane. 144 5.7 Yield of products vs time on Ti-OMS-2. 146 5.8 Homolytic pathway to form radical from TBHP over catalyst. 148 5.9 Formation of cyclohexyl hydroperoxide. 148 5.10 The heterolytic pathway of the formation of cyclohexanone from cyclohexyl hydroperoxide. 5.11 The homolytic pathway of the formation of cyclohexanol from cyclohexyl hydroperoxide. 5.12 148 149 The heterolytic pathway of the formation of cyclohexanol and cyclohexanone from 1,4-dicyclohexyltetraoxidane. 149 5.13 Reaction condition of cyclohexene and its products. 150 5.14 The conversion and product selectivity of oxidation of cyclohexene with tert-butyl hydroperoxide (TBHP) using TiO2, OMS-2, Ti-OMS-2(0.18), and Ti-OMS-2(0.67). {All reactions were carried out at 70 ºC for 2 h with cyclohexene (5 mmol), 70% aqueous TBHP (10 mmol), acetonitrile (15 ml) and catalyst (50 mg). The conversion and the amount of product obtained in blank experimental have been subtracted}. 5.15 Oxidation of styrene and its product on catalysts using TBHP as oxidant. 5.16 151 The conversion and product selectivity of oxidation styrene with tert-butyl hydroperoxide (TBHP) using TiO2, TiO2-OMS-2, Ti-OMS-2, OMS-2 and TS-1. All reactions were carried out at 153 xix 70 oC with styrene (5 mmol), 70% aqueous TBHP (10 mmol), acetonitrile (15 ml) and catalyst (50 mg) with vigorous stirring. 5.17 154 Consecutive oxidation and acid reaction to form of 1,2 octane diol from 1-octene on sulphated Ti-OMS-2 catalyst. 159 5.18 Yield of epoxyoctane and 1,2 octane diol after 24 h reaction. 159 6.1 Assignments of modified OMS-2 in oxidation of cyclohexane and consecutive reaction of 1-octene to 1,2 octanediol consecutive reaction of 1-octene to 1,2 octanediol 6.2 167 The role of the location of Ti sites in Ti-OMS-2 in oxidation of cyclohexane, cyclohexene and styrene. 169 xx LIST OF SYMBOLS/ABBREVIATIONS % - percent ~ - approximately μ - micron (10-6) 2 - Bragg angle Å - angstrom (10-10) a.u. - arbitrary unit BET - Brunnauer, Emmett and Teller c.a. - about (Latin:circa) cm-1 - per centimeter 2+ Co - Cobalt ion Co-OMS-2 - Cobalt substituted OMS-2 Cu K - X-ray diffraction from copper K energy levels Cu2+ - Copper ion Cu-OMS-2 - Copper substituted OMS-2 d - distance DTG - Differential thermogravimetry e.g. - example (Latin: exempli gratia) EDAX - Energy dispersive analysis by X-ray equilibrium pressure and vapour pressure Po of the adsorbate at the temperature where the isotherm is measured et al. - and others (Latin: et alia) eV - electrovolt Fe - Iron ion Fe-OMS-2 - Iron substituted OMS-2 FESEM - Field Emission Scanning Electron Microscope FID - Flame ionisation detector FTIR - Fourier transform infrared 3+ xxi g - grams h - hour H2 O2 - Hydrogen peroxide HF - Hydrofluoric acid i.e. - that is (Latin : id est) IUPAC - International Union of Pure and Applied Chemistry K - degree Kelvin KBr - Potassium bromide M - Molar mg - meter square per gram mA - milliampere min - minute mL - millilitre mol - mole M-OMS-2 - Metal substituted OMS-2 N2 - Molecular nitrogen nm - nanometer (10-9) o - degree celcius OL-1 - OL-1 is manganese oxide Octahedral layered with interlayer spacing of ~7 Å. OMS-1 - Manganese oxide octahedral molecular sieve with 3 x 3 tunnel structure. OMS-2 - Manganese oxide octahedral molecular sieve with 2 x 2 tunnel structure. OMS-5 - Manganese oxide octahedral molecular sieve with 2 x 4 tunnel structure. OMS-6 - Manganese oxide octahedral molecular sieve with 2 x 3 tunnel structure. OMS-7 - Manganese oxide octahedral molecular sieve with 1 x 1 tunnel structure. OTS - Octadecyltrichlorosilane P/Po - relative pressure; obtained by forming the ratio of the equilibrium pressure and vapour pressure po of the adsorbate at the temperature where the isotherm is measured SO42-/Ti-OMS-2 - Sulphated Ti-OMS-2 ST150-Ti-OMS-2 - Sulphated Ti-OMS-2 with 150 L concentrated H2SO4 using toluene as solvent 2 -1 C xxii ST200-Ti-OMS-2 - Sulphated Ti-OMS-2 with 200 L concentrated H2SO4 using toluene as solvent SW150-Ti-OMS-2 - Sulphated Ti-OMS-2 with 150 L concentrated H2SO4 using water as solvent SW200-Ti-OMS-2 - Sulphated Ti-OMS-2 with 200 L concentrated H2SO4 using toluene as solvent t - crystallite size TBHP - tert-butyl hydroperoxide TGA - Thermogravimetry analysis Ti - Titanium ion TiO2 - Titanium dioxide 3+ TiO2-OMS-2 (imp) - Impregnation of TiO2 on OMS-2 surface TiO2-OMS-2 (mix) - Physical mixture of rutile TiO2 and OMS-2 Ti-OMS-2 - Titanium substituted OMS-2 TS-1 - Titanium Silicate-1 wt % - weight percentage - wavelength xxiii LIST OF PUBLICATIONS AND PRESENTATIONS 1. H. Nur, F. Hayati, H. Hamdan, "On the location of different titanium sites in TiOMS-2 and their catalytic role in oxidation of styrene", Catalysis Communications, 8 (2007) 2007-2011. 2. F. Hayati, H. Nur, H. Hamdan, "Titanium Doped Octahedral Manganese Oxide Hybrid Catalyst in the Oxidation of Cyclohexene", Buletin Kimia, 21 (2005) 4954. 3. F. Hayati, H. Hamdan, H. Nur, "Synergetic effect of titanium and OMS-2 as TiOMS-2 hybrid catalyst in oxidation of cyclohexene", Book abstract of Annual Fundamental Science Seminar 2005, 4-1 July 2005, Johor Bahru, Malaysia. p. 67. 4. F. Hayati, H. Nur and H. Hamdan, "Synthesis and characterization of octahedral molecular sieves (OMS-2)", Book of abstract of Annual Fundamental Science Seminar 2004, 14-15 June 2004, Johor Bahru, Malaysia. p. 73. CHAPTER 1 INTRODUCTION A catalyst is a substance which accelerates a chemical reaction. The basic principle of how a catalyst works for a chemical reaction is shown by the example in Figure 1.1. It does so by forming bonds with the reacting molecules (i.e. adsorption), followed by breaking and weakening of the intramolecular bonds. Next, the adsorbed species react on the surface to a particular product, often in several consecutive steps. Finally, it detaches itself from the catalyst (i.e. desorption) and leaves the catalyst unaltered so that it is ready to interact with the next set of molecules. A catalyst cannot alter the chemical equilibrium of a given reaction; it only creates a favourable reaction pathway. This is done by decreasing the activation barrier (Ea,cat) compared to non catalytic reaction (Ea,non) and thus increasing the reaction rate. In general, a successful catalyst increases the yield of the desired product while decreasing that of other products, which has advantages for both economic and environmental reasons [1]. The first introduction of the word ‘catalysis’ was by Berzelius in 1836, while Ostwald presented the first correct definition of a catalyst in 1895 [2]. He described a catalyst as a substance that changes the rate of a chemical reaction without itself appearing in the products. Today, catalysis lies at the heart of our quality of life: the reduced emissions of modern cars, the abundance of fresh food at our stores, and the new pharmaceuticals that improve our health are made possible by chemical reactions controlled by catalysts [3]. It covers multidisciplinary science that serves a broad range of chemical industries covering specialty, fine, intermediate, commodity, 2 and life science chemicals [4]. It played a major role in establishing the economic strength of the chemical and related industries in the first half of the 20th century and an estimated 90% of all of the chemical processes introduced since 1930 depend on catalysis [5]. According to a report from Freedonia [6] the world catalyst demand will rise by 3.6% per year to $12.3 billion in 2010. Figure 1.1: Potential energy diagram for a catalytic reaction (solid line), i.e. reaction of A and B to form AB, compared with the non-catalytic reaction (dashed line). The presence of a catalyst lowers the activation energy (Ea) considerably [7]. Catalysis plays an important role in the green chemical processes, which is to minimize environmental impact and to reduce costs of the process. It is crucial to achieve the “ideal synthesis” which would be atom efficient, safe, one step, involving no wasted reagents, based on renewable resources, and environmentally acceptable to overcome some of the biggest problem areas in synthetic methodology such as in oxidation and acid-catalyze reactions [8]. In catalytic process, for example catalytic oxidation, there are some options to consider, i.e. phase (gas or liquid) and catalytic system as shown in Figure 1.2. Firstly, the selection of either a gas or liquid phase depends largely on the boiling 3 point and thermal stability of the reactants, especially of a desired fine chemical intermediate. However, catalytic gas phase is limited to relatively simple molecules. More complicated molecules cannot be brought easily into the gas phase and many of these molecules are unstable at elevated temperatures. In addition, the gas phase needs special reactor and costly. Therefore, working in liquid phase is relatively easier than in gas phase, as simpler equipments are required which can be performed for more complicated molecules. Catalytic oxidation Gas phase x x High temperature x Most compounds not stable at elevated temperature Liquid phase Homogeneous Heterogeneous x Low thermal stability x Difficulty of recovery Easy recovery and regeneration of catalyst Simple molecule only and regeneration Figure 1.2: Process options in catalytic oxidation and some of their disadvantages. As also shown in Figure 1.2, another important choice to consider is whether the oxidation should be performed using homogeneous or heterogeneous catalyst. In homogeneous catalysis, the reaction mixture and the catalysts are all in the same phase, usually the liquid phase. The catalyst may be a metal complex, which is dissolved in a solvent together with the reagents. Homogeneous catalysts are characterized by high activity and selectivity. The main advantage of homogeneous catalysts is the ease of accessibility of the active site, resulting in a high activity, no mass transfer limitations and generally low temperature and pressure requirements. The main disadvantages of this type of catalysts are low thermal stability, difficulty of recovery and regeneration. In combination with high cost of catalyst, this makes homogeneous catalysis less popular in industry. Homogeneous catalysts are mostly found in batch processes where volumes are small and the added value is high, e.g., in pharmaceuticals. 4 The problems faced by homogeneous catalyst can be solved by using heterogeneous catalyst [1, 2, 5]. In heterogeneous catalysis, the catalyst is in a different phase than the reaction mixture. Usually, the catalyst is a solid substance (e.g., a metal or a metal oxide) and the reaction mixture can be a gas or a liquid. Heterogeneous catalyst may be easily recovered by filtration and recycle. The success of this type of catalyst is due to the ease at which it can be applied in all types of reactions, carried out in both continuous and batch mode. It is relatively easy to separate the catalyst from the reaction mixture and reuse it. The solid, heterogeneous phase, however, may introduce mass transfer limitations, presumably resulting in a lower activity or selectivity. Therefore, the selection of suitable active site for heterogeneous catalyst is a challenge in academic and industry. Many heterogeneous or solid catalysts are based on porous inorganic solids [5]. The important physical properties of these materials are surface area (often very large and measured in hundreds of m2/g), pore volume, pore size distribution (which can be very narrow or very broad), the size and shape of the particles and their strength. The solid catalyst provides a surface, usually large internal, for the substrates to adsorb and react on. Thus the surface characteristics (roughness, functional groups, organophilicity, hydrophobicity, etc.) are also vital to performance. Following the definition accepted by the International Union of Pure and Applied Chemistry (IUPAC), porous materials can be grouped into three classes based on their pore diameter (d): microporous, d < 2.0 nm; mesoporous, 2.0 < d < 50 nm; macroporous, d > 50 nm. In the class of microporous materials, zeolites and related materials such as aluminophosphates have found for a long time applications outside the traditional areas of acid and bifunctional catalysis [9-14]. With the introduction of the ordered micelle-templated inorganic materials, the choice of available supports has been considerably extended into the mesoporous domain. Examples are M41S groups materials such as MCM-41, MCM-48, MSU, HMS, FSM-16, and various SBA type materials [9, 10]. Macroporous metal oxides such as titania, zirconia and alumina are example for macroporous material. However, all of the above materials in general are insulating materials. In addition they are often 5 synthesized with charge compensation in mind. For example, Al3+ substituted for Si4+ in zeolites lead to an inherent cation-exchanged capacity. Another way to approach the generation of microporous material is to generate element with mixed valencies in a structure which should also lead to cation exchange capacity [11]. Porous manganese oxides with mixed valencies could be a challenge due to wider application over aluminosilicate materials e.g. redox catalysis, battery and sensor [12]. 1.1 Research Background 1.1.1 Porous Manganese Oxide Materials Porous manganese oxide is one of the largest families of porous materials. These are two major structures of porous manganese oxides, tunneled and layered materials with pore size from ultra-micropores to mesopores [13]. Natural manganese oxides are found abundant as manganese nodules which exist in a wide variety of locations such as the ocean floor, the beds of many fresh-water lakes, rocks, and soil. Manganese oxide minerals have been used for thousands of years as pigments and to clarify glass, and more recently as ores of Mn metal, catalysts and battery material [14]. Porous manganese oxide materials have been described as useful catalysts in the oxidation of carbon monoxide, methane and butane, the reduction of nitric oxide with ammonia and demetallation of topped crude in the presence of hydrogen [15]. However, the structures and properties of the natural source are usually not uniform and its activity is not reproducible. The porous manganese oxides have been extensively investigated for their economic value and their potential applications. Due to their excellent cation- exchange and molecule adsorptive properties, like the aluminosilicates, these 6 manganese oxides can be used as ion-sieves, molecular-sieves, and catalysts similar to the aluminosilicates. In contrast to aluminosilicate based zeolite, these manganese oxides are mixed valencies materials (mainly, 4+, 3+, or 2+). The mixed valency materials are important in biology, chemistry, and physics which occur in manganese redox enzymes, in natural manganese oxide nodules, chemiluminescence systems, electron transfer and electrocatalysis, electrochromism, secondary nonaqueous rechargeable batteries, magnetics, ceramics, and biological systems. In chemistry, mixed valency in mixed metal oxides is important for electron transport. For example, the effectiveness of metal oxides as catalysts for redox reactions, as electrode materials for electrochemical processes, and as chemical sensors for reductive gases are usually governed by their ability and tendency to cycle between different valence states of relevant cations and the mobility of oxygen ions. From this standpoint, manganese oxide materials have distinct advantages over aluminosilicate molecular sieve materials for applications in redox catalysis, batteries, and chemical sensors [12]. The synthetic manganese oxides have been intensively studied by Suib and his group since 1990’s[16-18]. The synthetic names, the natural minerals and unit structures are listed in Table 1.1. In general manganese oxide materials can be classified into two kinds i.e. tunnel and layered structures. The term of octahedral molecular sieves (OMS) and octahedral layered (OL) materials are referred to the synthetic manganese oxides with tunnel and layered structure, respectively [19]. OMS materials comprise infinite 3-D crystalline frameworks with molecule-sized tunnels similar to the naturally occurring zeolites. The structural frameworks of the manganese oxides consists of MnO6 octahedral units shared by corners and/or edges in comparison with, in general, SiO4-AlO6 frameworks of the porous aluminosilicates. OMS-1 and OMS-2 are two types of OMS materials which are built of 3x3 and 2x2 MnO6 octahedral units, respectively. The other types are pyrolusite-, ramsdellite-, and romanechite-type manganese oxides which have onedimensional (1×1), (1×2) and (2×3) tunnel structures, respectively. OL materials have interlayer spacings that are similar to clay-type materials. OL-1 is synthetic name of birnessite which has the interlayer spacing of ~7 Å and it contains exchangeable cations and water molecules. 7 Table 1.1: The synthetic manganese oxides, their natural counterpart and structures as reported by Suib [12, 20-22]. Synthetic Name Natural counterpart MnO6 octahedral units OMS-1 todorokite 3x3 OMS-2 cryptomelane 2x2 OMS-5 - 2x4 OMS-6 romanechite 2x3 OMS-7 pyrolusite 1x1 OL-1 birnessite 2x~ Among most of the OMS and OL materials, OMS-1 and OMS-2 materials have been used widely as catalysts, chemical sensors and batteries due to their mixed-valency, high porosity, thermal stability, surface areas, and inexpensive manufacturing cost. However, OMS-2 materials have been used as catalyst even more widely than OMS-1 materials since they are more active and selective in catalytic oxidation [20]. 1.1.2 Manganese Oxide Octahedral Molecular Sieves (OMS-2) As mentioned before OMS-2 materials are synthetic manganese oxide which is built of 2x2 MnO6 octahedral and also known as synthetic cryptomelane [11]. The tunnel size of this material is 4.6 Å x 4.6 Å and potassium ion exists in the tunnel to balance the charge of the structure as shown in Figure 1.3. General composition of OMS-2 materials is K0.8–1.5Mn8O16. These materials are not expensive and easy to prepare compared to other manganese oxide type materials. For its application in catalytic reaction, OMS-2 materials were reported to be selective in oxidation of benzyl alcohol [23, 24]. Besides, OMS-2 materials are potential electrocatalysts for the oxidation of methanol, for fuel cell applications 8 [25], are also active catalyst for the total oxidation of benzene and ethanol [26] and epoxidation of olefins [27, 28]. The mild oxidants i.e. molecular oxygen and terbutyl hydrogen peroxide (TBHP) used as oxidizing agents on OMS-2 materials have let to these materials being promising catalysts in fulfilling the environmental concern and regulations for clean environment. Figure 1.3: Crystal structure of cryptomelane-type OMS-2: potassium atoms are shown as green spheres; MnO6 octahedra are shown in brown. The high oxidation ability of OMS-2 materials has mainly been related to two factors: the presence of Mn2+/Mn4+ or Mn3+/ Mn4+ redox couples and the ability of active participation of the lattice oxygen in these systems in oxidation process leading to a Mars van Krevelen type of oxidation mechanism [23, 24]. The mechanism involves two steps as illustrated in Figure 1.4. Firstly, the lattice oxygen oxidizes the substrate molecule, followed by a reoxidation of the partially reduced catalyst by molecular oxygen in order to regenerate the catalyst. To date, however, the report on liquid phase oxidation via Mars-van Krevelen-type mechanism is limited, i.e., homogeneous catalysis by phosphovanadomolybdate [29] and heterogeneous catalysis by OMS-2 catalyst [23, 9 24]. The high selectivity of product makes this mechanism interesting and worthwhile for intensive study. Figure 1.4: Mars van Krevelen mechanism. 1.1.3 Modification of OMS-2 Materials In order to alter their structures and properties and generate better electronic and catalytic performance, the modification was done by doping cation into OMS-2 materials. The cations doped into OMS-2 materials are listed in Table 1.2. These cations can substitute potassium and/or manganese ions which exist in the tunnel and framework structure, respectively [30]. Doping of alkali metals or NH4+ or H+ was reported to substitute some potassium ions which exist in the tunnel structure of OMS-2 materials. Transition metal oxides were incorporated in the framework and/or exist in the tunnel structure. The existence of metal oxides in the framework indicated that those metal oxides were substituted for manganese. When metals exist in a tunnel structure, it suggests that the metals have replaced potassium. The location of doped metal depends on the preparation method [31]. The transition metal cations are mostly situated in the framework positions of the OMS-2 structure with a priori incorporation. In a posteriori incorporation, the cations are situated in either the tunnel positions (if prepared by ion-exchange) or in extraframework positions (if prepared by homogeneous precipitation). 10 Table 1.2: The cations were doped OMS-2 and their location in OMS-2 material. Doping cations Location References H+/NH4+ In tunnel structure [23, 24, 32] Alkali metals In tunnel structure [32-34] Transition metals In framework and/or tunnel structure [27, 30, 31, 35-46] The summary on incorporation of transition metals into OMS-2 materials is shown in Figure 1.5. Among these transition metals, titanium incorporated OMS-2 has not been reported and therefore is appropriate to be explored. Ti V Cr Mn Fe Co Ni Cu Zn Zr Nb Mo Tc Ru Rh Pd Ag Cd Have been doped into OMS-2 Not studied Figure 1.5: Partial periodic table showing transition metals that have been incorporated into OMS-2 materials. The physical and chemical properties of doped OMS-2 are greatly influenced by the type, amount and location of the doping ions, where the properties of doped materials are significantly different from those of the undoped ones [12]. In the tunnel cations position, H+ doped OMS-2 exhibits the surprising activity in oxidation of alcohols and acid-catalyze condensation of phenylhydroxylamine with aniline to 2-aminodiphenylamine. Among the transition metal doped OMS-2 materials, FeOMS-2 seems to be the best catalyst in some catalytic reactions such as oxidative dehydrogenation of 1-butene [40], oxidation of toluene [41], decomposition of the cyanine dye and pinacyanol chloride [44], and oxidative dehydrogenation of ethanol [46]. However the activity of Fe-OMS-2 is lower than Co-OMS-2 in oxidation of styrene [27]. It suggests that the activity of metal doped OMS-2 also depends on the type of substrates involved in the reactions. 11 Titanium oxide attracts much attention in catalysis as well as photocatalysis and it has been used in the synthesis of many chemical compounds. Titanium incorporated material shows outstanding catalytic properties, particularly in liquid phase oxidation process [47-49]. Shell catalyst (Ti(IV)/SiO2) is the basis of the commercial process for the epoxidation of propene with ethylbenzene hydroperoxide. Thus, the discovery of titanium silicate-1 (TS-1) by Taramasso et al. [50] exhibited the remarkable catalytic activity, selective epoxidations with 30% aqueous hydrogen peroxide under very mild conditions, constituted a milestone in oxidation catalysis. The discovery of TS-1 led to the study on incorporation of titanium into porous materials: microporous material such as silicoaluminophosphate (SAPO-5) and aluminophosphates (AlPO-5, AlPO-11 and AlPO4-36) and mesoporous materials such as MCM-41 and MCM-48. The incorporation of titanium into manganese oxide molecular sieve frameworks is feasible because of similar sizes, charges, and coordination tendencies of manganese and titanium cations. Ionic radii of octahedral Mn3+, Mn4+, and Ti4+ of 0.65 Å, 0.53 Å and 0.61 Å, respectively [54-55] in crystals are close to one another, hence Ti4+ can easily substitute either Mn3+ or Mn4+ without causing much structural disorder and serious charge imbalance. Due to the mixed-valence character of manganese in OMS-2, this material has distinct advantages over silicate, aluminosilicate and aluminophosphate molecular sieve materials in catalytic applications [12, 30]. The incorporation of titanium into OMS-2 is expected to give excellent performance in oxidation reaction. The previous sections have demonstrated that OMS-2 and metal doped OMS2 materials are potential catalysts in oxidation reactions. The catalytic activity of metal doped OMS-2 varies depending on the substrate. Further study of the metal doped OMS-2 is required in order to understand the physicochemical propertiescatalytic activity of the catalysts. Besides that, the effect of titanium incorporated OMS-2 has not been reported. Therefore, the study on the effect of titanium incorporated OMS-2 to the physicochemical properties and catalytic activity of OMS-2 material in oxidation reaction is needed. Further investigation in the potential application of Ti incorporated OMS-2 in consecutive oxidation and acid 12 reaction should be done. The next section describes how Ti incorporated OMS-2 is modified to bifunctional oxidative and acidic catalyst. 1.1.4 Design of Sulphated Ti-OMS-2 as Bifunctional Oxidative and Acidic Catalyst A bifunctional or multifunctional catalyst is the catalyst which has two or more active sites. The catalysts have active sites which can catalyze two or more transformations which are carried out as a “one-pot” process. They offer a number of advantages to the organic chemist: in particular, they result in a reduced number of operations, giving significant time-cost benefits, but they also often allow “difficult” intermediate compounds (i.e., those that are volatile, toxic, or otherwise noxious) to be prepared and elaborated in situ, thus preventing problems associated with their isolation and handling [51]. A bifunctional oxidative and acidic catalyst catalyzes for oxidation and acidreaction, respectively. The catalyst effectively directs transformation of alkenes to alcohols. The oxidative sites catalyze the oxidation of alkenes to epoxide followed by transformation of epoxide to diols in the presence of Brönsted acid site in the catalyst. At present, 1,2-diols are manufactured industrially by a two-step sequence consisting of epoxidation of an olefin with a peracid followed by hydrolysis of the resulting epoxide as shown in Figure 1.6. Figure 1.6: Transformation of alkenes to diols via two step reaction. Recently, Prasetyoko et al. [52-54] reported that titanium silicalite (TS-1) loaded with sulphated zirconia or niobium oxide demonstrated bifunctional oxidative and acidic properties. However, TS-1 is expensive and difficult to prepare. The 13 purpose of this research is to design a novel bifunctional oxidative and acidic catalyst. Combining oxidative OMS-2 and acidity, one can come up with the bifunctional catalyst for consecutive liquid phase oxidation and acid reactions. Sulphated titanium oxide is a solid superacid which exhibits both Lewis and Brönsted acidity [55] which are considered as acid sites. The solid acid is used to overcome the problems of using homogeneous acid. One approach is to create acidity via incorporation of acidic sites in the framework of OMS-2. It is expected that bifunctional oxidative and acidic catalyst can be created by sulphation of TiO2 phase in titanium supported OMS-2 (Ti-OMS2) to form SO42--TiO2 superacid active site. Based on this consideration, a catalytic model of consecutive oxidation and acid catalyst (sulphated Ti-OMS-2) to catalyze consecutive reactions of alkenes to alcohols e.g. 1-octene to 1,2-octanediol could be synthesized. MnO6 octahedra as a basic of OMS-2 unit acts as an oxidative site for epoxidation of 1-octene to 1,2 epoxyoctane and SO42--TiO2 superacid acts as an acidic site for transformation of 1,2 epoxyoctane to 1,2 octanediol. Hence the use of OMS-2 materials which are relatively cheap and easily prepared and use a mild oxidant for oxidation process and sulphated TiO2 environmentally friendly as acid sites is expected to be the best solutions to overcome the related problems. 1.2 Research Questions and Scope of Research Based on the above descriptions, one considers that OMS-2 is the potential heterogeneous catalysts for liquid phase oxidation. Although modification of OMS-2 by doping of transition metal oxides have been reported but Ti incorporated OMS-2 has not been reported yet [30-46]. The effect of the location, amount and synthesis method of titanium incorporated OMS-2 to the catalytic properties of OMS-2 in some oxidation reactions have not been studied yet. The relationship between physicochemical properties and catalytic activity of titanium and other metals incorporated system are not well understood. In addition, design of bifunctional oxidative and acidic catalyst in order to form the more effective and efficient catalyst 14 for consecutive oxidation and acid reactions is a challenge that has less studied. Modification of OMS-2 by enhancement of the oxidative ability and creation of the acid site may together be beneficial and advantageous to create the bifunctional catalyst for synthesis of diols from alkenes. Figure 1.7 shows the research approach and questions of the research. The research includes the study on effect of synthesis method and some modifications to physicochemical properties of OMS-2 and their correlation to catalytic activity of the catalysts. Generally, there were two categories of modifications of OMS-2; firstly, metals substituted and H-exchanged, aimed to improve the catalytic oxidation of OMS-2 catalyst. The catalysts were characterized by several characterization methods to find out their physicochemical properties. Thus, those were correlated to the catalytic activity of the catalysts in oxidation of cyclohexane. However, this study emphasized on titanium incorporated OMS-2. The effect of synthesis method, amount and location of the titanium substituted to the physicochemical properties of OMS-2 were evaluated. Characterizations were done by several methods such as structural and acidity analysis, surface area, hydrophobicity-hydrophylicity, etc. Thus, their physicochemical properties were correlated to catalytic activity of the catalyst in several oxidation reactions such as cyclohexane, cyclohexene and styrene. The second modification was sulphation of titanium supported OMS-2 aimed to create acid sites on the oxidative catalyst. The catalysts were characterized and tested in consecutive oxidation and acid reactions for direct synthesis of diols from alkene. Briefly, the research attempted to answer some questions as depicted in Figure 1.7. 1.3 Research Objectives The research proposed is a fundamental study of heterogeneous catalytic system for both oxidation process and consecutive reaction of alkenes to alcohols under mild conditions. Since it is a one step, inexpensive and easy preparation of the 15 Synthesis x OMS-2 synthesized in different methods x Synthesis of metals substituted OMS-2 (M = Ti, Fe, Cu, Co) x H-exchanged OMS-2 x Sulfation of Ti-OMS-2 Characterization x x x x Structural analysis Acidity analysis Surface area Hydrophobicityhydrophilicity Research Questions Catalytic activity x Oxidation reactions x Consecutive oxidation and acid reactions 1. Which method gives the best performance of OMS-2 in oxidation of benzyl alcohol. 2. What are the effects of metals substituted and H-exchanged to the physicochemical properties of OMS-2 and its correlation to catalytic activity of the catalyst in oxidation reaction? 3. How are the catalytic activities of Ti-OMS-2 compared to the other metals incorporated OMS-2? 4. What are the effects of the synthesis method, amount and location of titanium sites to physicochemical properties of Ti-OMS-2 and its correlation to catalytic activity in oxidation reaction? 5. How is catalytic activity of Ti sites in different substrates in oxidation reaction? 6. Is sulfation of Ti-OMS-2 catalyst successfully designed as a bifunctional oxidative and acidic catalyst for synthesis of diols from alkene? Figure 1.7: The schematic of the research approach and research questions. 16 catalyst using mild oxidant as oxidizing agent, these systems are more economical and environmentally friendly (see Figure 1.8). In light of the issues described previously, the research was conducted with the following objectives: 1. To synthesize manganese oxide octahedral molecular sieve (OMS-2), titanium supported OMS-2 (Ti-OMS-2), metal supported OMS-2 (MOMS-2) and H doped metal supported OMS-2 (H-M-OMS-2). 2. To modify the surface of Ti-OMS-2 by sulphation as a bifunctional catalyst. 3. To study the physicochemical properties-catalytic activity relationship in model reactions. 17 Multidisciplinary Science Catalysis x x x x Oxidation reaction Importance Success of the industry in 20th century Greener industry in 21th century The heart of quality of life A healthy economy growth Biggest problem areas x OMS-2 as a promising heterogeneous catalyst (active with mild oxidant, inexpensive and easy preparation) x TS-1 as a milestone in oxidation catalysis Acid-catalyze reaction Sulfated metal oxide as a strong solid acid catalyst which more environmentally friendly Catalyst Design Catalyst Design Sulfated Ti-OMS-2 Ti-OMS-2 A new heterogeneous catalyst for oxidation reaction under mild condition Bifunctional oxidative and acidic catalyst for consecutive reaction of alkene to diols (one pot reaction) Significant More economic and environmentally friendly catalysts Figure 1.8: The significant of the use Ti-OMS-2 in heterogeneous oxidation and acid reactions. CHAPTER 2 LITERATURE REVIEW 2.1 Introduction This chapter presents brief introduction of classification porous manganese oxide, synthesis of porous manganese oxide and OMS-2 materials, an overview of modification and catalytic activity of OMS-2 materials, oxidation reaction, titanium incorported materials, sulphated metal oxides as solid acid and synthesis of diols. The modification and characterization techniques and reactions which used as model in this research are described. 2.2 Classification of Porous Manganese Oxide The tunnel and layered manganese oxide minerals can generally be classified into three groups i.e. the pyrolusite-ramsdellite family with (1×n) tunnel structure, the hollandite-romanechite family with (2×n) tunnel structure, and the todorokite family with (3×n) tunnel structure, respectively. All structures contain infinite chains of edge-sharing MnO6 octahedral structural units, and the numbers 1, 2, 3, and n correspond to the number of octahedra in the unit chain width. The chains are linked by corner sharing to form a one dimensional tunnel structural network. This structure comprises of infinite 3-D crystalline frameworks with molecule-sized tunnels similar to those found in zeolites. When n = , the network corresponds to a 19 layered structure which is similar to clay-type materials. Table 2.1 shows the nomenclature of major tunnel and layered manganese oxides proposed by Turner and Buseck [13]. Table 2.1: Classification of tunnel and layered manganese oxides and their crystallographic data [13]. Mineral or compound Approximate formula Crystal system (space group) Lattice constants (Å) Structural features (1×n) family Pyrolusite (-MnO2) MnO2 Tetragonal (P42/mn2) a=4.39; c=2.87 (1×1) tunnel Ramsdellite MnO2 Orthorhombic a=4.53; (Pbnm) b=9.27; c=2.87 (1×2) tunnel Nsutite (MnO2) [MnII, MnIII, MnIV] (O,OH)2 Hexagonal a=9.65; c=4.43 (1×1)/(1×2) complex tunnel LiMn2O4 Cubic (Fd3m) a=8.25 (1×3) tunnel Li1.09Mn0.91O2 Hexagonal (R3m) a=2.85; c=14.26 (1×) layer of 4.75 Å wide Vernadite ( MnO2) MnO2 ·H2O·R2O,RO, R2O3 (R=Na,Ca, Co, Fe, Mn) Hexagonal a=2.86; c=4.7 (1×) layer of 4.75 Å wide (2×n) family Hollandite (R)2[Mn8]O16·xH2O (R=Ba, K, Na, NH4) Tetragonal (I4/m) or monoclinic (I2/m) a=9.96; c=2.86 a=10.03; b=5.76; c=9.90; =90.42 (2×2) tunnel Continued in Page 20 20 Romanechite or psilomelane (R)2[Mn5]O10·xH2O Monoclinic (A2/m) a=9.84; b=2.88; c=13.85 =92.30 (2×3) tunnel RUB-7 (R=Ba, K, Na) (R)2[Mn6]O12·xH2O Monoclinic (C2/m) a=14.19; b=2.85; c=24.34; =91.29 (2×4) tunnel Rb0.27MnO2 (R=Rb, K, Na) (Rb)4[Mn7]O14 Monoclinic (A2/m) a=15.04; b=2.89; c=14.64; =92.4 (2×5) tunnel Birnessite Na4Mn14O27·9H2O and RyMnO2·xH2O (R=monovalent or divalent metal ions) Orthorhombic a=8.54; b=15.39; Hexagonal, c=14.26 monoclinic, orthorhombic, or triclinic (3×n) family Todorokite ( R)[Mn6 ]O18·xH2O (R=divalent metal ions and Na, K) Monoclinic (P2/m) a=9.76; b=2.84; c=9.55; =94.1 (3×3 ) tunnel Buserite NayMnO2·xH2O and RyMnO2·xH2O (R=divalent metal ions) Hexagonal a=8.41; c=10.01 (3×) layer of ~10 Å wide Lithiophorite [MnIII2MnIV4O12] [Li2Al4(OH)12] Monoclinic (C2/m) a=5.06; b=8.70; c=9.61; =100.7 Sandwich layer of ~9.5 Å apart Na0.44MnO2 Orthorhombic a=9.10; (Pbam) b=26.34; c=2.82 MnO6/MnO5 complex tunnel Li0.44MnO2 Orthorhombic a=8.93; (Pbam) b=24.44; c=2.83 MnO6/MnO5 complex tunnel (2×) layer of ~7 Å wide Other families Continued in Page 21 21 Ba6Mn24O48 Tetragonal (I4/m) a=18.17; c=2.824 (1×1)×(2×2) complex tunnel -NaMnO2 Monoclinic (C2/m) a=5.63; b=2.86; c=5.77; =112.9 Layer 5.3 Å wide -LiMnO2 Monoclinic (C2/m) a=5.44; b=2.81; c=5.39; =116.0 Layer 4.8 Å wide -NaMnO2 a=2.85; b=6.31; Orthorhombic c=4.77 (Pmnm) Layer 6.31 Å wide -LiMnO2 Orthorhombic a=2.81; (Pmnm) b=5.75; c=4.57 Layer 5.75 Å wide Mesoporous OMS-1 (MOMS-1) Mn2O3/Mn3O4 Hexagonal a=54; c= Mesopore size about 30 Å Mesoporous OMS-2 (MOMS-2) Mn2O3/Mn3O4 Cubic (Ia3d ) -a Mesopore -a not reported The schematic structures of manganese oxide with one-dimensional tunnel and layered structures are shown in Figure 2.1. Pyrolusite-, ramsdellite-, hollandite-, romanechite- and todorokite-type manganese oxides, and synthetic Rb0.27MnO2 have one-dimensional (1×1), (1×2), (2×2), (2×3), (3×3), and (2×5) tunnel structures, respectively. On the other hand, birnessite- and buserite-type manganese oxides have layered structures with basal spacings of about 7 and 10 Å, respectively. The tunnels and the interlayer spaces of the manganese oxides materials can be occupied by metal ions and crystal water molecules. The birnessite and buserite structures contain a single crystal water sheet and a double crystal water sheet between the MnO6 octahedral sheets, respectively. Most Mn are tetravalent in these 22 manganese oxides, but a part of Mn are trivalent in order to balance the charge of the foreign ions in the tunnels and interlayers. MnO6 octahedron a. (1x1) tunnel pyrolusite d. (2x2) tunnel hollandite g. (2x5) tunnel Rb0.27MnO2 b. (1x2) tunnel ramsdellite e. (2x3) tunnel romanechite h. (2x) layer birnessite c. (1x) layer Li1.09Mn0.91O2 f. (2x4) tunnel Rub-7 i. (3x3) tunnel todorokite j. (3x) layer buserite Figure 2.1: Schematic structures of one-dimensional tunnel and layered manganese oxides [13]. 23 The intergrowth of two or more tunnel phases occurs also in the manganese oxides as shown in Figure 2.2 [13]. An irregular intergrowth of (1×1) tunnels (pyrolusite) and (1×2) tunnels (ramsdellite) in the structure of -MnO2 (nsutite) is well known to electrochemists. Complex intergrowths of the (2×2) tunnels (hollandite) and the (2×3) tunnel (romanechite) are found in fibrous manganese oxide minerals. Almost all the intergrowths are random, so that regular periodicity or superstructure may not be apparent. a. (1x1)/(1x2) tunnel intergrowth (nsutite) b. (2x2)/(2x3) tunnel intergrowth Figure 2.2: Intergrowth tunnels of (a) (1×1) and (1×2), and (b) (2×2) and (2×3) in the tunnel structure of manganese oxides. The naturally occurring manganese oxides, e.g. manganese nodules are found on the sea floor at depths of thousands of meters and occur as vast deposits, which are estimated to be as much as 500 billion tons in all oceans [56]. Manganese nodules are generally composed of very small particles of poorly crystalline manganese (IV) oxide minerals, which are related to terrestrial minerals, such as binnersite, todorokite, cryptomelane, and nsutite. The compositions of manganese nodules are not homogeneous and vary considerably from place to place on the ocean floor. The occurrence of different mineral phases in manganese nodules depends on the location, marine environment, and sediment type. Compositional variations occur even among samples collected at the same place. Manganese nodules consist of a variety of metal oxides, mainly transition metal oxides. The major elements in manganese nodules are Mn and Fe (~50 wt% each, excluding O), and the minor elements (less than 2 wt%) are Co, Ni, Cu, Ti, Si, Al, Ca, and Mg. In addition to being potentially important mineral sources for useful metals due to the vast deposits and varied compositions of metal oxides, manganese nodules are also valuable as potential cheap natural sources of adsorbents and catalyst due to 24 the porous structure (porosity in the order of 60%) and large specific areas (90 to 400 m2/g) [56]. Several promising processes utilizing manganese nodules have already been proposed and applied in the petroleum industry and pollution control plants. Manganese nodules are described as powerful catalysts in oxidation of carbon monoxide, methane and butane, the reduction of nitric oxide with ammonia and demetallation of topped crude in the presence of hydrogen [15]. Although some catalytic studies of manganese nodules have been conducted, there are a few fundamental and systematic studies on their catalytic properties due to the complexity and uncertainty of their chemical and mineralogical composition. Suib and co-workers [11] prepared the synthetic manganese oxide materials with octahedral molecular sieves and octahedral layer structures. They claimed that these materials are pure and more crystalline, and have more uniform and homogeneous structure than their natural counterparts. OMS materials normally have an open framework tunnel structure with sizes ranging from a few angstroms to hundreds of angstroms. This unique property makes them excellent shape selective catalysts and ion exchange materials, which make the application of OMS materials more consistent and reproducible results [57, 20]. 2.3 Material Synthesis 2.2.1 Synthesis of Porous Manganese Oxide Porous manganese oxides, as well as layered claylike manganese oxide materials are prepared via a variety of routes. Many of the materials with similar gross structural features nevertheless show a diversity of properties depending on the specific synthetic route. These differences may be attributed to variations in particle size and the type and amount of defects in the structures. Thus, small changes in synthetic parameters can result in materials with novel catalytic, electrochemical, and ion-exchange properties [12]. 25 The tunnel and layered manganese oxides can be prepared by a variety of processes [12, 13]. These processes are classified into: 1. Dry process solid state reaction melting salt flux processes 2. Wet process redox precipitation / reflux method hydrothermal hydrothermal soft chemical processes 3. Wet–dry process sol–gel processes. Stable phases with small tunnel and narrow layered structures, such as spinel, hollandite, romanechite, and birnessite can be synthesized by any of these processes. However, metastable phases with large tunnel and wide layered structures, such as Rb0.27MnO2, RUB-7, todorokite, buserite, and mesoporous manganese oxides, can only be prepared by the wet processes [13]. Metal ions and organic surfactants are usually used as templates (see Table 2.2). The wide tunnel and layered structures are easily formed under high template concentrations. Generally, the ion-sieve, electrochemical, and catalytic properties of the manganese oxides depend on the synthetic process. When Ba2+ is used as template for 2x2 tunnel structure of manganese oxide the materials are called hollandite. If the template is K+, Na+, Pb2+, the related materials are named cryptomelane, manjiroite and coronadite, respectively. Generally, all those materials are grouped in hollandite type materials. The synthetic layered and tunnel manganese oxide materials are also known as octahedral layer (OL) structure materials and octahedral molecular sieves (OMS) materials, respectively [19]. Thus, the synthetic binnersite, hollandite and todorokite are labelled as OL-1, OMS-2 and OMS-1 material, respectively. In this work, we emphasised on OMS-2 (holandite type) materials due to their advantages over other manganese oxide materials in catalytic oxidation. 26 Table 2.2: Synthesis of tunnel and layered manganese oxides with various templates [13]. Template Compound Wet process Dry or wet–dry process Pyrolusite (-MnO2) H+ No template Ramsdellite H+ -a Spinel Li+ Li+, Mg2+ A0.44MnO2 -a Na+ -AMnO2 -a Na+ -AMnO2 -a Li+, Na+ Hollandite K+, Rb+ , Pb2+, Ba2+, NH4+, H3 O+ K+, Rb+ , Pb2+, Ba2+ Romanechite Ba2+ Na+ RUB-7 Rb+, K+, Na+ -a Rb0.27MnO2 Rb+ -a Birnessite Alkali metal ions Na+, K+ Ba6Mn24O48 -a Ba+ Buserita Na+ -a Todorokite Mg+, divalent transition metal ion -a MOMS (Mesoporous material) CH3(CH2)15(CH3)3N+, [CH3(CH2)3]4N+ -a - a Not reported 2.2.2 Synthesis of OMS-2 Materials In the solid state reaction, OMS-2 may be synthesized by calcinations of MnO2 at 600 oC for 2 h using Na+ or K+ as template [58]. However, using Na+ as a template for the hollandite is rare, because the size of Na+ is somewhat smaller than the tunnel sizes of hollandite types. 27 The precipitation process or also known as reflux method involving oxidation of Mn salts and/or reduction of MnO4 salts in solutions is a typical method for the 2+ production of birnessite and hollandite. Birnessite can be obtained in alkaline or weak acidic solution, while the hollandite can be obtained only in acidic solution. The hollandite-type manganese oxides may be prepared directly by oxidation of Mn2+ salts or reduction of AMnO4 (A = alkali metal) in acidic aqueous solution. The oxidants can be used for synthesis of hollandite type material by oxidation of Mn2+ salts are KMnO4, KClO3, O2, H2O2 and K2S2O8 [30, 37, 59-62]. K+ and NH4+ are usually used as the template for the preparation of hollandites. K-hollandites are prepared by reacting MnSO4 or Mn(NO3)2 solution with KMnO4 or KClO3 in a 1 M H2SO4 solution at above 60 oC. NH4-hollandites are obtained by using NH4S2O8 as the oxidant. The hollandites without template metal ions in the tunnel sites could be directly prepared by reacting Mn(NO3)2 with LiMnO4 or NaMnO4, or oxidation of MnSO4 with ozone (O3) gas in a H2SO4 solution. In these reactions, H3O+ may act as the template for the hollandite structure. A higher reaction temperature (>70 °C) and higher concentration of H2SO4 (>4 M) are necessary for the hollandite formation with small diameter of template metal than when using a K+ or NH4+ template which has bigger diameter [32]. Alternatively, hollandite is formed by hydrothermal decomposition of KMnO4 at higher temperature and lower pressure as compared to the conditions for birnessite formation. A K2Mn4O9 phase is obtained in this hydrothermal decomposition reaction. Ohzuku et al. [63] prepared hollandites with NH4+, K+, and Rb+ templates by treating Mn2O3 in 0.5 M H2SO4 solutions containing these ions at 100 °C. The hollandite without the template ions in the tunnel can be obtained by leaching Mn2O3 in a 4-8 M H2SO4 solution at 105 °C. The treatment of Mn2O3 in acidic solution results in the disproportionation of Mn2O3 into soluble Mn2+ and MnO2 with a hollandite structure. Rossouw et al. [64] have synthesized a highly crystalline hollandite without metal ion or NH4+ in the tunnel by leaching Li2MnO3 in a 1-4 M H2SO4 solution at 90 °C. Similar to the redox precipitation reaction, the H3O+ ions may act as the template in these reactions. Hollandite is also formed in the hydrothermal reacting -MnO2 in 5 M RbOH at >400 °C [65]. The structure of the manganese oxide strongly depends on the concentration of RbOH. The product 28 is hollandite of low RbOH concentration, while the product is (2×5) manganese oxide or birnessite-type manganese oxide at high RbOH concentration. This suggests that large tunnel and layered structures are preferred in alkaline solution at high template concentrations. A hydrothermal soft chemical process is a useful and unique method for the preparation of the tunnel manganese oxides. This process comprises of two steps: the first step is the preparation of a framework precursor with layered or analogous structure and insertion of template ions or molecules into the interlayer space by a soft chemical reaction, and the second step is the transformation of the templateinserted precursor into a tunnel or other structure by hydrothermal treatment. The dimension of the resulting tunnel can be designed and predicted easily from the dimensions of the template. Hollandite can be obtained by using K+ templates. The schematic procedure is shown in Figure 2.3. Layered structure Layered structure Tunnel structure K+ Hydrothermal reaction Ion-exchange reaction Na-binnersite K-binnersite Hollandite Figure 2.3: Transformation reaction from birnessite to hollandite under hydrothermal conditions. The hydrothermal reaction, however, may be most effective for the transformation reaction, because under hydrothermal conditions the transformation reaction can be carried out at low temperature to prevent destruction of metastable tunnel structures, and the template ion or molecule can be incorporated/liberated into/from the precursor from/to the solution phase to reach a composition for the desired structure during the transformation reaction. 29 The sol–gel synthesis hollandite has been prepared by the reaction of KMnO4 with organic reducing agent [19, 33]. Two different groups of organic species were investigated as reducing agents: multifunctional carboxylic acids, such as fumaric acid and maleic acids, and polyols, including sugars. The authors have investigated the effect of cations (Na, K), ratio of reducing agent to permanganate, temperature of calcinations, and effect of pH on the structure of the materials produced, along with their ion-exchange properties. In other ways, OMS materials were also prepared by microwave heating method [66, 67]. The microwave heating method gave different catalytic activity to the thermal heating. The comparison of preparation of two sets of materials, using the same starting reagents and conditions, but one heated by microwaves and the other thermally, showed the different catalytic activity in oxidation of benzene [20]. The samples prepared in microwave heating are more active for total oxidation. Among the preparation method used in the synthesis of OMS-2 material, a reflux method was one route mostly used to prepare bulk OMS-2 materials [60, 62]. Study on the effect of different media i.e. acidic, neutral and basic condition in the synthesis of OMS-2 by Maknawa et al. [23] showed that OMS-2 prepared in acidic medium resulted in material with the highest surface area and maximum activity in oxidation of alcohol. 2.4 Modification of Manganese Oxide 2.4.1 Overview of Modification of OMS-2 materials Modification of OMS-2 materials was aimed at achieving better electronic and catalytic performance of OMS-2. Some authors reported the incorporation of metal cations into OMS-2 materials as summarized in Table 2.3. It shows that metal incorporated OMS-2 alters the structures and properties of OMS-2 materials. Other metal cations can substitute potassium ion in the tunnel or manganese ions in the Cu2+ A sol-gel and Framework reflux method -a Adding Mg2+, Ni2+, Cu2+, dopant prior to refluxing Co2+, and Fe3+ Location -a Synthesis route Ni2+, Cu2+, Adding Fe3+, Co2+, dopant prior to refluxing and Zn2+ Metal/Ion Doping Thermal stability and resistivity of Cu-OMS-2 decrease and increase, respectively, as amount Cu incorporated increase. -a -a Properties Liquid phase Gas phase Type reaction Gas oxidative phase dehydrogenation of ethylbenzene to styrene Decomposition of H2O2 Oxidative dehydrogenation of ethanol Catalytic Application [68] [45] [46] References Continued in Page 31 Copper content influenced styrene selectivity. The trends is Ni- > Cu- > Fe> Co- > Mg-OMS-2 to decompose H2O2. Fe-OMS-2 had the highest conversion and selectivity to acetaldehyde Results Table 2.3 : Review of Metal Doped into OMS-2; its synthesis routes, location, properties and catalytic application. 30 Reflux method Hydrothermal method Hydrothermal method Reflux method Ag+, Co2+, Cu2+, and Fe3+ Fe3+ Cr3+ Cu2+, Zn2+, Ni2+, Co2+, Al3+, or Mg2+ Metal cations probably in framework except Cu2+ (mostly in the tunnel) Tunnel and framework Framework Framework All M-OMS-2 materials have similar acidities but varied strength and amount of basic sites. M-OMS-2 has slightly higher surface areas and pore volumes M-OMS-2 has similar thermal stabilities but slightly lower than undoped OMS-2. -a The thermal stability of cryptomelane was improved by iron doping. The strength of basic sites increased with increasing amounts of iron doping. Doped OMS-2 catalysts are very stable over long times on stream during reaction. -a -a -a Oxidation of carbon monoxide -a -a -a Gas phase [30] [37] [36] Continued in Page 32 Phiscochemical properties of M-OMS-2 Synthesis of Cr-OMS-2, from Na-binnersite and ion exchange with Cr. Fe was in the framework structure of OMS-2 Catalytic activity [31] depended on oxidation number of Mn and the position and nature of the doped cations. K–OMS2 shows negligible activity. 31 Ion-exchange Hydrothermal Tunnel method H+ Li+, Na+, K+, Rb+, and H+ Tunnel Tunnel Ion-exchange H+ -a Reflux method Cu2+, Zn2+, Ni2+, Co2+, Al3+, or Mg2+ The larger cations led to better crystallinity and more ordered tunnel structure. Larger cations lead to higher thermal stability. A-OMS-2 had nanofibrous morphologies with various widths on the order of nm and various lengths on the order of μm. The exchange of the tunnel cation with the smaller H+ ions leads to weakening of the Mn–O bond The catalyst has Lewis acid sites -a Aerobic oxidation of cyclohexanol to cyclohexanone Oxidation of alcohol Oxidation of benzyl alcohol with O2 Decomposition of 2-Propanol, product: acetone and propene. Liquid phase Liquid phase Liquid phase Gas phase [32] [23] [24] [69] Contiuned in Page 33 Higher surface areas lead to better catalytic performance. Selectivity 100% to corresponding aldehyde/keton, the highest conversion for oxidation benzylic compound. High activity and 100% selectivity to benzaldehyde and proven via Mars van Krevelen mechanism. The highest: Cu-OMS-2 due to phase transition during reaction and hausmannite phase responsible to the highest activity which is not observed on others metal doped. 32 Hydrothermal method Sol-gel assisted solidstate method Hydrothermal method V5+ and Nb5+ Fe3+, Co2+, Ni2+, and Cu2+ Cu2+ Framework Framework and tunnel Framework -a The thermal heating of nanomaterials were as high as 800 °C. Fe(III) cations were most suitable to be associated into the structure of nanoscale OMS-2. M-OMS-2 lowers the manganese oxide average oxidation state and increases its electrical resistivity. -a Oxidation of toluene to produce benzyl alcohol, benzylaldehyde, and benzoic acid -a -a Liquid phase -a [43] [41] [42] Continued in Page 34 XAS (XANES and EXAFS) showed the presence of octahedral Mn4+ and that Cu2+ ion were supposed to interact with lattice oxygen and manganese atoms which consistent with copper being sited within the channel of the OMS-2 framework. Compared with onventional OMS-2 catalysts, the nanoscale OMS-2 showed exceptional catalytic activity in the green oxidation of toluene. Physicochemical properties M-OMS-2 33 Ion exchange Ion exchange and impregnation. Reflux, hydrothermal, ion-exchange and impregnation method. Soft chemical method H+ Ce3+ Ce3+ Rb+ Tunnel Framework, tunnel, and extraframework Tunnel and nonframework Tunnel Synthesized OMS 3D architectures with tunable tunnel structures. The thermal stability of the material varies depending on preparation method. -a Wet oxidation of phenol compounds -a Liquid phase Gas phase Liquid Acid-catalyzed phase condensation of phenylhydroxylamin e with aniline to produce 2aminodiphenylamine. BET surface area of Ce- Oxidation and dehydrogenation of OMS-2 is higher than cyclohexanone OMS-2. The average oxidation state (AOS) of Mn decreases with an increasing extent of K+ exchange by H+ Successive exchanges with HNO3 resulted in better H+-exchanged material. [72] [71] [34] [70] Contiuned in Page 35 Control of crystal forms, morphologies, and tunnel sizes. Best performance: well crystallized OMS-2 where all the accessible potassium ions were exchanged for cerium cations. The selective product formation depended on the nature of the acidic and basic properties of the catalyst. H-K-OMS-2 with the highest %H+ gave the highest selectivity for the ortho isomer. High selectivity to ortho isomer compared with para isomer. 34 Reflux, hydrothermal, ion-exchange and impregnation methods Zr4+ -a Not reported framework Hydrothermal method Cu2+ Framework, tunnel, and extraframework -a Fe3+, Ni2+, Reflux, Cu2+, and Hydrothermal, Solvent free Co2+ technique Sol-gel and combustion method. The morphology of Zr-OMS-2 materials depends on the preparation method. Synthesis of CuOMS-2, from Kbinnersite using potassium permanganate and potassium persulphate and ion exchange with Cu -a Oxidation of side chain ethylbenzene, benzyl alcohol and cyclohexanol. -a Oxidation of styrene Liquid phase -a Liquid phase The Zr-K-OMS-2 synthesized by the impregnation method showed efficient conversion for oxidation of ethylbenzene, benzyl alcohol and cyclohexanol. Cu-OMS-2-permanganate material shows fibrous and needle shape particle morphology, whereas CuOMS-2-persulphate shows only the globular type aggregated particle morphology. Transition metal doped alters the acidity of OMS-2, leading to higher selectivity to styrene oxide. [39] [38] [27] 35 36 framework structure of OMS-2 materials. The physical and chemical properties of doping ions, and the properties of doped materials are significantly different from doped OMS-2 are greatly influenced by the type, amount, and the location of the those of the undoped ones [12]. Different metals doped OMS-2 materials have catalytic properties in different reactions and different amounts of doping affect the catalytic abilities of the materials and different doping locations such as framework or tunnels can provide different active sites for catalytic reaction [31, 32, 45, 46]. Table 2.3 also shows that the position of ion doping depended on the synthesis method. Substitution of manganese in framework sites of OMS-2 materials by other metal ions is possible by doping the initial solution with doping ions [69] or isothermal/framework substitutions [28, 36]. Ion exchange of OMS-2 via post synthesis treatment could only substitute the potassium in the tunnel structure while impregnation could create agglomeration of the non-framework metal oxide on surface OMS-2 particle. The catalytic activities of metal doped OMS-2 materials are varied, depending on the physicochemical properties of metal doped materials. The materials are active for oxidation, dehydrogenation and acid catalyze condensation. Extensive research is needed to explore the potential applications of metal doped OMS-2 in the catalytic oxidation reaction. Heterogeneously catalyzed partial oxidation of organic compound is widely applied in numerous chemical, biological and pharmaceutical industries. Products of selective oxidation of olefins are important starting materials towards the production of many other fine chemicals [73]. The previous section shows that the modifications on OMS-2 materials mainly are based on doped of metals (alkali, alkali earth and transition of metal). In order to enhance the physicochemical properties and catalytic activity of OMS-2 materials some modification was used. The following sub-section describes several modification techniques used in this research. 37 2.4.2 Ion Exchange Ion exchange is a natural phenomenon occurring continually in inorganic substances and in living bodies on the earth’s surface. The ion exchange reaction can be described as the interchange of ions between a solid phase and a liquid surrounding the solid. Initially, ion exchange was confined to surface reactions, but these were gradually replaced by gel type structures where the exchanged sites were available throughout the particle. The process is shown graphically in Figure 2.4: Na+ Na+ + Ca++ Ca++ + Na+ Figure 2.4: The ion exchange reaction. The sites exhibit affinity for certain ions over others and this phenomenon is very helpful in removing objectionable ionic materials from process streams. The affinity relationship can also be expressed by equilibrium (selectivity) equations based on the reversibility of ion exchange reactions and the law of mass action. K+ ions in OMS-2 can be ion-exchanged by treatment with HNO3. All alkali metal ions can enter the lattice of K+-removed OMS-2. The K+ extraction and metal ion adsorption reactions are topotactic, preserving the hollandite structure. Feng et al. [74] have proposed that hollandite-type manganese oxides can be expressed by a general formula {An}[xMn8x]O16 (n2, x1), where {}, [], and A denote the (2×2) tunnel sites, octahedral sites for Mn, octahedral vacant sites, and metal ions in the tunnel, respectively. The hollandite-type manganese oxides can be classified also as redox-type and ion-exchange-type similar to the spinel-type manganese oxides. {K2}[Mn3+2Mn4+6]O16 is a redox-type OMS-2, and {K2}[0.5Mn4+7.5]O16 is an ion exchange-type OMS-2. The redox and ion-exchange extraction/insertion reactions for OMS-2 can be written as follows: 38 8{K2}[Mn3+2Mn4+6]O16 + 32H+ 7{ }[Mn4+8]O16 + 16K+ + 8Mn2+ + 16H2O (Equation 2.1) 7{ }[Mn4+8]O16 + 2KOH {K2}[Mn3+2Mn4+6]O16 + H2O + (1/2)O2 (Equation 2.2) while for the redox reactions; {K2}[0.5Mn4+7.5]O16 + 2H+ {H2}[0.5Mn4+7.5]O16 + 2K+ (Equation 2.3) for the ion-exchange reactions. The redox-type extraction of one K+ is attended by the disproportionation of one Mn3+ to 0.5Mn2+ and 0.5Mn4+. The number of ionexchange-type sites is four times the number of manganese defects at the octahedral site, similar to the spinel system. 2.4.3 Impregnation Impregnation as a means of supported catalyst preparation is achieved by filling the pores of a support with a solution of the metal salt from which the solvent is subsequently evaporated (See Figure 2.5). The catalyst is prepared either by spraying the support with a solution of the metal compound or by adding the support material to a solution of a suitable metal salt, so that the required amount of active component is incorporated into the support without the use of excess solution. This Figure 2.5: Impregnation of porous catalyst. 39 is then followed by drying and subsequent decomposition of the metal salt at an elevated temperature, either by thermal decomposition or reduction. Impregnation is an extremely versatile technique (although it is not applicable to insoluble reagents) which can be controlled to give good dispersion and a known loading of reagent. 2.4.4 Alkylsilylation Alkylsilylation is modification of external surface of the catalyst by attachment of organofunctional silane or organosilane. Octadecyltrichlorosilane (OTS) which has a chemical formula of C18H37SiCl3 is a type of family of organosilane. The alkyl groups of OTS possess a hydrophobic characteristic as depicted in Figure 2.6. It is widely used for surface modification and functionalization. OTS is preferred over other organic compounds since the silanes can form bonds via several mechanisms [75]. Organosilane deposition has additional benefits over other preparation methods because of its fast preparation, stable finish and applicability to a wide range of substrates. Polar head Cl Nonpolar tail Si Cl Cl Figure 2.6: Chemical structure of OTS. Nur et al. [76] demonstrated that amphiphilic Ti-loaded NaY zeolite prepared by partial modification with OTS successfully catalyzed epoxidation of 1octene in this system. In this catalytic system, part of the external surface of which was covered with hydrophobic alkyl groups and the rest being left hydrophilic. The amphiphilic catalyst exhibited much higher catalytic activity than that of hydrophilic titanium-loaded NaY, without modification by OTS, or of a hydrophobic catalyst with almost full coverage by the alkyl groups [76, 77]. 40 The assembly or coating occurs by hydrolysis reaction of free surface hydroxyl group on the support and the organosilanes. Figures 2.7 and 2.8 show the mechanisms of alkylsilane using OTS on the surface of support. In case of complete formation of assembled monolayer using OTS (Figure 2.7), all chlorine species hydrolyze to form bonds with other hydroxyl groups. OTS reacts with the substrate surface through only one hydroxyl group, while two other groups produce uniform surface coverage through cross polymerization. Figure 2.8 represents incomplete mechanism of OTS. In case of incomplete reaction, there are some unreacted chlorosilane groups remaining on the surface which might be hydrolyzed later, especially by trace amount of water in the reaction vassel. This type of reaction favours deposition of multilayer alkylsilane, instead of a monolayer deposition. CH3 Cl H O + n Cl Si (CH2)17 CH3 (CH2)17 - HCl O Cl surface CH3 Si O (CH2)17 Si O O O n OTS surface Figure 2.7: Mechanism of complete reaction of OTS on the support surface to form a well ordered layer. CH3 Cl H O surface + n Cl Si Cl OTS (CH2)17 CH3 - HCl CH3 CH3 CH3 (CH2)17 (CH2)17 (CH2)17 (CH2)17 CH3 (CH2)17 O Si O Si O Si O Si OH OH Si O O HO HO O surface Figure 2.8: Mechanism of incomplete reaction of OTS on the support surface. 41 2.4.5 Sulphation Sulphated metal oxides are usually made by the precipitation-sulfation (PS) method [78]; where the respective metal hydroxide is precipitated from a salt solution by increasing the pH through addition of concentrated ammonium hydroxide; the filtered, washed, and dried solid is reacted with a sulfate source, e.g., sulfuric acid, and then calcined at elevated temperatures (normally above 700 K). The highly acidic (or “superacidic”) system is obtained at a narrow temperature range, allowing a loose, metastable oxide structure to be formed, e.g., the tetragonal phase in the case of ZrO2, with a few percent sulfate residing on its surface. Another apparent way of making sulphated metal oxides is by controlled thermal decomposition of a metal sulfate salt. Arata et al. [79] made sulphated zirconia and titania by decomposing the respective sulfate salts and claimed that the obtained systems were superacidic. The sulfate decomposition (SD) method avoids the complications associated with the PS method (type of starting salt, concentration and pH effects, various factors in the sulfation stage, etc.). As in the PS case, the eventual sulphated metal oxide product is shaped up by the calcination/decomposition temperature. However, in the SD method the sulphated metal oxide is an intermediate in the thermochemical transformation of a sulfate salt to the corresponding oxide, through eliminating the sulfate, essentially as free SO3. Thus, the SD-derived sulphated metal oxide product is created only when most of the sulfate has been removed, at temperatures usually higher than those needed for making sulphated metal oxide of similar sulfate level by the PS method. On the other hand, an SD-derived sulphated metal oxide might be expected to be more ordered and structurally uniform than the parallel sulphated metal oxide obtained by the PS method. 42 2.5 Catalytic Activity 2.5.1 Catalytic Activity of OMS-2 Materials Cryptomelane OMS-2 (K-OMS-2) materials have promising applications as ion-exchange materials, battery materials, chemical sensors, electromagnetic materials, and catalysts [20]. Recent studies have been reported in detail on the catalytic application of OMS-2 materials in several reactions, such as low temperature carbon monoxide oxidation, total oxidation of methanol, acetone and 2propanol, oxidative dehydrogenation of 1-butene, oxidation of ethanol to acetaldehyde, oxidation or dehydration of 2-propanol to acetone or propylene respectively, dehydrogenation of ethylbenzene to styrene, oxidative dehydrogenation of cyclohexane, partial oxidation of cyclohexane, decomposition of H2O2, and oxidation of CO to CO2 for fuel cells. The details of catalytic applications of OMS-2 materials are shown in Table 2.4. Table 2.4: Some catalytic applications of OMS-2 materials reported in the last decade. Catalyst Application Results Author (s) MnO2, M-OMS-1, M-OMS-2 (M=Mg2+, Ni2+, Cu2+, Co2+, and Decomposition of H2O2 The catalytic activity Zhou et al. in trend of: M-OMS- [45] 2>M-OMS-1>MnO2 M-OMS-1 and MOMS-2 (Ni, Cu, Fe, Co, Mg) Oxidative dehydrogenation of ethanol Most better CoOMS-1 OMS-1, OMS-2 and OL-1 Oxidative Dehydrogenation of 1-Butene The materials are not Krishnan et stable during the al. [40] reaction. Fe3+) Zhou et al. [46] Continued in Page 43 43 Cu-OMS-1 and CuOMS-2 Oxidative dehydrogenation of ethylbenzene to styrene Copper content influenced styrene selectivity. Tolentino et al. [68] M-OMS-2 (M=Ag, Co, Cu) Oxidation of carbon monoxide Catalytic activity of doped OMS-2 catalysts toward CO oxidation shows a correlation among average oxidation state of Mn ion and the position and nature of the doped cation. Xia et al. [31] Hydrophobic OMS2, synthesis in buffer solution, OMS-1, and OL-1 Total oxidation of benzene Luo et al. These materials be [26] very active in total oxidation catalysts or selective adsorbents for VOCs M-OMS-2 (Cu, Zn, Ni, Co, Al, or Mg) Decomposition of 2-Propanol to acetone and propene. The highest active is Cu-OMS-2 K-OMS-2 and H-KOMS-2 Oxidation of H-K-OMS-2 > Kbenzyl alcohol with OMS-2, via Mars O2 van Krevelen mechanism A-OMS-2 (A=Li, Na, K, Rb, H) Aerobic oxidation of cyclohexanol to cyclohexanone OMS-2 and H-OMS- Oxidation of 2 alcohol Chen et al. [69] Makwana et al. [24] Higher surface areas lead to better catalytic performance. Liu et al. [32] 100% Selectivity to aldehyde/keton and the highest conversion for oxidation benzylic compound. Makwana et al. [23] Continued in Page 44 44 OMS-2 Ghosh et al., Oxidation of cyclic Cyclooctene [28] selective to olefins and cyclooctane epoxide. benzylic double bonds with tertiarybutyl hydroperoxide (TBHP) as the oxidant. OMS-2 and H-OMS-2 Acid-catalyzed condensation of phenylhydroxylami ne with aniline to produce 2aminodiphenylamine. High selectivity to ortho isomer compared with para isomer. Kumar et al., [70] Nano-OMS-2, Synthesized by reduction of KMnO4 with H2O2. Oxidation of benzyl alcohol and fluorene Lower activity in oxidation of benzyl alcohol and higher to fluorine compared to OMS-2 conventional. Villegas et al. [61] Ce-OMS-2 Oxidation and dehydrogenation of cyclohexanol Catalytic activity of Jothiramamaterials depends on lingam et al, synthetic method. [34] OMS-2, synthesized by low temperature solvent-free method Oxidation of 2OMS-2solvent free > thiophenemethanol, OMS-2reflux furfuryl alcohol and cyclohexanol Ding et al. [80] Microwave synthesis Oxidation of 2of OMS-2 thiophenemethanol The microwave synthesis of OMS-2 more active than OMS-2 synthesized by conventional heating. However, OMS-2 precursor is the most active catalyst. Malinger et al. [81] OMS-2 and MOMS-2 (Fe, Ni, Cu, Co) Transition metal doped alters the acidity of OMS-2, leading to higher selectivity to styrene oxide. Ghosh et al. [27] Oxidation of styrene Continued in Page 45 45 OMS-2 Total oxidation formaldehyde Catalytic activity is closely related to the morphology of the catalysts. Xingfu et al. [82] OL-1 and OMS-2, with different Mn precursor Total oxidation of acetone OMS-2 with manganese carboxylates as precursor has high surface area and were extremely active for acetone total oxidation Frías et al. [83] Zr-OMS-2 Liquid phase oxidation ethylbenzene, benzyl alcohol and cyclohexanol. The activity of catalyst depends on synthesis method and material prepared by impregnation more active than direct synthesis. Jothiramalingam et al. [39] OMS-2 Total oxidation of ethyl acetate OMS-2 is an active ethyl acetate complete oxidation catalyst with 100% selectivity to CO2 at about 673 K. Gandhe et al. [84] Ce-OMS-2 Wet oxidation of phenol compounds OMS-2 materials containing pure CeO2 phase and excess active xygen species has lower catalytic activity. -Wolfovich et al. [71] 2.5.2 Catalytic Oxidation As known, the scope of partial oxidation catalysis is wide, ranging from the large-scale production of commodities to the synthesis of minute amounts of pharmaceuticals and fine chemicals. Compared with other chemical processes, the oxidation process is complex and difficult to be controlled or stopped at a certain 46 stage. For these reasons, the selective catalytic oxidation is an active field of research. The driving forces for the industrial and academic research are: (i) the formulation of alternative or new catalysts, (ii) reduction of the number of process steps, (iii) elimination of waste by-products and (iv) development of new processes. Catalytic oxidation is an important method of transforming hydrocarbon feedstock like alkanes, alkenes and aromatics into more sophisticated oxygenated products. Million tons of the oxidation product such as alcohol, carbonyl compounds and epoxides are annually produced worldwide [85]. These compounds are used in all areas of chemical industries ranging from pharmaceuticals to large-scale commodities [86-88]. Furthermore, catalytic oxidations can be used to eliminate a series of pollutants, thus producing healthier environment. Traditionally, an oxidation process is conducted by using a stoichiometric process with classic oxidants such as dichromate/sulfuric acid, chromium oxides, permanganates, periodates, and osmium oxide. The process produces large amount of waste. For example, using a stoichiometric amount of potassium dichromate or permanganate results in 5 to 100 times weight of waste (mostly inorganic salts) per kilogram of product [89]. This also means that most of the reactants introduced at the start of the reaction are not converted to the desired product, leading to very low atom-selectivity. Consequently, the use of traditional stoichiometric process is not environmentally friendly and inefficient as well. In contrast, a clean synthetic technology should proceed with a high atom-economy and the overall synthesis must be accomplished with low E-factor (by waste per kg product), thereby minimizing the cost of waste disposal [85, 90]. As a result, currently there is considerable pressure to replace these outdated methods by cleaner catalytically driven technologies. Epoxidation is specific to oxidation of alkene to form epoxide. Traditionally the main methods for performing epoxidations in organic synthesis have been oxidation with organic peracids (See Figure 2.9a) or, to a lesser extent, the chlorohyrins route as shown in Figure 2.9b. The latter method has several drawbacks such as the use of corrosive condition and production of pollutants with a 47 considerable environment impact. However, the use of many organic peracids, e.g. peracetic acid, is problematical owing to restrictions on their transport, storage, and handling. These shortcomings have focused attention on the development of methods using hydrogen peroxide or tert-butyl hydroperoxide (TBHP) as the oxidant (Figure 2.9c). O + R'CO3H + OH base + HOCl (a) R'CO2H Cl O (b) O + R'COOH + R'OH (c) R' = H, t-Bu Figure 2.9: Epoxidation of alkenes using (a) organic peracids, (b) chlorohyrins route and (c) H2O2 or TBHP as the oxidant. Besides that, common products of partial oxidation of alkanes are alcohol and carbonyl compounds. Carbonyl compounds are also formed by the oxidation of alcohol compounds. This catalytic oxidation process of both alkanes and alcohol were studied in homogeneous and heterogeneous catalyst [91-96]. Homogeneous catalysts are single-molecule catalysts, which typically consist of a transition-metal atom, surrounded by one or more ligands. Such well-defined catalysts allow for good mechanistic understanding and consequently, for catalyst performance enhancement by ligand or metal tuning. As a result, homogeneous catalysts are characterized by high activity and selectivity, and a good reproducibility. Additionally, mild operating conditions may be applied, which allows for the production of complex and temperature-sensitive substances. The major drawback of homogeneous catalysis is the recovery of the catalyst from the reaction products. As a result, the industrial applications of homogeneous catalysis are limited, and 48 successful industrial application is mainly due to the high chemo-, regio-, and enantio-selectivities attainable with homogeneous catalysts. In heterogeneous catalysis, the catalytically active substance is deposited on solid support material with a large surface area and a high porosity. Catalyst recovery is relatively simple, although macroscopic diffusion limitations can be a serious problem. Heterogeneous catalysts are robust, with a long service life and often operate at relatively harsh conditions. The disadvantages of homogeneous catalyst are low thermal stability and the difficulty of recovery and regeneration which may be solved with heterogeneous catalyst; which are easily recovered by filtration and recycled [5]. Catalytic oxidations in the liquid phase generally employ soluble metal salts or complexes in combination with clean, inexpensive oxidants such as O2, H2O2, or RO2H [47]. However, heterogeneous catalysts have the advantage, compared to their homogeneous counterparts, of facile recovering and recycling. Moreover, site- isolation of active metal ions or complexes in inorganic matrices precludes their dimerization/oligomerization to less reactive μ-oxo species and therefore endow them with unique activities. Recently, focus has been given to the use of heterogeneous catalysts for the liquid phase epoxidation of olefins with milk oxidant such as alkyl hydroperoxides, hydrogen peroxide or molecular oxygen [97]. The use of heterogeneous catalysts may avoid the difficulties concerning separation, recovery and recycling. As in solution, the concept of “site isolation” should provide very active catalysts [98]. Early transition metal-substituted molecular sieves are of particular interest: (i) they have very high surface areas and a great accessibility to the active sites, and (ii) the high dispersion of the cations is usually obtained by incorporating low amounts of metal directly in the preparation mixture. The comparison of those mild oxidants is demonstrated in Table 2.5. Alkyl hydroperoxide is generally more active than hydrogen peroxide, but it is more expensive and the active oxygen content is rather low. This reaction generates 49 stoichiometric amounts of corresponding alcohols, which in most cases are quite easily recycled via a reaction with hydrogen peroxide. Unfortunately, this process requires at least two extra separations and one extra reaction step. Hydrogen peroxide, with respect to active oxygen content (47%) and the nature of by-product (only water), seems to be the oxidant of choice in catalytic liquid phase oxidations. However, the inherent co-production of water poses some serious difficulties. Most transition metal catalysts are very sensitive to water, and are prone to leaching. In cases where the catalysts are stable, water has serious retarding affect on the oxidation reaction, making the search for new environmentally friendly, effective oxidation catalyst a challenge. Dioxygen which contains 100% active oxygen and generates no by-products is a very attractive choice. Therefore, the development of selective oxidation of hydrocarbon using dioxygen or air in mild condition could be the best option [99, 100]. Table 2.5: The comparisons of the common mild oxidant are used in oxidation process [47, 97]. Alkyl hydroperoxide Hydrogen peroxide Dioxygen Activity More active Active Less active Cost More expensive Expensive Less expensive By-product Corresponding alcohol Water No Active oxygen content Lower than others 100 % 2.5.3 47 % Titanium Incorporated Materials Shell catalyst titanium (IV) silicon dioxide [Ti(IV)SiO2] catalyst and titanium silicalite 1 (TS-1) are titanium based catalysts. Those catalysts were the most successful heterogeneous catalysts in liquid phase epoxidation [47]. Table 2.6 provides a comparison of the characteristics of Shell’s epoxidation catalyst compared with titanium silicalite 1. The Shell catalyst was patented in 1971 by Shell Oil and is 50 industrially used for the epoxidation of propene using an organic peroxide (Figure 2.10). This was the first truly heterogeneous epoxidation catalyst useful for continuous operation in the liquid phase [86]. The catalyst was prepared by impregnating silica with TiCl4 or an organo-titanium compound, followed by calcination. TiIV/SiO2 + O + PhCH(CH3)O2H PhCH(CH3)OH Figure 2.10: Epoxidation of propene on Shell catalyst using an organic peroxide. Table 2.6: Characteristic features of Shell’s epoxidation catalyst compared with titanium silicalite 1. Shell catalyst TS-1 Structure Amorphous, silica based Crystalline microporous structure Ti incooporation Several Ti siloxy sites: from monopodal to tetrapodal Isomorphous replacement of T-atom sites Substrate size No limitations with regard to substrate size Micropores of 5.6 Å diameter impose severe limitations Oxidant Limited to organic peroxides Aqueous hydrogen peroxide The superior catalytic activity of Ti(IV)SiO2 was attributed to both an increase in Lewis acidity of the Ti(IV), owing to electron withdrawal by silanol ligand, and to site isolation of discrete Ti(IV) centers on the silica surface preventing oligomerization to unreactive oxo species (which occurs readily with soluble Ti(IV) compounds). Interestingly the catalyst has been reported to leach catalytically inactive titanium species during the initial stages of the epoxidation reaction after which the catalyst becomes truly heterogeneous. The catalyst was quite unique in that it is heterogeneous, unlike other supported metal oxides such as MoO3 and V2O5 supported on SiO2 or other inert carriers, which often owed their catalytic activity to soluble metal species that may rapidly leach out of the support. 51 TS-1 is an ordered crystalline microporous structure and it has made remarkable progress in the oxidation reactions [50, 101-103]. TS-1 contains Ti(IV) isomorphously substituted for silicon in the framework of silicalite-1, a hydrophobic molecular sieve possessing a three-dimensional system of intersecting elliptical pores with diameters of 5.3×5.5 and 5.1×5.5 Å. The catalyst is especially known for its ability to use aqueous hydrogen peroxide as the oxidant for the epoxidation of small alkenes [104-107]. The other reactions such as oxidation of alcohols, hydroxylation of aromatics, ammoximation of cyclohexanone, oxidation of alkanes to alcohols and ketones, oxidation of amines, oxidation of sulfur containing compounds, and oxidation of ethers, among others, also have been carried out selectively with TS-1. A summary of the reactions catalyzed by TS-1 is given in Figure 2.11 [108]. OH OH OH O OH ArOH OH ArH OH O R C H CH2 TS-1 30% H2O2 R-CH=CH2 OH R C H R-CH2-OH R-CH=O O R' R2NH N OH O R C R' R2NOH Figure 2.11: Oxidation reactions catalyzed by TS-1. Although water is known to seriously retard epoxidation reactions, the hydrophobicity of the pores ensures a very low water content around the catalytic titanium centers, enabling this remarkable feature of the catalyst [91, 109]. The small pore size of the zeolite structure (about 6 Å) allowed shape selective catalysis [110, 111], but on the other hand, restricted the reaction to small alkenes only, which may be seen as a major drawback of the catalyst. To overcome the later problem Ti incorporated in larger porous materials were intensively studied of [103, 112]. 52 2.5.4 Sulphated Metal Oxides as a Solid Acid Catalyst An acid that is stronger than H0 = -12, which corresponds to the acid strength of 100% H2SO4, is known as a superacid [113]. Such a superacidity has been made up by mixing a fluorine containing Brönsted acid (HF, HSO3F, CF3SO3H, etc) and a fluorinated Lewis acid (BF3, SbF5, TaF5, etc.). These superacids have been developed since 1960s and have been applied to various organic syntheses, especially in the field of hydrocarbon chemistry. They are responsible for producing more than 1 x 108 MTon/year of products [114, 115]. However, these conventional industrial acid catalysts have unavoidable drawbacks because of their severe corrosivity and high susceptibility to water. The search for environmentally benign heterogeneous catalysts has driven the worldwide ongoing research of new materials as substitutes for current liquid acids and halogen-based solid acids. Among them sulphated oxides, such as sulphated zirconia, titania, and iron oxide, displaying high thermostability, very strong acidity, and high catalytic activity, have aroused increasing interest. Since the primary work of Arata and Hino [116], the superacids by sulphated metal oxides have been attracting more and more attention for investigations. Sulfation of metal oxides introduces quite strong Brønsted acidity and, in general, enhances the catalytic activity in acid catalyzed reactions. The superacid property of those materials is due to both Lewis acid and Brønsted acid sites [55]. The strong acid properties are related to sulfate ions. It has been proposed that the high electronegativity of sulfur could induce polarisation of the neighboring OH groups. The structure of the sulfate bound to the metal oxide is also a subject of many investigations. However, it is not clear until now. One of the proposed structure is that sulfate is coordinated to the metal in a chelated form, in which two of the oxygen atoms of the sulphate are bound to one metal atom. Brønsted acid is generated by cleavage of one of these bonds by water and formation of OH bond [117]. 53 2.5.5 Synthesis of Diols Preparative procedures in which two or more transformations are carried out as a “one-pot” process offer a number of advantages to the organic chemist: in particular, they result in a reduced number of operations, giving significant time-cost benefits, but they also can often allow “difficult” intermediate compounds (i.e., those that are volatile, toxic, or otherwise noxious) to be prepared and elaborated in situ, thus preventing problems associated with their isolation and handling [51]. To this end, the development of catalyst, which has bifunctional/multifunctional active sites for multiple chemical transformations in a one pot, is considered. Recently, there have been many reports of heterogeneous bifunctional catalysts used for one pot synthesis. Bifunctional palladium/amberlyst catalysts have been used to carry out dehydration of the tertiary alcohol and the hydrogenation of the in situ formed alkene in a single vessel [118]. In another example, caprolactam, a precursor for Nylon-6, which generally requires a two step synthesis, was synthesized in high yields in a single-pot using an aluminophosphate bifunctional heterogeneous catalyst [119]. Diols are important starting materials for polyurethane chemistry which are the important biomedical polymers, and are used in implantable devices such as artificial hearts, cardiovascular catheters, pacemaker lead insulation, etc. [120, 121]. Diols are very important intermediate materials for biologically active natural products such as compactin, macrolides, conduritols, cyclitols and others, as well as statins, an important class of pharmaceuticals [122, 123]. Diols also are useful materials for synthesis of physiologically-compound having specific uses in the field of the drugs and agrichemicals [124]. A number of 1,2-diols such as 2,3-dimethyl2,3-butanediol, 1,2-octanediol, 1,2-hexanediol, 1,2-pentanediol, 1,2- and 2,3butanediol are of interest to fine chemical industries. In addition chiral 1,2-diols are employed as intermediates for pharmaceuticals and agrichemicals [125]. The clean catalytic technology for the synthesis of diols are described in the literature; they can be obtained from the acid sites that catalyzes the hydrolysis of 54 epoxides [126-128], regio- and enantio-selective reduction of diketones using biocatalyst [129], bioconversion of heterocyclic which are linked with penyl or benzyl groups [130]. However, the direct transformation of alkenes which are found in great abundance in the realm of organic molecules to diols using the heterogeneous catalyst is rarely reported. The direct transformation of alkenes to diols can be produced using a bifunctional oxidative and acidic catalyst. The oxidative and Brönsted acid sites of the catalyst acts for oxidation of alkenes to epoxide and to hydrolyze epoxide to diols, respecticely. The heterogeneous catalytic process for one pot reaction of alkenes to diols has been reported by Prasetyoko et al. [52-54]. The catalytic process occurs with the existence of bifunctional oxidative and acidic catalysts. The system is the titanium silicate-1 (TS-1) loaded with sulphated zirconia or niobium oxide. They reported that sulphated zirconia or niobium loaded on TS-1 acted as acid site which has Brönsted acid sites whereas TS-1 itself acted as oxidative sites. The similar properties are also found in Ti-beta and Ti-Al-beta zeolites [131]. The presence of tetrahedral Ti4+ in the framework of zeolite acts as oxidative sites, while the incorporation of Al3+ into Ti-beta induces Brönsted acid sites. Ti-site itself has Lewis acidic character that can coordinate an alcohol molecule to give Brönsted acid. Incorporation of trivalent metal ions (Al3+, B3+, Ga3+ and Fe3+) and titanium ion (Ti4+) together in the framework of silica based molecular sieves display bifunctional oxidative and acid characteristic [132-134]. CHAPTER 3 EXPERIMENTAL 3.1 Synthesis of OMS-2 Materials The integral part of the research is the synthesis of OMS-2 materials followed by modification and characterization of the samples by the characterization techniques described in Chapter 2. The materials were prepared accordingly to the flow chart shown in Figure 3.1. Generally, there are two groups of materials: OMS-2 and modified OMS-2. The experimental procedures are described in the following sections. 3.1.1 Synthesis of OMS-2 without Buffer (OMS-2a) OMS-2 was prepared according to method reported by DeGuzman et al. [62]. A 0.4 M solution of KMnO4 (13.3 g in 225 mL of distilled, deionized water, DDW) was added to a mixture of a 1.75 M solution of MnSO4.H2O (19.8 g in 67.5 mL DDW) and 6.8 mL of concentrated HNO3. The resulting black precipitate was stirred vigorously and refluxed at 100 oC for 24 h. The precipitate was filtered and washed with DDW until neutral pH and dried at 110 oC. The sample was labelled OMS-2a. 56 without OMS-2 buffer with OMS-2a OMS-2b Ti-OMS-2(0.05) without buffer Fe-OMS-2(0.09) Cu-OMS-2(0.02) M-OMS-2 Direct synthesis Co-OMS-2(0.04) Ti-OMS-2(X) without Mn2+ ion Fe-OMS-2 Cu-OMS-2 Materials Co-OMS-2 impregnation Modified OMS-2 TiO2(imp)-OMS-2 physical mixing TiO2-OMS-2(mix) H-OMS-2 H-Ti-OMS-2(Y) Post synthesis ion exchange H-Fe-OMS-2(0.09) H-Cu-OMS-2(0.02) H-Co-OMS-2(0.04) OTS/OMS-2 alkylsilylation OTS/Ti-OMS-2(Z) water sulfation SW100-Ti-OMS-2(Z) SW150-Ti-OMS-2(Z) solvent toluene ST100-Ti-OMS-2(Z) ST150-Ti-OMS-2(Z) Note: X refer to Ti/Mn ratio = 0.18, 0.43 and 0.67; Y refer to Ti/Mn ratio = 0.05 and 0.67 and Z refer to Ti/Mn ratio = 0.67 Figure 3.1: Materials preparation and their labelling. 57 3.1.2 Synthesis of OMS-2 with buffer (OMS-2b) For comparison, OMS-2 was also prepared following the method reported by Luo et al. [26]. A 5.5 g of Mn(Ac)2.4H2O (from Fluka) in 40 mL of DDW was dissolved in a buffer solution consisting of 2.5 mL of glacial acetic acid and 2.5 g of KAc in 20 mL of DDW. A solution of 3.75 g of KMnO4 in 75 mL of DDW was added dropwise to the resulting solution, followed by solution reflux process at 100 o C for 24 h. The resulting brown precipitate was filtered and washed with DDW several times and dried at 120 oC overnight. The final product is labelled as OMS2b. 3.2 Modification of OMS-2 materials The next section explains the modification of OMS-2 materials in this work. The modification was done by direct and post synthesis. Metal substituted OMS-2 (M-OMS-2) was prepared via direct synthesis and the other modification such as ion exchange, impregnation, physical mixing, alkylsilylation and sulfation were prepared via post synthesis. 3.2.1 Synthesis of Metal Substituted OMS-2 (M-OMS-2) M-OMS-2 (M = Ti, Fe, Cu, and Co) was prepared by two methods i.e. following the work by Suib’s group [30] and a new method that is the only metal solution in acidic condition was oxidized by KMnO4 without addition of any manganese ion solution. Different metals were used to study the effect of metal types to properties and catalytic activities of OMS-2 materials. In the first method, metal (M) cations (Ti3+, Fe2+, Co2+ and Cu2+) was added to the acidic manganese ion solution before being oxidized by KMnO4. Metal ions 58 were doped into the OMS-2 structure with an a priori incorporation method, that is, adding aqueous solutions of dopant before refluxing to form the OMS-2 structure. The precursor of Ti3+, Fe2+, Co2+ and Cu2+ cations were Ti2(SO4)3 in H2SO4, FeSO4.7H2O, Co(CH3COO)2 and CuSO4.5H2O, respectively. The precursor was added prior to reflux, with M/Mn atomic ratios of 1:10 except Ti/Mn with ratio of 1:5, under agitation at room temperature. The mixture was oxidized by potassium permanganate and the resulting precipitate was stirred vigorously and refluxed following the procedure in Section 3.1.1. The samples are labeled as M-OMS-2(X), where X is the molar ratio of M/Mn is calculated with AAS analysis. In the second method, M-OMS-2 was prepared without the addition of any manganese ion solution. The metal solution was oxidized by potassium permanganate in acidic condition. Typically, a 0.4 M solution of KMnO4 (13.3 g in 225 mL of deionized water) was added to a mixture of 1.75 M solution of metal and 6.8 mL of concentrated HNO3, except for Ti-OMS-2, no concentrated HNO3 was added. The resulting black precipitate was stirred vigorously and refluxed following the procedure in section 3.1.1. The samples were labelled as M-OMS-2. 3.2.2 Ion Exchange of OMS-2 and M-OMS-2 Samples H doped catalyst was prepared by ion exchange of potassium in the tunnel structure with acid. About 50 mL of concentrated HNO3 was added to 2.0 g of OMS-2 and M-OMS-2 samples. The slurry was stirred vigorously at 80 oC for 6 h. The product was filtered and washed several times with DDW. This procedure was repeated for successive ion exchanges to obtain greater H+ exchange in samples. The product was dried at 120 oC for 12 h, and then calcined at 280 oC for 6 h. The samples were labelled as H-OMS-2 and H-M-OMS-2. 59 3.2.3 Synthesis of Ti Incorporated OMS-2 (Ti-OMS-2) Titanium incorporated OMS-2 (Ti-OMS-2) was prepared in several Ti/Mn ratios. The samples were prepared by stepwise addition of solution KMnO4 (13.3 g in 225 mL of deionized water) to different amount of Ti2(SO4)3 15 % v/v in H2SO4 (25, 50 and 75 mL). The ratios of Ti:Mn were 0.18; 0.43 and 0.67 (as analyzed by atomic absorption spectrometer). Upon completion, the mixture was stirred, refluxed, filtered, washed, and dried following the procedure in section 3.1.1 and it is labeled as Ti-OMS-2 (0.18), Ti-OMS-2 (0.43) and Ti-OMS-2 (0.67), where the number in parenthesis is the molar ratio of Ti/Mn. 3.2.4 Synthesis of Ti Impregnated OMS-2 [Ti-OMS-2 (imp)] Titanium(IV) tetra-2-propoxide [Ti(OPri)4] was impregnated into OMS-2 powder in order to prepare the extraframework titanium on OMS-2. 69.6 mg of Ti(OPri4) was dissolved in 10 mL of toluene. About 1 g of OMS-2 sample was added to the solution and stirred overnight. After the evaporation of the solvent, the solid sample was calcined at 500 oC for 3 h. Here, this modified OMS-2 is labelled as Ti-OMS-2 (imp). The molar ratio of Ti/Mn was 0.18. 3.2.5 Preparation of TiO2-OMS-2 (mix) The physical mixture of OMS-2 and TiO2 rutile was prepared as comparison. The sample was prepared by mixing of 1 g of OMS-2 and 72.8 mg of TiO2 rutile. The mixture was grinded to homogenize the sample. The calculated molar ratio of Ti/Mn ratio is 0.67. The molecular weight of cryptomelane OMS-2 material was collected from JPSDS-29, 1020 that is 734.59 g/mol. It is labeled TiO2-OMS-2 (mix). 60 3.2.6 Synthesis of Sulphated Ti-OMS-2 (SO42-/Ti-OMS-2) Sulfation was done in order to create the acid site on Ti-OMS-2(0.67) sample. Sulfation was carried out by addition of 150 μL or 200 μL of H2SO4 (18 M) in 20 mL of solvent (water and toluene) to 1 g of catalyst and then stirred for 1 h. The product was filtered, washed with DDW several times, then dried at 120 oC overnight and calcined at 450 oC for 2 h. The samples were labelled as SW150-Ti-OMS-2 and SW200-Ti-OMS-2 for 150 μL and 200 μL of H2SO4, using water as solvent. Using toluene as solvent, the samples are labeled as ST150-Ti-OMS-2 and ST200-Ti-OMS2 for 150 μL and 200 μL of H2SO4, respectively. 3.2.7 Surface Modification by Alkylsilylation Alkylsilylation of n-octadecyltrichlorosilane (OTS) was done on selected samples; OMS-2 and Ti-OMS-2(0.67). Typically, a powder sample was immersed in 5 ml toluene containing 500 mol of OTS and the suspension was shaken for 5 min at room temperature. Then, the solid was collected by centrifugation and dried at 110 ºC for 5 h. 3.3 Characterization Techniques Characterization is a central aspect of catalyst development [Richards, 2006]. The elucidation of structures, compositions and chemical properties of both the solids used in heterogeneous catalysis and the adsorbates and intermediates present on the surface of the catalysts during reaction is vital for a better understanding of the relationship between catalyst properties and catalytic performance. The characterization techniques were in this work are described in the following subsection. 61 3.3.1 X-Ray Diffraction (XRD) Spectroscopy 3.3.1.1 Introduction X-ray diffraction (XRD) is a powerful method to define the crystallographic structure of crystalline materials whereby no other means is feasible or even possible. Each of the crystal materials has their own specific pattern that can be used as references for the determination of solid crystal phase and it is used as a fingerprint for every materials. It also provides information on the long range order, phase purity, change in lattice parameter with changing composition and enables one to assess preferred orientation effects, index diffraction patterns as well as to evaluate background and line broadening effects [135, 136]. The presence or absence of some peaks of the diffractogram indicates the existence of other crystal phase or the sample was contaminated with other phases. The crystal structure of the OMS-2 materials under investigation is considered similar to the structure of the reference sample, if all of the observed peaks are present with similar diffraction peaks as the reference sample. Extra or missing peaks observed indicate the presence of other crystalline phases, impurities or changes in the structure of the sample. A powdered XRD pattern is a plot of intensity of the diffracted beams as a function of 2 which satisfies the Braggs equation: n= 2d sin (Equation 2.4) where is the diffraction angle, d is the interplanar spacing, is the wavelength of the beam and n is an integer number. Wavelength of the CuK radiation is 1.5418 Å. The spacing of planes (hkl) or Miller indices is related to the unit cell parameters of lattice. The position of the diffraction peaks changes with the composition of the lattice. OMS-2 materials have the cryptomelane structure with tetragonal geometry. 62 In the tetragonal lattice, the unit cell parameter, ao and co can be calculated by using the following equation: 1 d2 h2 k 2 l 2 2 a2 c (Equation 2.5) 3.3.1.2 Experimental The synthesized samples were characterized by two model XRD instruments. First, powder XRD patterns were collected on a Bruker Advance D8 using Siemens 5000 diffractometer with Cu KD radiation (O = 1.5418Å,) operated at 40 kV and 40 mA. It was used to characterize the crytallinity, structure and phase of the samples. Typically, powder samples were ground, spread into a sample holder, and finally analyzed. The pattern was scanned in the ranges between 5° to 70° at a step of 0.020° and step time of 1s (scanning speed of 1.2°/min). As comparison, some selected samples were also characterized using Shimadzu XRD 6000 diffractometer with the Cu K ( = 1.5405 Å) radiation and the diffracted monochromatic beam at 30 kV and 30 mA at Universiti Putra Malaysia. 3.3.2 Atomic Absorption Spectroscopy (AAS) 3.3.2.1 Introduction Atomic absorption spectroscopy is one of the important instrumental techniques for both quantitative and qualitative analysis of metallic and nonmetallic elements in inorganic or organic materials [137]. This method provides a total metal content of sample and is almost independent of the molecular form of the metal in the liquid. The absorption of energy by ground state atoms in the gaseous state forms the basis of this technique. When a solution containing metallic species is introduced into a flame, the vapour of metallic species will be obtained; some of the metal atoms may be raised to a sufficiently high energy level to emit the 63 characteristic radiation of the metal. But a large percentage of the metal atoms will remain in the non-emitting ground state. These ground state atoms of a particular element are receptive of light radiation of their own specific resonance wavelength. Thus, when a light of this wavelength is allowed to pass through a flame having atoms of the metallic species, part of that light will be absorbed and proportioned to the density of the atoms in the flame. Therefore, once the absorption is known, the concentration of the metallic element can be determined [138]. 3.3.2.2 Experimental A Perkin-Elmer model AAnalyst 400 spectrophotometer was used to carry out the analyses of the amount of metal ions in the samples. Prior to analysis, the solid sample was diluted by decomposition using hydroflouric acid method. Approximately, 50 r 0.01 mg of a prepared, representative sample (200 mesh) was placed in a Teflon decomposition vessel with 0.5 mL aqua regia (HNO3 : HCl = 1 : 3 v/v). 3 mL of HF (48%) was added and the vessel was tightly sealed and placed in an oven at 110oC for 1 h. After cooling, the dissolution products are quantitatively transferred to a 50 mL plastic beaker containing 2.8 g of HBO3. 10 mL of distilled water was added and the mixture stirred magnetically to dissolve any insoluble fluorides. Finally, the clear solution was diluted to 100 mL and stored in a plastic bottle ready for major elemental analysis. 3.3.3 Fourier Transform Infrared (FTIR) Spectroscopy 3.3.3.1 Introduction Field emission scanning electron microscope (FESEM) is a scientific instrument that uses a beam of highly energetic electrons to examine objects on a very fine scale. The magnification produced by a scanning microscope is the ratio between the dimension of the final image display and the field scanned on the specimen. The FESEM does not produce a true image of the specimen. Instead, it 64 produces a point by point reconstruction of the specimen. The image of FESEM is produced by the signal which comes from the interaction of an electron beam with the specimen in the column of FESEM [139]. FTIR is a most widely applied analytical tool for identifying types of chemical bonds in a molecule by producing an infrared absorption spectrum that is like a molecular "fingerprint". By interpreting the infrared absorption spectrum, the chemical bonds in a molecule can be determined. Small differences in structure may result in significant changes in the spectra observed, and absorption in this region is probably unique for every molecular species. The region is extremely useful for the purpose of identifying a molecule. Molecular bonds vibrate at various frequencies depending on the elements and the type of bonds. For any given bond, there are several specific frequencies at which it can vibrate [140]. In general, the infrared region between 4000-200 cm-1 can be divided into four regions [135]: a. The X-H stretch region (4000-2500 cm-1), where strong contributions from OH, NH, CH and SH stretch vibrations are observed. b. The triple bond region (2500-2000), where the contributions of gas phase CO (2143 cm-1) and linearly adsorbed CO (2000-2200) are seen. c. The double bond region (2000-1500), where in catalytic studies bridge bonded CO, as well as carbonyl groups in adsorbed molecules (around 1700 cm-1) absorbs. d. The fingerprint region (1500-500), where all single bond between carbon and elements such as nitrogen, oxygen, sulphur and halogens absorb. e. The M-X or metal-adsorbate region (around 200-450 cm-1), where the metal-carbon, metal-oxygen and metal-nitrogen stretch frequencies in the spectra of adsorbed species are observed. IR is employed to study the framework structure, hydroxyl group and also molecules adsorbed in the heterogeneous catalyst. It provides a meaningful information in the mid-infrared region (1400-400 cm ) attributed to Mn-O vibration; -1 the characteristic of cryptomelane peaks. This is analogous with the various T-O 65 vibrational modes associated with the zeolites or mesoporous materials dominating most of the region between 1400-400 cm-1. 3.3.3.2 Experimental FTIR spectroscopy analysis was applied to determine the structure of OMS-2, M-OMS-2, H-OMS-2 and SO4/Ti-OMS-2. The infrared spectra were recorded on Spectrum One FTIR Spectrometer with 4 cm-1 resolution. Approximately 1 mg of sample was ground together with 100 mg of kalium bromide using pestle and mortar. The fine powder was then transferred to the ‘dye’ and 10 tonne of pressure was applied for 2 minutes. The resulting pellet was put in the sample holder and the FTIR spectrum of the sample was recorded in the range of 400 cm-1 to 4000 cm-1. 3.3.4 Total Specific Surface Area (BET) and Pore Volume Analysis 3.3.4.1 Introduction Adsorption isotherm is a unique and useful technique in measuring surface area and pore structure of a solid. The principal method of measuring total surface area of porous structures is by physical adsorption of a particular molecular species from a gas (typically nitrogen) onto the surface, maintained at a constant temperature (usually at liquid nitrogen temperature of 77K). One of the most commonly used measurements in molecular sieves research is the specific surface area as measured by Brunauer Emmet Teller (BET) method. This method involves adsorbing a monolayer of liquid nitrogen onto a surface of sample followed by measuring the amount of nitrogen that is released when the monolayer is vaporized. Based on this quantity, the surface area of the sample is calculated. Given the complexity of the pore structure in high-surface area catalyst, six types of adsorption isotherms (see Figure 2.12) have been identified according to a 66 classification advanced by IUPAC. However, only four type of adsorption isotherms usually found in catalysis [136, 141]: a. Type II, typical of macroporous solid where the prevailing adsorption processes are the formation of a monolayer at low relative pressures, followed by gradual and overlapping multilayer condensation as the pressure is increased. b. Type IV, often seen in mesoporous solids, where condensation occurs sharply at a pressure determined by Kevin-type rules. c. Type I, characteristic of microporous solids, where pore filling takes place without capillary condensation, and is distinguishable from the monolayer formation process. d. Type VI, corresponding to uniform ultramicroporous solids, where the pressure at which adsorption takes place depends on surface-adsorbate interactions, and shows isotherm with various steps each corresponding to adsorption on one group of energetically uniform site. nad P/Po P/Po P/Po VI nad V nad IV nad III nad II nad I P/Po P/Po P/Po Figure 3.2: The IUPAC classification for adsorption isotherms, where nad = amount of adsorbed and P/Po= relative pressure. 67 3.3.4.2 Experimental In this work, the isothermal N2 adsorption/desorption experiments were conducted on a Quantachrome Autosorb 1 series instrument. Before analysis, samples were outgassed at 200 °C under vacuum for 22 h. The relative pressure P/Po (P and Po are the pressures of N2 vapor at adsorption and its saturation vapor pressure at 77 K, respectively) used for the calculation is in the range of 0–0.3. 3.3.5 Thermal Gravimetry and Differential Thermal Analysis (TG-DTA) 3.3.5.1 Introduction Thermal gravimetry and differential thermal analysis (TG-DTA) is a technique whereby the weight of a sample and the phase transitions or chemical reactions can be followed through observation of heat absorbed or released over a period of time while its temperature is being raised linearly. Briefly, the sample is placed on the balance and the furnace for sample heating is installed beneath the balance. Then it will be heated and the electronic recording mechanism will plot a graph of weight and energy adsorbed (endothermic) or released (exothermic) against time, termed thermogram. TG-DTA can be used in studying: thermal degradation of a sample, chemical reaction resulting in changes of mass such as absorption, adsorption, desorption and also sample purity. 3.3.5.2 Experimental In this work, TG-DTA was performed using a Mettler Toledo TGA/SDTA 851 instrument under N2 atmosphere with a flow rate of 20 mL min-1 using about 25 mg of sample. The sample was heated in the temperature range of 30 to 900 °C at a heating rate of 10 °C min-1. 68 3.3.6 Field Emission Electron Scanning Microscopy (FESEM) 3.3.6.1 Introduction Field emission scanning electron microscope (FESEM) is a scientific instrument that uses a beam of highly energetic electrons to examine objects on a very fine scale. The magnification produced by a scanning microscope is the ratio between the dimension of the final image display and the field scanned on the specimen. The FESEM does not produce a true image of the specimen. Instead, it produces a point by point reconstruction of the specimen. The image of FESEM is produced by the signal which comes from the interaction of an electron beam with the specimen in the column of FESEM [139]. 3.3.6.2 Experimental FESEM experiments were carried out for the determination of particle size and morphology of the samples. Prior to sample scanning, the sample in powder form was attached to sample holder by using double sided tape. The samples were then coated with gold using BIO-RAD Polaron Division SEM Coating System machine. Samples were scanned using Zeiss Supra 35VP FESEM operating at 35 100 kV. 3.3.7 Photoluminescence 3.3.7.1 Introduction Photoluminescence spectroscopy is a contactless, nondestructive method of probing the electronic structure of materials. Photoluminescence (PL) is the spontaneous emission of light from a material under optical excitation [142]. PL spectroscopy provides electrical characterization, and it is a selective and extremely sensitive probe of discrete electronic states. Features of the emission spectrum can 69 be used to identify surface, interface, and impurity levels and to gauge alloy disorder and interface roughness. The intensity of the PL signal provides information on the quality of surfaces and interfaces. It gives a measure of the relative rates of radiative and nonradiative recombination. All solids, including semiconductors, have so-called “energy gaps” for the conducting electrons. In order to understand the concept of a gap in energy, first consider that some of the electrons in a solid are not firmly attached to the atoms, as they are for single atoms, but can hop from one atom to another. These loosely attached electrons are bound in the solid by differing amounts and thus have many different energies. Electrons having energies above a certain value are referred to as conduction electrons, while electrons having energies below a certain value are referred to as valence electrons and labeled as conduction and valence bands, respectively (see Figure 2.13). Furthermore, there is an energy gap between the conduction and valence electron states. Under normal conditions electrons are forbidden to have energies between the valence and conduction bands. Figure 3.3: The physical process following absorption of a photon by a molecule. 70 If a light particle (photon) has an energy greater than the band gap energy, then it can be absorbed and thereby raise an electron from the valence band up to the conduction band across the forbidden energy gap (see Figure 2.13). In this process of photoexcitation, the electron generally has excess energy which it loses before coming to rest at the lowest energy in the conduction band. At this point the electron eventually falls back down to the valence band. As it falls down, the energy it loses is converted back into a luminescent photon which is emitted from the material. Thus the energy of the emitted photon is a direct measure of the band gap energy, Eg. The process of photon excitation followed by photon mission is called photoluminescence. A spectrometer is used for measuring the intensity of light as a function of wavelength. 3.3.7.2 Experimental In this work PL analysis was done to study the effect of titanium incorporated OMS-2 to PL spectra of OMS-2 material. It should prove is it titanium exists in the framework structure or as non-framework TiOx particles. PL spectra were recorded in air at room temperature on Perkin-Elmer LS 55 spectrometer. The emission spectra were observed at excitation wavelength was 430 nm. 3.3.8 X- Ray Photoelectron Spectroscopy (XPS) 3.3.8.1 Introduction Surface analysis by XPS is accomplished by irradiating a sample with monoenergetic soft X-rays and analyzing the energy of the detected electrons. The photons have limited penetrating power in a solid on the order of 1-10 micrometers. They interact with atoms in the surface region, causing electrons to be emitted by photoelectric effect. The kinetic energy (KE) of the emitted electrons is given by following equation: 71 KE hv BE ) s (Equation 2.6) where hv is the energy of the photon, BE is the binding energy of the atomic orbital from which the electron originates, and s is the spectrometer work function. The binding energy may be regarded as the energy differences between the initial and final states after the photoelectron has left the atom. Because there is variety of possible final states of the ions from each type of atom, there is a corresponding variety of kinetic energies of the emitted electron. Besides, each element has a unique set of binding energies, XPS can be used to identify and determine the concentration of the elements in the surface. Variations in the element binding energies (the chemical shifts) arise from the difference in chemical potential and polarizability of compounds. These chemical shifts can be used to identify the chemical state of the materials being analyzed. 3.3.8.2 Experimental In this study, X-ray photoelectron spectra (XPS) were recorded using a Krastos XSAM–HIS (SAC) electron spectrometer fitted with a Mg K source. The anode was operated at 120 W (12 kV, 10 mA) and the analyzer was operated at constant pass energy of 40 eV. The binding energy shifts due to surface charging were corrected using the C 1s level at 284.6 eV. 3.3.9 Pyridine Adsorption 3.3.9.1 Introduction Another application of FTIR spectroscopy is that it has been combined and used in the characterization of nature of the acid sites (Brönsted and Lewis). The acidity of the solids characterized by absorbed base probe molecule i.e. pyridine. The choice of pyridine is due to its strong basicity property and its ability to interact 72 with a wide scale of acid strength as well as it can differentiate between the Brönsted and Lewis sites. A solid acid is capable of converting an adsorbed basic molecule into its conjugated acid form. The acid site is able to transfer a proton from the solid to the adsorbed molecule (Brönsted acid site) or an electron pair from the solid to the adsorbed molecule (Lewis acid site) [143]. By using this method of analysis, the Brönsted and Lewis acid sites could be differentiated. Pyridine selectively interacts with Brönsted and Lewis acid sites, giving rise to infrared absorptions in the region of 1400-1700 cm-1. The characteristic IR bands of adsorbed pyridine protonated by Brönsted acid sites (pyridinium ions) appear at 1545 cm-1 while the bands from pyridine coordinated to Lewis acid sites appear at 1450 cm-1 as shown in Table 2.7. Table 3.1: Position of bands for linkages of pyridine. Position of band (cm-1) Linkage of pyridine 1638 Brönsted acid sites 1620 Lewis acid sites 1580 Physisorbed 1545 Brönsted acid sites 1490 Brönsted and Lewis acid sites 1450 Lewis acid sites 1439 Physisorbed Lewis acid sites occur at the electron deficient sites and can accept a pair of electrons. Pyridine with its nitrogen lone pair of electrons acts as an electron donor while the metal acts as an electron acceptor. This interaction results in a peak at ~1455 cm-1 and can be shown as in Figure 2.14. The Brönsted acid is chemical species that is able to lose, or "donate" a hydrogen ion (proton). The existence of hydroxyl groups will give a peak at ~3600 cm-1. If the Brönsted acid site exists in the samples, the peak disappears after 73 introducing the pyridine and at the same time, a new peak at ~1545 cm-1 appears. This new peaks is due to pyridine bounded to Brönsted acid sites. The mechanism is depicted in Figure 2.15. O + M O Mn M OO O N N Mn + OO O Figure 3.4: The mechanism of interaction between pyridine molecules with Lewis. H O + N O Mn M O OO N H O O Mn M O OO Figure 3.5: The interaction between pyridine molecules with Brönsted acid sites. 3.3.9.2 Experimental The wafer of the sample (10–12 mg) was locked in the cell equipped with CaF2 windows, and evacuated at 400 oC under vacuum condition for 4 h. Pyridine as probe molecule was introduced into the evacuated sample for a minute at room temperature, followed by desorption of physisorbed pyridine at 150°C under vacuum for 1 h. IR spectra of the pyridine vibration region were monitored in the range of 1700–1400 cm-1 at room temperature. 74 3.3.10 Adsorption Capacity of Adsorbed Water A study on the adsorption capacity of adsorbed water was carried out in order to determine hydrophilicity relative of the samples. About 1 g of sample was dehydrated under vacuum at 100 ºC overnight. After dehydration, the sample was exposed to water vapour at room temperature, followed by the determination of the percentage of adsorbed water as a function of time. 3.3.11 Gas Chromatography (GC) Analysis Gas chromatography is widely used in most qualitative and quantitative analysis and it is basically limited to organic compounds that are volatile and thermally labile. The GC analysis technique was used to separate, identified and quantify the amount of reaction products. Gas chromatography is a physical separation method in which the components in a mixture are selectively distributed between the mobile phase, which is an inert carrier gas, and a stationary phase, which is present as a coating of either column packing particles of the inert column wall. The chromatographic process occurs as a result of repeated sorption/desorption step during the movement of the analytes along the stationary phase by the carrier gas. The separation is due to the differences in distribution coefficients of the individual components in the mixture. The instrumentation for GC consists of a gas control unit, a sample introduction system or injector, a column housed in the temperature-programmable column oven, and a detector or transfer line and/or interface to mass spectrometer. The gas control unit performs flow-rate or pressure control of the gas flows through the injector, the column, and the detector of carrier gas and, if required auxiliary gases. The carrier gas (hydrogen, helium, or nitrogen) typically is applied at a pressure below 0.3 MPa. 75 In this study, the GC analysis technique was used to separate and quantify the amount of reaction product. The analysis was performed using Thermo Finnigan Trace-GC instrument with Equity-1 capillary column (30 m x 0.25 mm ID, 0.25 m film thickness) and connected to FID detector. 3.3.12 Gas Chromatography-Mass Spectrometry (GC-MS) Analysis Gas chromatography-mass spectrometry (GC-MS) is a combination of two powerful analytical tools: gas chromatography for the highly efficient gas-phase separation of component in complex mixture, and a mass spectrometry for the confirmation of identity of these components as well as for the identification of unknowns. Compared to FID detector, this detector separates the compounds according to their mass-to-charge ratio. In this study, the GC-MS analysis technique was used to evaluate the reaction products. GC-MS analysis technique was applied using Agilent model G1540N GCMS instrument with HP-5MS capillary column (30 m x 0.25 mm ID, 0.25 m film thickness). 3.4 Catalytic Testing The catalytic reactions were done to study the physicochemical propertiescatalytic activities relationship of samples in oxidation and acid reactions. The reactions were carried out in oxidation of alcohol, alkanes and alkenes and also in consecutive oxidation and acid reactions for direct transformation of alkenes to diols as depicted in Figure 3.2. The figure also shows the detail of the catalyst used in each reaction. The following sections are the description of the reactions carried out to examine the catalytic properties of catalysts synthesized. 76 Benzyl alcohol to benzaldehyde with air as oxidant OMS-2a OMS-2b OMS Ti-OMS-2(X) Oxidation reaction Fe-OMS-2(0.09) Cyclohexane to cyclohexanol and cyclohexanone with TBHP as oxidant Cu-OMS-2(0.02) Co-OMS-2(0.04) H-OMS-2 H-Ti-OMS-2(X) H-Fe-OMS-2(0.09) H-Co-OMS-2(0.04) H-Cu-OMS-2(0.02) Cyclohexene to 2-cyclohexenone and 2-cyclohexenol Catalytic activity Alkenes with TBHP as oxidant OMS-2 TiO2 Ti-OMS-2(Y) TS-1 OMS-2 Ti-OMS-2(Z) Styrene to styrene oxide, TiO2-OMS-2(imp) benzaldehyde and phenylacetaldehyde TiO2-OMS-2(mix) OTS/OMS-2 OTS/Ti-OMS-2 SW150-Ti-OMS-2(Y) Consecutive oxidation and acid reaction 1-octene to 1,2 octanediol with TBHP as oxidant SW200-Ti-OMS-2(Y) ST150-Ti-OMS-2(Y) ST200-Ti-OMS-2(Y) Note: X refer to Ti/Mn ratio = 0.05 and 0.67; Y refer to Ti/Mn ratio = 0.18 and 0.67 and Z refer to Ti/Mn ratio = 0.18; 0.43 and 0.67. Figure 3.6: Catalytic reactions over various modified OMS-2. 77 3.4.1 Oxidation of Benzyl Alcohol 3.4.1.1 Introduction Solid-gas catalyzed-liquid reactions are often encountered in the chemical process industry, most frequently in hydroprocessing operations and in the oxidation of organic liquid phase [144-147]. In respect of the latter reaction type, the system has been applied for mineralization of toxic organics in wastewaters. However, synthesis of useful organic materials with this system is also one of particular fields of application. In this work the model reaction is the oxidation of benzyl alcohol with air as oxidant. The use of air as oxidant is advantageous due to its environmentally friendly nature and no increase in cost. Furthermore, oxidation of benzyl alcohol is one of the important oxidation processes since benzaldehyde as product is an important intermediate for the processing of perfume and flavouring compounds (almond flavour) and also as starting materials for other organic compound ranging from pharmaceutical to plastic additives. One of the active materials for oxidation of benzyl alcohol with molecular oxygen is OMS-2 material. The material shows the rare Mars-van Krevelen type mechanism for oxidation with molecular oxygen in the liquid phase [24]. The process usually occurs in gas phase oxidation reaction using metal oxide/mixed metal oxide as a catalyst [148-152]. However, the other liquid phase oxidation in polyoxometalate was reported via that mechanism [29]. This mechanism is interesting because by controlling the lattice oxygen diffusion (with a proper selection of oxide), an effective catalyst for liquid-gas system could be obtained (as catalyst/active site). 3.4.1.2 Experimental Catalytic reaction of samples was studied for oxidation of benzyl alcohol. A 15 mL of toluene, 1 mmol of benzyl alcohol, and 0.5 mmol of methyl benzoate as an internal standard were put in a round-bottomed flask. The catalyst (~50 mg (0.5 eq), considering one manganese atom to be one active site) was added and a reflux 78 condenser was attached. This assembly was placed in a silicon oil bath. The reaction was thus carried out under semibatch conditions for 4 h. The reflux conditions ensured a constant temperature equal to the boiling point of toluene at 110 o C. Samples of the reaction mixture were collected and centrifuged to separate the solid before they were analyzed by GC and/or GC-MSD. 3.4.2 Oxidation of Cyclohexane 3.4.2.1 Introduction Functionalisation of alkanes is of great interest in both theory and practice due to their abundance and extremely low activity. Petroleum and natural gas are the main sources of alkanes and about of 5% of nearly a billion ton of petroleum extracted in the world is chemically processed to alkanes [153]. The inertness of alkanes is the main obstacle to the development of methods for their functionalisation in reactions with these compounds occurs under drastic conditions. The main pathway of alkanes functionalization is their oxidation. The oxidation of alkanes takes place by two main paths: compounds containing oxygen (partial oxidation) and CO2 and H2O (extensive oxidation). The partial oxidation of alkanes is significant to the chemical industry because these oxidation reactions are used to convert petroleum hydrocarbon feedstocks into chemicals important in the polymer and petrochemical industries. Among various alkanes oxidation, partial oxidation of cyclohexane is much attractive because of its main products i.e. cyclohexanol and cyclohexanone are very important chemical intermediates. They are as starting materials to produce adipic acid and caprolactam which are used in the manufacture of nylon-66 and nylon-6 polymer, respectively. Currently industrial process of oxidation of cyclohexane is its liquid-phase oxidation with air at 150 oC in the presence of soluble cobalt catalyst [154]. The drawbacks of the process are the poor conversion of cyclohexane (about 79 4%) and selectivity for cyclohexanone and cyclohexane (K-A-oil) of just about 70– 85%, depending on the conversion of cyclohexane. The intensive study was done in oxidation of cyclohexane in homogeneous and heterogeneous catalysts [155-159]. A review of oxidation of cyclohexane by Schuchardt et al. [160] concluded that further work for finding the active catalyst in the oxidation of cyclohexane is still a great challenge. 3.4.2.2 Experimental The oxidation reactions were carried out at atmospheric pressure, as follows: 100 mg of catalyst was suspended in a mixture of 27.8 mmol (3 mL) of cyclohexane (Merck), 10 mmol of TBHP (70% in water). The reaction mixture was heated under reflux with magnetic stirring in an oil bath at 60 oC. After reaction, the catalyst was removed and gas chromatography coupled to mass spectrometry (GC–MS, Agilent model G1540N, DB-1MS 20M capillary column) was used to identify the reaction products, which were quantified by gas chromatography (Trace GC) coupled to a flame ionization detector, using an internal standard (cyclooctane, Fluka) and calibration curves. 3.4.3 Oxidation of Cyclohexene 3.4.3.1 Introduction Alkenes, either derived from natural resources or generated as products of the chemical industry, are found in great abundance in the realm of organic molecules. Heterogeneously catalyzed oxidations of alkenes have been widely applied in numerous chemical, biological and pharmaceutical industries. The selective oxidation products are important starting materials towards the production of several fine chemicals and polymers [28, 161]. One of the most useful transformations of alkenes is epoxidation. The epoxides can be transformed into a variety of functionalised products. For example, reductions, rearrangements or ring-opening 80 reactions with various nucleophiles give diols, aminoalcohols, allylicalcohols, ketones, polyethers etc. However, the epoxidation process in oxidation of alkenes is competed by C-H bond oxidation, thus making the process more complex and the selectivity lower. Selective oxidation of cyclohexene is one of the important alkenes oxidation. The oxygenated products of cyclohexene and their derivatives are very important in organic synthesis owing to the existence of a highly reactive ,-unsaturated carbonyl group, which are extensively used in the preparation of a range of chemical intermediates and products [162]. 3.4.3.2 Experimental The catalyst performances were tested in the oxidation of cyclohexene using tert-butyl hydroperoxide (TBHP) as oxidant. Cyclohexene (5 mmol), 70% (wt. %) TBHP in water (10 mmol), catalyst (50 mg), cyclooctane (0.5 mmol) as internal standard and acetonitrile (15 ml) as solvent were placed in a round-bottomed flask with a reflux condenser and the reaction was performed with stirring at 80 °C in an oil bath for 24h. In order to evaluate qualitatively the reaction products, the GC-MS analysis technique was applied using Agilent model G1540N GC-MS instrument with HP-5MS capillary column (30 m x 0.25 mm ID, 0.25 m film thickness). The analysis of products quantitatively is using Thermo Finnigan Trace-GC instrument with Equity-1 capillary column (30 m x 0.25 mm ID, 0.25 m film thickness) and connected to FID detector. 81 3.4.4 Oxidation of Styrene 3.4.4.1 Introduction The other interesting alkenes oxidation is the liquid-phase oxidation of styrene; from which produced high value products such as styrene oxide and phenylacetaldehyde. The rearrangement of styrene oxide to benzaldehyde also yields valuable compounds or intermediates for production of fragrances, pharmaceuticals, insecticides, fungicides and herbicides [163-166]. 3.4.4.2 Experimental The oxidation of styrene was carried out using the above catalysts. Styrene (5 mmol), 70% (wt. %) in water of ter-butyl hydroperoxide (TBHP) (10 mmol), catalyst (50 mg) and acetonitrile (15 ml) as solvent were placed in a round-bottomed flask with a reflux condenser and the reaction was performed with stirring at 70 oC in an oil bath. The products were collected after 3 h of reaction and analyzed by GC and GC-MS. 3.4.5 Transformation of 1-octene to 1,2-octanediol 3.4.5.1 Introduction Diols are important starting materials for polyurethane chemistry, natural products, drugs and agrichemicals [120, 121, 124, 167]. A number of 1,2-diols such as 2,3-dimethyl-2,3-butanediol, 1,2-octanediol, 1,2-hexanediol, 1,2-pentanediol, 1,2and 2,3-butanediol are of interest to fine chemical industries. In addition chiral 1,2diols are employed as intermediates for pharmaceuticals and agrichemicals [125]. Industrial diols are synthesized by two step reaction that is epoxidation of alkenes followed by hydrolysis of epoxide in the presence of Brönsted acid sites. 82 The direct transformation of alkenes to diols was reported by Prasetyoko et al. [5254] using a bifunctional oxidative and acidic catalyst using aqueous H2O2 as oxidant. The catalyst is the titanium silicate-1 (TS-1) loaded with sulphated zirconia or niobium oxide and it was proven that TS-1 acted as oxidative site and sulphated zirconia or niobium loaded on TS-1 as acidic site, respectively. Therefore it is interesting to determine whether the sulphated Ti-OMS-2 act similarly to the TS-1 catalyst. 3.4.5.2 Experimental The oxidation of 1-octene was carried out using ter-butyl hydroperoxide (TBHP) as an oxidant. A typically reaction mixture was composed of 50 mg catalysts, 1-octene substrate 1 mL, 1 mL of TBHP, 10 mL of aceton as solvent, and 0.5 mmol of cyclooctane as internal standard. The mixture placed in round-bottomed flask with reflux condenser. This assembly was placed in a paraffin oil bath. The reaction mixture was vigorously stirred at 60 °C. The products were collected after 24 h of reaction and analyzed by GC and GC-MS. CHAPTER 4 PHYSICOCHEMICAL PROPERTIES OF OMS-2 AND MODIFIED OMS-2 CATALYSTS 4.1 Introduction This chapter describes the physicochemical properties of OMS-2 and modified OMS-2 catalysts. The samples were characterized by XRD, FTIR, surface area analyzer, FESEM, TG-DTA, XPS, pyridine adsorption and adsorption of vapour water. These characterizations are very important for the study of the structurecatalytic activity relationship. The physicochemical properties were studied including structural analysis, thermal stability, surface area and acidity study. Besides using the previous method for incorporation of metal into the framework of cryptomelane OMS-2, a new method was attempted. The only solution of transition-metal ion (without any manganese ion solution) in acidic condition was oxidized by solution of potassium permanganate. This method gives higher M/Mn molar ratio in M-OMS-2. However, this method was only suitable for Ti2(SO4)3 as source of metal. The other metal sources such as FeCl3, FeSO4.7H2O, CoSO4, CuSO4.5H2O and TiCl4 can not retain the cryptomelane structure of OMS-2. 84 4.2 Physicochemical Properties of Prepared OMS-2 by Different Methods As mentioned before in Section 2.2.2, OMS-2 prepared in acidic condition gave the best catalytic activity in oxidation reaction. Based on that, OMS-2 was prepared in acidic condition. However, there were only two methods to prepare OMS-2 in acidic condition reported i.e. with and without buffer solution. Hence in this work, OMS-2 was prepared with and without buffer solution in order to study the physicochemical properties-catalytic activity relationship of both samples and then to choice the best method to prepare OMS-2. The preparation methods are described in Section 3.1.1 and 3.1.2. OMS-2 prepared without buffer solution was labeled as OMS-2a and OMS-2 prepared with buffer solution as OMS-2b. The synthesis of cryptomelane from KMnO4 and Mn2+ depended on pH, temperature and the type of countercation. The only variable is pH while temperature and the type of countercation are the same for both prepared OMS-2 samples. The buffer solution was used to control the pH of the mixture during the synthesis. pH played an important role in determining the success of growing cryptomelane structure during the formation of OMS-2 materials. The amount of counteraction also influenced the synthesis of OMS-2. The molar ratios of KMnO4/[KMnO4] + [MnSO4]) for OMS-2 prepared with and without buffer were about 0.52 and 0.41, respectively. This was based on the work by DeGuzman et al. [62] who studied the effect of potassium concentration under reflux in acidic condition as shown in Table 4.1. If the molar ratio of KMnO4/[KMnO4] + [MnSO4]) is 0.32 or lower, OMS-2 was formed with low crystallinity. In the synthesis, when KMnO4 was added to Mn2+ solution in acidic condition, amorphous MnO2 was formed. The color of the reaction mixture turned immediately from light pink to light brown and finally a dark brownish or black precipitate was obtained. The addition of KMnO4 by dropwise to mixture of Mn2+ solution increased the crystallinity of OMS-2 formed. After several hours of heating (in this work 100 oC), the cryptomelane phase was formed. After washing and drying, a black and brown solids for OMS-2a and OMS-2b were obtained. 85 Table 4.1: Effect of potassium concentration under reflux on the synthesis of OMS-2 [62]. KMnO4/[KMnO4] + [MnSO4]) The phase formed 0.88 Cryptomelane 0.74 Cryptomelane 0.41 Cryptomelane 0.36 Cryptomelane 0.32 Cryptomelane (poor crystallinity) 0.28 Cryptomelane (poor crystallinity) 0.26 Cryptomelane (poor crystallinity) The X-ray diffraction patterns of OMS-2a, OMS-2b and standard synthetic cryptomelane (KMn8O16) [168] are shown in Figure 4.1. XRD patterns of both synthesized OMS-2 samples are similar to the standard pattern of cryptomelane, Q. This indicated that both synthesized OMS-2 materials were of cryptomelane type structure. Cryptomelane consists of a well-defined 2 x 2 tunnel structure having a pore size of about 4.6 Å and composed of double chains of edge-sharing MnO6 octahedra and corner-sharing double chains. The 2 x 2 tunnel structure needs some large ions, such as K+ ion in the tunnels, to prevent collapse of the framework. In addition, no other peaks are observed in both samples which indicate that no other phase of MnO2 such as a pyrolusite (a one-dimensional tunnel manganese oxide with 1×1 MnO6 units) was observed; an impurities phase that is normally observed in highly acidic condition especially in pH < 1. The pH of the mixture during synthesis of OMS-2a and OMS-2b sample were in the range of 1-2 and 4.5, respectively. Therefore, both conditions were suitable for the synthesis of pure phase cryptomelane. 86 Intensity / a.u. (c) 5 10 20 50 (541) (002) (521) (600) (510) 40 30 (411) (301) (211) (310) (220) (200) (110) (b) (a) 60 70 2 /º Figure 4.1: XRD patterns of (a) reference pattern of crypromelane, Q JCPDS 29, 1020, (b) OMS-2a and (c) OMS-2b. There are no significant differences in the relative peaks heights of OMS-2b sample and cryptomelane-Q; indicated that both samples have similar preferred orientation of crystal structure. However, there is a difference of relative peaks heights of OMS-2a compared to OMS-2b and cryptomelane Q, in which the relative intensity of (200) to (310) plane (as indicated by the arrow in Figure 4.1) in OMS-2a > 1 while in OMS-2b and Cryptomelane-Q are that is <1. This indicates that there is a difference between preferred orientation of the crystal in OMS-2a sample compared to OMS-2b and cryptomelane-Q. It is also observed that the peak intensity of OMS-2b sample is higher than OMS-2a sample. The higher peak intensity indicates the more crystalline is the sample. Therefore, OMS-2b is more crystalline than OMS-2a. This may be caused by the acidity (pH) of the mixture during synthesis. As mentioned above, the reaction mixture in the synthesis of OMS-2b with the buffer solution was maintained at pH of around 4.5 and synthesis of OMS-2a without buffer had pH in the range of 1-2. The high acidity of the mixture in the synthesis of OMS-2a rendered 87 crystallization of OMS-2a faster than crystallization of OMS-2b. The consequence is lower crystallinity of OMS-2a compared to OMS-2b. From the figure it is also observed that diffraction peaks of OMS-2a sample as boarder as OMS-2b sample implying that particles size of both samples is similar. The thermal stability of OMS-2 was studied by calcination of OMS-2 sample in a series of temperatures. XRD pattern of the calcined OMS-2b sample is shown in Figure 4.2. It shows that no extra line is observed for calcined sample until 500 oC which indicates that the sample was not decomposed at this temperature. Extra lines are observed when the sample was calcined at 600 oC for 5 h; indicating the formation of the other phases. The appearance of these peaks indicate that a part the samples have been transformed from cryptomelane structure to hausmannite structure (Mn3O4). This phase was also observed by Chen et al. [69] during oxidation process of 2-propanol on OMS-2, where cryptomelane was reduced to hausmannite by 2-proponal, which at the same time is oxidized to CO2. Compared to the other synthetic manganese oxides such as OMS-1, generally the thermal stability OMS-2 material is higher. The thermal stability of OMS-1 is less than 400 oC [56]. OMS-2 is more thermally stable than OMS-1 due to the smaller pore opening of OMS-2 than OMS-1. OMS-2 utilizes two MnO6 octahedra on each side to form a 2x2 square tunnel with a pore size of about 4.6 Å and similarly, OMS-1 has a 3x3 square tunnel with a pore size of about 6.9 Å. There are some significant differences in the relative peaks heights of OMSOMS-2b after calcination. Table 4.2 shows the relative intensity and ratio of I(110)/I(200) planes (as indicated by arrow in Figure 4.2) of OMS-2b and calcined OMS-2b samples. The relative intensity of peaks of 200 to 110 planes decreased by increasing of temperature up to 500 oC, indicating the existence of different preferred orientation of crystal after calcinations until that temperature. However, the relative intensity of (110)/(200) plane became higher after calcination at 600 oC, implying the changing of the preferred orientation on of the crystal after that temperature. This may be due to the transformation of cryptomelane to hausmanite. 88 110 200 * hausmannite * * (d) Intensity / a.u. (c) (b) (a) 5 10 20 30 40 50 60 70 2 /º Figure 4.2: Effect of calcination on XRD patterns of OMS-2b materials, (a) before calcination, (b) calcination at 400oC, (c) at 500oC, and (d) at 600oC. Table 4.2: The relative intensity and ratio of I(110)/I(200) plane of OMS-2b and calcined OMS-2b samples calculated by XRD. Relative intensity (hkl) samples I(110)/I(200) 110 200 OMS-2b 75 68 0.90 OMS-2b (400) 87 56 0.64 OMS-2b (500) 44 24 0.54 OMS-2b (600) 66 65 0.98 89 FTIR spectra of samples are shown in Figure 4.3. The bands in the range of 500-700 cm-1 and 400-500 cm-1 are assigned to the Mn-O stretching of MnO6 octahedra and the Mn-O-Mn bending vibration in the MnO2 octahedral lattice, respectively [169]. Both samples have peaks at 600 and 520 cm-1 which are characteristic of IR spectra for cryptomelane structure [170]. This is in agreement with XRD data in which both samples have cryptomelane type framework structure regardless of different synthesis conditions. The bands around 2800-3600 cm-1 and 1620 cm-1 were observed in OMS-2a sample and not observed in OMS-2b sample. Relative strong of these bands are due to water adsorbed on the surface of OMS-2a and some hydroxyl groups not from hydrates but those directly bound to metal ions [171] due to the defect on the structure as shown in Figure 4.4. This is in agreement to XRD analysis in which OMS-2a has the lower crystallinity of than OMS-2b. These bands are not observed in OMS-2b samples indicating that there was no adsorb water in OMS-2b and vibration of Mn-OH in OMS-2b. Consequently, OMS2b is more hydrophobic compared to OMS-2a. This is also supported by the capacity of adsorbed water vapour calculated on both OMS-2 samples as shown in Figure 4.5. 600 520 Figure 4.3: FTIR spectra of (a) OMS-2a and (b) OMS-2b. 90 O O O O Mn O 4+ Mn 4+ O OO 4+ O O O O O OO O 4+ Mn O O Mn O 4+ O Mn O O O OH 4+ 4+ Mn3+ O O O Mn4+ O O Mn Mn4+ HO O Mn4+ Mn O Mn4+ OO O Mn4+ O O 4+ O Mn O Mn O O O O 4+ O OO 4+ Mn Mn O O Mn3+ Mn4+ O O OO O 4+ O Mn4+ O O Figure 4.4: Defect on OMS-2 structure. Figure 4.5: Amount of adsorbed water on the surface of OMS-2a and OMS-2b samples. 91 As shown in the figure, unbuffered sample (OMS-2a) absorbed water vapour almost four times more than buffered sample (OMS-2b), confirming that the surface of OMS-2a is more hydrophilic than OMS-2b. The thermal stability of the sample was tested by TG analysis. Figure 4.6 shows thermograms of OMS-2a and OMS-2b in nitrogen atmosphere. The weight loss below 200 ºC is due to the removal of water molecules present on the surface of manganese oxide. It shows that the adsorbed water on the surface of OMS-2a and OMS-2b are about 2.5% and 0.9%, respectively. It indicates that the surface of OMS-2 with buffer (OMS-2b) is more hydrophobic than without (OMS-2a), which is in agreement with the IR analysis. This is also supported by the adsorbed water vapour calculated on the surface of both OMS-2 samples as shown in Figure 4.5. Figure 4.6: TGA plots for OMS-2 materials in N2 atmosphere. The weight losses at higher temperatures (>200 ºC, stage II) are due to the evolution of oxygen from the lattice, as confirmed by temperature programmed decomposition with mass spectroscopic detection done by Luo et al. [26]. Some of 92 the oxygen in the OMS-2 framework could be easily dislodged from the framework and evolved into air without destroying the overall framework structure when the sample was heated in nitrogen flow at as high as 520°C. The evolution of oxygen resulted in the formation of framework oxygen vacancies, which may be catalytically active sites for the oxidation reactions. At this stage OMS-2a sample lost about 2.5% of its weight and OMS-2b only about 0.8%, implying that it was easier to release oxygen lattice on OMS-2a than OMS-2b. At 520 ºC (stage III) both materials were decomposed into a stable lower valence manganese oxide like Mn3O4 (hausmannite) and lost of O2 due to the collapse of tunnel structure. Both samples showed that their structure collapsed at similar start points. However, the different environments may cause different stability of OMS-2 sample that decomposed at lower temperature in nitrogen or helium compared to in air [26]. At higher temperature (>720 ºC, stage IV) this material further decomposed to other phase i.e. bixbyite. OMS-2b sample started to decompose at a lower temperature than OMS-2a. The phase transformation of manganese oxide during the heating is given below: OMS-2 Mn3O4 Mn2O3 Base on the discussion above, it shows that OMS-2a and OMS-2b samples which were prepared without and with buffer, have different physicochemical properties. OMS-2a is less crystalline and less hydrophobic compared to OMS-2b. The thermal stability of both OMS-2 samples is similar. However, it is suggested that oxygen lattice of OMS-2a was easier to be released than OMS-2b. 4.3 Physical Properties of Metal Substituted OMS-2 Material The XRD patterns of synthesized M-OMS-2(X) samples prepared using the first method (the mixture of metal source and Mn2+ in acidic condition was oxidized by potassium permanganate solution) are shown in Figure 4.7. It shows that all 93 Figure 4.7: X-ray diffractograms of (a) OMS-2; (b) Ti-OMS-2(0.05); (c) Fe-OMS2(0.09); (d) Co-OMS-2(0.02); and (e) Cu-OMS-2(0.04). synthesized M-OMS-2(X) have similar XRD patterns to that of OMS-2 pattern which indicates that all M-OMS-2(X) have cryptomelane-type structure. No new peaks due to metal species were observed in M-OMS-2(X) confirming that metals are well incorporated in the framework structure of OMS-2 as depicted in Figure 4.8. Figure 4.8: Schematic incorporation of metals in M-OMS-2 materials. 94 The figure shows that the peaks intensity of M-OMS-2(X) except Ti-OMS2(0.05) are lower than that of pure OMS-2 which indicate that crystallinity of MOMS-2(X) samples are lower than OMS-2. Ti-OMS-2(0.05) has higher intensity compared to OMS-2 which indicates that Ti-OMS-2(0.05) is more crystalline than OMS-2. Ti-OMS-2(0.05) sample has narrower peaks compared to pure OMS-2 which implies a bigger crystallite size in Ti-OMS-2 than OMS-2 sample. Fe-, Cu-, and Co-OMS-2 samples have broader peaks compared to pure OMS-2 implying the smaller crystallite size of those M-OMS-2 than the pure one. The average crystallite sizes of all materials were calculated from the Scherrer equation and are listed in Table 4.3. Table 4.3: The ionic radii of metals ion and average crystallite size of OMS-2 and M-OMS-2 samples. Samples a Average crystallite size (Å)a OMS-2 208 Ti-OMS-2(0.05) 270 Fe-OMS-2(0.09) 199 Co-OMS-2(0.04) 178 Cu-OMS-2(0.02) 186 average crystallite size was calculated from the Scherer equation OMS-2 is the mixed valence of manganese oxide which has an average oxidation state (AOS) of manganese at around 3.8-3.9. This means that large amount of manganese exists as Mn4+. The amount of Mn3+ and or Mn2+ is less than 10% of the total quantity of manganese ion. One expects that any metal ion could replace manganese ion in the same oxidation state. The initial oxidation state of metal ions are Ti3+, Fe2+, Co2+ and Cu2+. The synthesis using potassium permanganate as oxidation agent faced the possibility of metal ions being oxidized to the higher oxidation state (Table 4.4). All transition metals may be oxidized to higher oxidation state except Cu since the maximum oxidation state of Cu is +2, consequently Cu favoured to exist as Cu2+. XPS analysis confirms that titanium exists as Ti4+ and there was no Ti3+ was observed in Ti-OMS-2 sample (discussed in Section 4.4.5). 95 Table 4.4 also shows the lattice parameters and cell volume of OMS-2 and M-OMS-2. The lattice parameters and cell volume of OMS-2 changed after incorporation of metal indicating that metal were incorporated in the framework structure of OMS-2 material. In Ti-OMS-2 sample, the possibility of titanium with 4+ charge to be substituted to Mn4+ caused Ti-OMS-2 to has bigger cell volume size, since the ionic radii of Ti4+ was bigger than Mn4+. As shown in Table 4.4, for the other metal ions, ionic radii of metals with the same charge with Mnn+ ions are smaller than ionic radii of Mnn+. This may due to the cell volume of Fe, Co and CuOMS-2 being smaller than OMS-2 sample. Table 4.4: Lattice parameter (a and c) and cell volume (V) of OMS-2 and M-OMS-2 samples. a (Å) OMS-2 9.81 c (Å) 2.82 V (Å3) 271 The Ionic Possibility possibility radii of of Mnn+ of Mn+ ion octahedra ion was in the n+ M (Å) substituted framework Mn2+ 0.83 3+ Mn 0.65 Mn4+ 0.53 - 0.61 Mn4+ 4+ Ti-OMS-2 9.82 2.82 272 Ti Fe-OMS-2 9.80 2.81 270 Fe2+/Fe3+ 0.61/0.55 Mn2+/Mn3+ Co-OMS-2 9.81 2.81 270 Co2+/Co3+ 0.65/0.55 Mn2+/Mn3+ Cu-OMS-2 9.81 2.79 268 Cu2+ 0.73 Mn2+ The intensity of the (310) and (211) reflections as shown by the arrows in Figure 4.7 are varied by incorporation of metals. The calculation of the relative peaks heights of I(310)/I(211) plane of M-OMS-2(X) samples are shown in Table 4.5. The table shows that only Ti-OMS-2(0.5) has significantly different in ratio of I(310)/I(211) plane and become higher indicated that Ti incorporated caused preferred orientation in OMS-2 crystal. For the other metals such as Fe-OMS2(0.09) and Cu-OMS-2(0.02) have lower of ratio of I(310)/I(211) plane than OMS-2. This indicated that preferred orientation of Ti-OMS-2(0.5) is opposite to Fe-OMS- 96 2(0.09) and Cu-OMS-2(0.04). While Co-OMS-2 has ratio of I(310)/I(211) plane very close to OMS-2 (i.e. 0.74 and 0.73, respectively) indicated that no preferred orientation in OMS-2 crystal in that plane after incorporation of Co. Table 4.5: The relative intensity and ratio of I(310)/I(211) plane of OMS-2 and MOMS-2 samples. Relative intensity of (hkl) Samples I(310)/I(211) 310 211 OMS-2 73 100 0.73 Ti-OMS-2(0.05) 100 93 1.07 Fe-OMS-2(0.09) 65 100 0.65 Co-OMS-2(0.02) 74 100 0.74 Cu-OMS-2(0.04) 56 100 0.56 The XRD patterns of M-OMS-2 prepared without any manganese ion solution are shown in Figure 4.9. The preparation was done in order to get higher M/Mn ratio. Apparently, the peaks due to cryptomelane structure collapse upon substitution by the Cu, Co and Fe metal ions. Substitution by Cu caused a decrease in the intensity of the main peaks of cryptomelane and is broaden indicating its amorphous nature. Substitution of Co is truly amorphous as well. Fe- OMS-2 patterns have new peaks corresponding to MnO and Fe2O3. This demonstrates that the high amount of Cu, Co and Fe renders the structure of cryptomelane amorphous. It is shown that only Ti-OMS-2 has maintained its cryptomelane structure. This suggests that OMS-2 structure is retained and confirms the ability of Ti to efficiently substitute manganese in the framework structure. This may be due to the factors, such as similar electronegativity, oxidation state of metal ions and the source of metal ions. 97 Figure 4.9: XRD pattern of (a) cryptomelane (JCPDS 29, 102) and (b) Ti-, (c) Cu-, (d) Co- and (e) Fe-OMS-2 was prepared by oxidation of its metal ions source by potassium permanganate in acidic condition without the addition of Mn2+ solution. Pauling electronegativity of elements is listed in Table 4.6. It shows that electronegativity of Ti is very close to electronegativity of Mn. This explained the fact that high content of titanium still retained in the cryptomelane structure of OMS2. Whereas the electronegativity of the other elements (Fe, Co and Cu) is very different from Mn which could caused the collapse the of cryptomelane structure after the incorporation of high content of these elements. Table 4.6 also shows the oxidation state of Mn in OMS-2 and the possibility of oxidation state of the substituted metal ions. It is clearly observed that only titanium fully fill all oxidation state of Mn in OMS-2 materials, due to the ability of Ti to efficiently substitute manganese in the framework structure. 98 Table 4.6: The physical properties of metal ions Element Electronegativity The possible of oxidation state of metal ions Mn 1.55 +2, +3, +4 (in OMS-2) Ti 1.54 +2, +3, +4 Fe 1.83 +2, +3 Co 1.88 +2, +3 Cu 1.90 +1, +2 The effect of source of metal ions was studied by the different source of Ti used i.e. 15% w/v Ti2(SO4)3 in H2SO4, TiCl4 and Ti(SO4)2 as listed in Table 4.8. It shows that cryptomelane structure collapse by using of TiCl4 and Ti(SO4)2 as metal ions source. This may be due to Ti4+ of those metal sources could not be oxidized to higher oxidation state by potassium permanganate. Since synthesis of OMS-2 was via redox reaction where Mn2+ was oxidized and potassium permanganate was reduced. The use of Ti3+ {source Ti2(SO4)3} could replace the position of Mn2+ and was oxidized to Ti4+, while Ti4+ could not. Although Cu, Co and Fe has the same charge with Mn2+, they could not retain the cryptomelane structure which may be caused by Table 4.7: Source of metal, its charge and their effect to cryptomelane structure. Metal source Metal ion Structure MnSO4. H2O Mn2+ OMS-2 15% w/v Ti2(SO4)3 in H2SO4 Ti3+ retain of OMS-2 structure TiCl4 Ti4+ collapse Ti(SO4)2 Ti4+ collapse CuSO4. 5H2O Cu2+ collapse CoSO4. 7H2O Co2+ collapse FeSO4. 7H2O Fe2+ collapse FeCl3 Fe3+ collapse 99 the inability of those metal ion to be in oxidation state +4, which is the main oxidation state in OMS-2. Similarly, the use of Fe3+ with the same charge with Ti3+ could not retain the structure of OMS-2 since it could not be oxidized to +4 oxidation state. The results of AAS elemental analyses for cations in all synthesized M-OMS2 are summarized in Table 4.8. Although the same initial molar ratio of metal cation was used in the reactants (0.1) in the syntheses of different metal doped M-OMS-2 materials except for Ti substituted, the final molar ratios of M:Mn in prepared materials are different from each other. For Ti substituted OMS-2 the initial Ti:Mn ratio (0.05) remain unchanged due to their physicoproperties such as electronegativity and oxidation state of Ti very closed to Mn. Based on the initial ratio of M:Mn, the doped metal content decreases in the order of Ti3+>Fe2+>Co2+>Cu2+. The trend corresponds to electronegativity of metal ion which electronegativity of Ti very closed to Mn and followed by Fe, Co and Cu, respectively (Table 4.6). The elemental analysis of Ti-OMS-2 was also determined. The ratio of Ti:Mn in Ti-OMS-2 sample was 0.67 and labelled as Ti-OMS-2(0.67). Table 4.8: AAS data of OMS-2 and M-OMS-2 samples. Initial M/Mn ratio M (mol)a Mn (mol)a M/Mn ratio K (mol)a K/(Mn+M) OMS-2 0.0 0.0 552.2 0.0 38.8 0.07 Ti-OMS-2 (0.05) 0.05 26.81 505.97 0.05 38.7 0.07 Ti-OMS-2 (0.67) -a 228.2 342.0 0.67 21.5 0.04 Fe-OMS-2(0.09) 0.10 42.2 457.9 0.09 40.6 0.08 Cu-OMS-2(0.02) 0.10 9.9 528.0 0.02 41.2 0.08 Co-OMS-2(0.04) 0.10 20.7 518.4 0.04 38.3 0.07 a no MnSO4.H2O was added to the initial reaction, see Section 3.2.3, the amount of Ti2(SO4)3 solution was 75 mL. The morphology of M-OMS-2 samples was analyzed by FESEM and its micrographs are shown in Figure 4.10. It shows the morphology of Ti-OMS-2(0.05); 100 Fe-OMS-2(0.09); Co-OMS-2(0.04) and Cu-OMS-2(0.02). The figures show that the morphology of Ti- and Co-OMS-2 are similar to fibrous morphology of OMS-2. The widths of Ti-OMS-2 particles are relatively the same with OMS-2 in the ranges of 20-30 nm. For Co- and Cu-OMS-2, the widths of the particles increase to about 100 nm and the length remain relatively the same, which is about several hundred nm. The different morphology was observed in Fe- and Cu-OMS-2 samples. FeOMS-2(0.09) has a dense globular morphology with size of 100-200 nm, whereas Cu-OMS-2(0.04) has mixed fibrous and dense globular morphology. The difference of morphology of OMS-2 and M-OMS-2 dissagree to difference of relative intensity of the peaks after incorporation of metal ions as shown by XRD pattern of sample in Figure 4.7 and ratio of I(310)/I(211) plane in Table 4.5. It showed that Ti incorporated gave significant higher ratio of I(310)/I(211) while Fe and Cu incorporated gave lower ratio than OMS-2. Co has similar ratio with OMS-2. However, the different values are observed in calculation of ratio of I(200)/I(211) plane of OMS-2 and M-OMS-2 samples which listed in Table 4.9. The table shows that Ti has ratio of I(200)/I(211) plane very close to OMS-2. Conversely, ratio of I(200)/I(211) plane of Co-OMS-2 was significantly lower than ratio of OMS-2. Whereas ratio of I(200)/I(211) plane of Fe-OMS-2 and Cu-OMS-2 be in agreement to ratio of I(310)/I(211) plane which significantly lower than the ratio of OMS-2. This could be due to the very different morphology observed in both samples which both (200) and (310) plane relatively decreases compare to (211) plane. However, it could concluded that there are differences in ratio of intensity of the peaks of samples after incorporation of metal ions which indicated that difference of preferred orientations in OMS-2 crystal after incorporation of metal ions. Moreover, the exactly preferred plane which count their different morphology could not defined. 101 OMS-2 Ti-OMS-2(0.05) Cu-OMS-2(0.04) Fe-OMS-2(0.09) Co-OMS-2(0.02) Figure 4.10: Morphology of OMS-2 and M-OMS-2 samples. 102 Table 4.9: The relative intensity and ratio of I(200)/I(211) plane of OMS-2 and MOMS-2 samples calculated by XRD. 4.3 Samples 200 211 I(200)/I(211) OMS-2 88 100 0.88 Ti-OMS-2(0.05) 81 93 0.87 Fe-OMS-2(0.09) 54 100 0.54 Co-OMS-2(0.02) 76 100 0.76 Cu-OMS-2(0.04) 58 100 0.58 Physical Properties of H-OMS-2 and H-M-OMS-2 Materials Hydrogen form of OMS-2 and M-OMS-2 (H-OMS-2 and H-M-OMS-2, respectively) were prepared by ion exchange of potassium ion in the tunnel structure with H+ ions using concentrated HNO3 at 60 oC. This is a short and simple method to give a higher amount of substituted H+ compared to using aqueous HNO3. The synthesis of H-OMS-2 and H-M-OMS-2 are schematically shown in Figure 4.11. Figure 4.11: Schematic synthesis of H-OMS-2 and H-M-OMS-2. 103 The ion exchange was repeated twice in order to maximise the amount of K+ that may be exchanged by H+ ion. The XRD pattern of OMS-2 and H-OMS-2 are shown in Figure 4.12. These XRD patterns of OMS-2 and H-OMS-2 are similar indicating that the catalyst retains its framework structure after even successive ion exchange with H+. However, there are differences in the XRD pattern of OMS-2 and H-OMS-2 in terns of the peaks intensity, relative peaks height and the peaks width of the sample, reflecting the changes in the crystallinity, preferred orientation and crystallite size of the cryptomelane structure of OMS-2 after ion exchange. The amount of K+ that was exchanged by H+ was calculated by EDAX and is given in Table 4.10. It is observed that after ion exchange, the percentage of potassium substituted by H+ on H-OMS-2 is about 44%. This amount is less than 50% of that reported by Kumar et al. [70]. The percentage of potassium substituted by H+ on M-OMS-2 varies. The highest percentage is observed in H-Ti-OMS2(0.67) sample and the lowest in H-Fe-OMS-2(0.09) sample. The highest potassium exchanged by H+ is due to the small amount of K+ in Ti-OMS-2 (0.67) sample (see Table 4.8). Figure 4.12: XRD pattern of (a) OMS-2 and (b) H-OMS-2. 104 Table 4.10: The percentage of potassium substituted by H+. Samples %K was exchanged by H+ (a) H-OMS-2 44 H-Ti-OMS-2(0.67) 93 H-Fe-OMS-2(0.09) 12 H-Co-OMS-2(0.04) 37 H-Cu-OMS-2(0.02) (a) + 62 + %K exchanged with H was calculated based on the decrease in the wt% of potassium in EDAX data assuming that the amount of Mn does not change after the ion-exchange. The morphology of H-OMS-2, H-Ti-OMS-2(0.67), H-Fe-OMS-2(0.09), HCo-OMS-2(0.04) and H-Cu-OMS-2(0.02) samples are depicted in Figure 4.13. FESEM micrograph of H-OMS-2 and H-Co-OMS-2 reveal a fibrous needle-like morphology; similar to the morphology of coresponding parent samples (before ion exchanged). H-Fe-OMS-2(0.09) keeps a dense globular morphology of Fe-OMS2(0.09) and H-Cu-OMS-2(0.02) also has similar morphology to its unexchanged sample [Cu-OMS-2(0.02)]. These show that metal substituted samples, prepared by mixing of metals ion with manganese ion have similar morphology to their ionexchanged samples indicated that the tunnel cation substitution has no effect on the morphology of those samples. This is in agreement with the previous report by Kumar et al. [70]. The particle sizes of all H-M-OMS-2 samples are relatively similar to its M-OMS-2 samples. However, the difference was observed on the morphology of H-Ti-OMS-2(0.67) which shows dense globular morphology similar to Fe-OMS-2. Its morphology changed after ion-exchanged by H+ ions. 105 H-OMS-2 H-Ti-OMS-2(0.67) H-Fe-OMS-2(0.09) H-Co-OMS-2(0.04) H-Cu-OMS-2(0.02) Figure 4.13: Morphology of H-OMS-2 and H-M-OMS-2 samples. 106 4.4 Physicochemical Properties of Ti-OMS-2 Materials This section demonstrates in more detail the physicochemical properties of Ti-OMS-2 prepared by oxidation of titanium (III) sulfate solution in 15 % sulphuric acid by potassium permanganate. As a comparison, the extraframework TiO2-OMS2 was prepared by impregnation of titanium to the OMS-2. Physical mixing of OMS-2 and TiO2 rutile was also prepared. Table 4.12 summarizes the preparation method, the chemical composition and labeling of titanium containing OMS-2. 4.4.1 Structural Properties of Ti Substituted OMS-2 Catalyst XRD patterns of OMS-2, Ti-OMS-2 (0.18) and Ti-OMS-2 (0.43) show that the samples are pure and highly crystalline and match those of cryptomelane Q [168]; the natural counterpart of OMS-2 material (see Figure 4.14). The results confirm that OMS-2, Ti-OMS-2 (0.18) and Ti-OMS-2 (0.43) materials consist of the cryptomelane structure: 2 x 2 tunnels with a pore size of 4.6 Å, composed of double chains of edge-sharing and corner-sharing MnO6 octahedra [172]. The absence of other peaks in XRD patterns except the cryptomelane peaks suggests that Ti was successfully incorporated in the framework of Ti-OMS-2. In order to confirm the successful incorporation of titanium, the XRD patterns of Ti-OMS-2 was compared with mechanical mixture of TiO2 (rutile) and OMS-2, in which the mixture (TiO2OMS-2 (mix)) shows the presence of rutile phase (see Figure 4.13 and Table 4.11). If Ti were successfully incorporated in the framework of OMS-2, one expects that the bigger is the substituted atom, the bigger is the unit cell volume. Calculations of the unit cell volume of OMS-2, Ti-OMS-2 (0.18) and Ti-OMS-2 (0.43) as listed in Table 4.12 show that the unit cell volumes increase by the incorporation of Ti in the framework of OMS-2. Ti-OMS-2 (0.43) which has higher titanium content than Ti-OMS-2 (0.18), also has bigger cell volume than Ti-OMS-2 (0.18). This indicates that the amount of manganese in the framework structure which was substituted by titanium also increases. The lattice enlargement is due to 0.18 0.43 0.67 0.18 0.67 Ti-OMS-2 (0.18) Ti-OMS-2 (0.43) Ti-OMS-2 (0.67) TiO2-OMS-2 (imp) c TiO2-OMS-2 (mix) d 552.2 552.0 342.0 420.4 369.9 99.4 228.2 181.4 74.4 0.0 Molar amount of Tib mechanical mixing impregnation direct synthesis direct synthesis direct synthesis - Methods of introduction of Ti species non-framework non-framework framework and non-framework framework framework - Location of Ti species 152 - rutile amorphous rutile b n.d. e n.d. e 152 149 155 - - Surface area / m2g-1 Structure of nonframework Ti species Analysis was carried out by Atomic absorption spectrometer. The amount of OMS-2, Ti-OMS-2 and TiO2–OMS-2 was 50 mg. c Titanium(IV) tetra-2-propoxide (Ti(OPri4) was impregnated from its toluene solution into OMS-2 powder and calcined at 773 K for 3h. d Catalyst was prepared by addition of calculated amount of Ti from TiO2 powder to OMS-2. e Not determined. a 552.2 0.00 OMS-2 412.5 Molar amount of Mnb Molar ratio of Ti/Mna Samples Table 4.11 : Chemical composition and physicochemical properties of OMS-2, Ti-OMS-2 and TiO2–OMS-2. 107 108 * = TiO 2 (rutile) * * (g) * * (f) (e) * * * Relative intensity / a.u. * (d) (c) (b) (a) 10 20 30 40 50 60 70 2T/ o Figure 4.14: X-ray diffractograms of (a) cryptomelane (JCPDS 29, 102), (b) OMS-2, (c) Ti-OMS-2 (0.18), (d) Ti-OMS-2 (0.43), (e) Ti-OMS-2 (0.67), (f) TiO2-OMS-2 (imp) and (g) Ti-OMS-2 (mix). 109 the replacement of the smaller Mn4+ ions (ionic radius is 0.53Å) by the relatively larger Ti4+ ions (ionic radius is 0.61 Å). It strongly indicates that isomorphous substitution of Mn atoms by Ti into the framework of Ti-OMS-2 has occured. Moreover, it is found that the incorporation of titanium caused a decrease to the full width at half maximum, indicating an increase in the grain size. It is observed that the peaks of rutile phase for TiO2 appeared at the ratio of Ti:Mn higher than ca. 0.5. In contrast, it is not observed at the ratio of Ti:Mn less than ca. 0.5 (see Figure 4.14). By considering the upper limit of the titanium that could be incorporated into the framework, one would expect non-framework titanium species to be formed when the Ti:Mn ratio reached 0.5. The amount of Ti located in non-framework is 25% in Ti-OMS (0.67). This argument is supported by the presence of rutile phase of TiO2 in Ti-OMS-2 (0.67) (see Figure 4.14). However, no reflection for rutile phase of TiO2 was observed in TiO2-OMS-2 (imp) where the catalyst was prepared by impregnation method. This result implies that the structure of TiO2 in TiO2-OMS-2 (imp) is in the amorphous form. Table 4.12: The lattice parameters (a and c) and cell volume (V) of OMS-2 and TiOMS-2 samples. Samples a (Å) c (Å) V (Å3) OMS-2 9.74 2.86 271 Ti-OMS-2(0.18) 9.84 2.84 275 Ti-OMS-2(0.43) 9.83 2.87 277 Figures 4.15 and 4.16 show the FTIR spectra of the samples. The lower wavenumber region the five bands between 400 cm-1 - 800 cm-1 are assigned to MnO and Ti-O vibrations that are clearly seen (Figure 4.15) and the details of these peaks are summarised in Table 4.13. The characteristic cryptomelane peaks at 464, 522, 600 and 714 cm-1 are similar in OMS-2 and Ti-OMS-2(0.18) samples, which is in agreement with XRD analysis, indicating that both samples have pure cryptomelane structure. However, the intensity of vibration bands due to Mn-O lattice at OMS-2 is stronger than that Ti-OMS-2(0.18); indicating that titanium was 110 successfully incorporated into the framework structure of OMS-2. Figure 4.15 also shows that bands of Ti-OMS-2 (0.67) are significantly shifted to higher wavenumber and its intensity become lower than that of Ti-OMS-2(0.18) and OMS-2. The new band observed at around of 581 cm-1 is due to stretching of Ti-O-Ti [173] indicating the existence of TiO2 peak in Ti-OMS-2 (0.67) which agrees with XRD data. 1100 600 714 464 522 Transmittance / a.u. (c) (b) (a) 1200 1100 1000 900 800 700 600 500 400 Wavenumber / cm-1 Figure 4.15 : IR spectra at lower wavelength region of (a) OMS-2, (b) Ti-OMS-2 (0.18), (c) Ti-OMS-2 (0.67). The band around 714 cm1 which is assigned to the characteristic band of manganese oxides with tunnel structure was also observed. It was observed that in Ti-OMS-2(0.67), the corresponding band diminished indicating that there was 111 Table 4.13: Vibrational spectroscopy feature of samples. Wavelength (cm-1) Assignment 464 (O-Mn-O) 522 (Mn-O) 581 (O-Ti-O) 714 (Mn-O) 1600 (c) 3400 Transmittance / a.u. (b) (a) 4000 3000 2000 1500 1200 Wavenumber / cm-1 Figure 4.16: IR spectra at higher wavelength region of (a) OMS-2, (b) Ti-OMS-2 (0.18), (c) Ti-OMS-2 (0.67). 112 change in the tunnel structure of OMS-2 due to insertion of the titanium rutile phase. The bands at 1000- and 1150 cm1 were also observed for all samples, which were assigned to the vibration of the Mn3+-O bond [174]. This suggests that the oxidation state of Mn in the structural framework is not only four but also include Mn3+ occupying the Mn4+ sites in the crystal of the tunnel structure. Figure 4.16 shows the FTIR spectra of OMS-2 and Ti-OMS-2 at high wavenumber. The bands at 3400 cm-1 and 1600 cm-1 are due to vibration and stretching OH, respectively, designated to the presence of OH groups that probably belong to the M-OH as well as to the absorbed water in the sample. It shows that by incorporating titanium, the band at 3400 cm-1 and 1600 cm-1 are stronger than the original OMS-2 which indicates that more hydroxyl groups are in the titanium incorporated sample and are responsible to the surface of being more hydrophilic. The photoluminescence (PL) was also used to confirm the absence of nonframework TiO2, since non-framework TiOx with very small crystallite size cannot be detected by XRD. The PL spectra are useful to disclose the efficiency of charge carrier trapping, migration and transfer, and to understand the nature of electron-hole pairs in TiO2 semiconductor particles since PL emission results from the recombination of photo-excited free carriers [175]. In this study, the 430 nm excited PL spectra of all powder samples pressed at room temperature were examined in the range of 560–680 nm. The PL spectra of OMS-2, Ti-OMS-2(0.43) and TiO2-OMS-2 (mix) are shown in Figure 4.17. The results indicate that the photoluminescence intensity of TiO2-OMS-2 (mix) (em(max) ~600 ± 10 nm; FWHM ~40 nm) was substantially higher than that of OMS-2 and Ti-OMS-2 (0.43) (see Figure 4.17). The relative intensity of Ti-OMS-2 (0.43) and TiO2-OMS-2 (mix) is similar suggesting that there is no TiOx particle in Ti-OMS-2 (0.43) sample. This result further supports our suggestion that Ti was incorporated in the framework of Ti-OMS-2. The incorporation of Ti in the framework of OMS-2 was further supported by the surface area analysis (see Table 4.12), which revealed that the surface area of OMS-2 and TiOMS-2 is almost the same. 113 Ti-OMS-2 (0.43) Intensity / a.u. TiO2 -OMS-2 (mix) OMS-2 560 580 600 620 640 660 680 Wavelength / nm Figure 4.17: Photoluminescence spectra of OMS-2, Ti-OMS-2 (0.43) and TiO2OMS-2 (mix). The excitation wavelength is 430 nm. 4.4.2 Acidity Properties IR spectrum of acidity study by pyridine adsorption after evacuation under vacuum at 400 oC and 150 oC are shown in Figure 4.18. The figure shows that Lewis acid sites are formed in Ti-OMS-2 (0.67) as indicated by the appearance of peaks at 1447 cm-1, 1489 cm-1 and 1604 cm-1. In contrast, no peaks are observed for OMS-2 sample in Figure 4.18(b). The absence of peaks at 1540 cm-1 confirms that there are no Brønsted acid sites in both samples. The data indicates that the insertion of Ti into the framework of OMS-2 created Lewis acids in the sample. 114 1447 1604 (a) Absorbance / a.u. 1489 (b) 1700.0 1600.0 1500.0 Wavenumber / cm-1 1400.0 1300.0 Figure 4.18: FTIR spectra of (a) Ti-OMS-2 (0.67) and (b) OMS-2 after evacuation under vacuum at 400 oC for 4 h followed by pyridine adsorption at room temperature and evacuation at 150 oC for an hour. 4.4.3 Morphology, Surface Area and Textural Properties Figure 4.19 reveals the morphology of Ti-OMS-2 (0.18), Ti-OMS-2 (0.67), and TiO2-OMS-2 (imp) materials. As shown by Figure 4.19 (a) and (b), the incorporation of Ti changes the fibrous morphology of OMS-2 (see Figure 4.10) to spherical morphology and its particle size was reduced. The particle sizes are in the range of 50 - 80 nm in Ti-OMS-2 (0.18). The particle sizes of Ti-OMS-2 (0.67) are bigger than Ti-OMS-2 (0.18) which are in the range of 120 - 200 nm. The difference in morphology of OMS-2 and Ti-OMS-2 are due to preferred orientation of Ti-OMS2. The Ti-OMS-2 materials tend to lie on their (110) plane, as revealed by the XRD data (see Table 4.14). The table shows the relative intensity and the ratio of (110) planes to (211) planes of OMS-2, Ti-OMS-2 (0.18) and Ti-OMS-2 (0.67). The ratio of I(110)/I(211) of Ti-OMS-2 that was higher than OMS-2 indicates that Ti-OMS-2 is preferred in (110) plane than OMS-2. The higher ratio of I(110)/I(211) of Ti- 115 (a) (b) (c) Figure 4.19: Morphology of (a) Ti-OMS-2 (0.18), (b) Ti-OMS-2 (0.67) and (c) TiO2-OMS-2 (imp). OMS-2(0.18) than Ti-OMS-2(0.67) explaines the difference of the length of particle of both materials {see Figure 4.19 (a) and (b)}. Thus, the agglomeration of TiO2 rutile particle as identified by XRD is also observed in Ti-OMS-2(0.67). The preferred orientation was not observed in titanium impregnated materials. As shown by Figure 4.19 (c) the impregnation of Ti did not alter the fibrous morphology of OMS-2 but only showed that titania particle was finely dispersed on the fibrous morphology of OMS-2 materials. 116 Table 4.14: The relative intensity and ratio of I(111)/I(211) plane of samples calculated by XRD. Relative Intensity of (hkl) I(110)/I(211) (110) (211) OMS-2 56 100 0.56 Ti-OMS-2 (0.18) 88 100 0.88 Ti-OMS-2 (0.67) 84 100 0.84 The N2 adsorption isotherm of OMS-2 is shown in Figure 4.20. Type II adsorption isotherms with micropore filling at low P/Po and capillary condensation at high P/Po are observed on OMS-2 sample. At low P/Po, micropore filling of Type I adsorption isotherm occurs, indicating the existence of micropores in OMS-2. A hysteresis loop type H3 occurs at high P/Po close to saturation, suggesting the 600 adsorption 500 desorption Volume / cc g-1 400 300 200 100 0 0.0 0.2 0.4 0.6 0.8 P/Po Figure 4.20: N2 adsorption isotherm for OMS-2a at 77 K. 1.0 117 existence of slit-shaped mesopores with non-uniform sizes or shapes in OMS-2 [141]. In general the loop closes at P/Po0.8 indicating mesoporosity but probably due to pores that are generated by the spaces between the nanofibers of OMS-2 materials. The N2 adsorption isotherms of Ti-OMS-2(0.43) and Ti-OMS-2(0.67) are shown in Figure 4.21 (a) and (b). Adsorption isotherms of both Ti-OMS-2 samples are different from OMS-2. The type IV adsorption isotherm was observed in TiOMS-2 with micropore filling at low P/Po. Similar to OMS-2, at low P/Po, micropore filling of Type I adsorption isotherm occurs, indicating Ti-OMS-2 has micropores. However, the hysteresis loop type H2 observed in Ti-OMS-2 indicates the presence of agglomerates of spheroid particles with nonuniform sizes or shapes [141]. The surface area of samples are shown in Table 4.11. It shows that surface area of Ti-OMS-2 sample close to OMS-2 sample indicated that incorporation of Ti not significantly changes the surface area of OMS-2. 118 210 180 Volume / ccg-1 150 120 90 60 30 0 0.0 0.2 0.4 0.6 0.8 1.0 0.60 0.80 1.00 P/Po 200 180 160 Volume / cc g-1 140 adsorption desorption 120 100 80 60 40 20 0 0.00 0.20 0.40 P/Po Figure 4.21: N2 adsorption isotherm for (a) Ti-OMS-2 (0.43) and (b) Ti-OMS-2 (0.67) at 77 K. 119 4.4.4 Thermal Stability Figure 4.22a and 4.22b represent TGA profiles in original and differential forms recorded from OMS-2, Ti-OMS-2 (0.18) and TiO2-OMS-2 (imp). The thermal stability of the material changed after the incorporation of titanium. First weight loss occurred in the range temperature of 30–200 ºC which is attributed to the elimination of adsorbed water molecules from the samples. Ti-OMS-2 (0.18) has the highest weight loss (about 4.5%) in this stage indicating that more adsorbed water was present on surface, implying more hydrophilic properties of the sample. This is followed by OMS-2 and TiO2-OMS-2 (imp) with their weight losses of about 2.4% and 1.2%, respectively. The weight loss at temperature range of 200-350 ºC due to the evolution of oxygen from the lattice was observed in OMS-2 and Ti-OMS-2 (0.18) samples. The weight loss of OMS-2 and Ti-OMS-2 (0.18) samples are 2.0 and 2.5 %, respectively. It shows that the start point in OMS-2 sample was at lower temperature compared to Ti-OMS-2 (0.18), indicating that incorporation of titanium caused evolution of oxygen from the lattice more difficult. However, such weight loss was not clearly observed in TiO2-OMS-2(imp) sample indicating that oxygen evolution from the lattice was blocked by impreganation of titanium on the surface of OMS-2 sample. Figure 4.22b also shows significant weight loss of 9.4% from OMS-2 was observed in the temperature range of 520-800 ºC and contributed to the collapse of OMS-2 structure. Weight loss of Ti-OMS-2(0.18) and TiO2-OMS-2(imp) in this stage are 7.2% and 6.6%, respectively. It shows that in both samples, the start point shifts to higher temperature indicating that the thermal stability of samples increased either upon incorporation or impregnation of titanium. OMS-2 lost its weight at two 120 Figure 4.22: TGA profile (a) in original and (b) differential forms of OMS-2, TiOMS-2 (0.18) and TiO2-OMS-2(imp). 121 parts i.e. at around 585 and 780 ºC due to the decomposition of this material into stable lower valent manganese oxides, Mn3O4 (hausmannite) and Mn2O3 (bixbyite), respectively [30]. Ti-OMS-2(0.18) has about three parts of weight loss in this stage which indicates that the sample decomposed into three phases. The additional phase could be MnTiO3 (phyrophanite) which is also observed in Mn-TiO2 [176]. The addition of decomposition process suggests the existence of titanium incorporated in the framework since the presence of extraframework titanium in TiO2-OMS-2 (imp) sample had no additional weight loss. The dispersion of TiO2 on TiO2-OMS-2 (imp) also caused the slight increase of the thermal stability of OMS-2 in first decomposition to hausmannite and significant increase in decomposition to bixbyite but no other weight loss was observed. 4.4.5 More Evidence of the Location of Titanium on Ti-OMS-2 Materials XPS analyses were carried out in order to obtain information on the local chemical environments of manganese and titanium in cryptomelane from the variations in binding energies or chemical shifts of the photoelectron lines. Because chemical shifts are very uniform among the photoelectron lines of an element, line separations rarely vary by more than 0.2 eV [36]. A change in the separation of photoelectron lines suggests that a change in the local environment may has occured. Line separation of binding energy of Mn 2p and Ti 2p and its differences between selected samples are listed in Table 4.15. Mn 2p XPS spectra of OMS-2, Ti-OMS-2 (0.18), Ti-OMS-2 (0.67) and TiO2-OMS-2 (imp) are shown in Figure 4.24. The binding energies for Mn 2p1/2 and Mn 2p3/2 peaks of all samples were about 654 and 642 eV, which are very close to those of MnO2 Mn 2p peaks. The line separation for Mn 2p1/2 and Mn 2p3/2 peaks BE of OMS-2, Ti-OMS-2 (0.18), Ti-OMS-2 (0.67) and TiO2-OMS-2 (imp) were 11.83, 11.82, 11.91 and 11.70 eV, respectively. Difference of line separation (BE) of Ti-OMS-2 (0.18) and BE of OMS-2 was 0.01 eV, which very close to each other indicating that there is no 653.73 653.88 654.05 Ti-OMS2(0.18) Ti-OMS2(0.67) TiO2-OMS2(imp) 642.35 641.97 641.91 641.75 BE Mn 2p3/2 (eV) 0.01 0.08 0.13 11.91 11.70 Difference of BE Mn 2p of X and OMS-2 (eV) 11.82 11.83 BE Mn 2p (eV) 464.01 463.56 462.89 BE Ti 2p1/2 (eV) X refers to Ti-OMS-2(0.18), Ti-OMS-2(0.67) and TiO2-OMS-2(imp), respectively. Y refers to Ti-OMS-2(0.67) and TiO2-OMS-2(imp), respectively. 653.58 BE Mn 2p1/2 (eV) OMS-2 samples. 458.17 457.92 457.53 BE Ti 2p3/2 (eV) 5.84 5.64 5.36 BE Ti 2p (eV) 0.48 0.28 Difference of BE Ti 2p of Y and TiOMS-2 (0.18) (eV) Table 4.15: Binding Energies (eV) of Mn 2p, Ti 2p, and its line separation (BE) and difference of line separation from selected 122 123 2p1/2 2p3/2 (d) Intensity / a.u. (c) (b) (a) 658 656 654 652 650 648 646 644 642 640 638 Binding Energy / eV Figure 4.23: Detailed XPS spectra for the Mn 2p transition for (a) OMS-2, (b) TiOMS-2 (0.18), (c) Ti-OMS-2 (0.67) and (d) TiO2-OMS-2 (imp). change of the local environment of OMS-2 after incorporation of titanium. For other samples i.e Ti-OMS-2 (0.67) and Ti-OMS-2 (imp) the difference of line separation was also lower than 0.2 indicating that the changing of local environment Mn 2p could not observed upon the addition of titanium. This is in agreement with results by Cai et al. [36] which reported that the local environment of Mn 2p of Fe-OMS-2 124 did not show the significant change although the hematite phase was observed in that sample by using XRD analysis. However, the binding energies of Mn 2p of OMS-2 are lower than Ti-OMS-2 (O.18) (see Figure 4.23). This confirmed the substitution of some manganese by titanium in the framework of OMS-2. The assumption is supported by considering the binding strength on bridging oxygen atom in OMS-2 and Ti-OMS-2 as shown in Figure 4.24. Structures A, B, C is pure OMS-2, Ti incorporated OMS-2 and TiO2, respectively. The binding strength of (A) which is lower than (B) indicates that (A) is less stable than (B), and consequently the binding energy of Mn in pure OMS-2 is lower than in Ti-OMS-2 sample. This result is supported by TG analysis (section 4.4.4) which showed that incorporation of titanium into the framework of OMS-2 material caused an increase in thermal stability of OMS-2 material. As shown in Table 4.15 the separation of titanium 2p1/2 and 2p3/2 peaks BE increased from 5.36 to 5.64 eV (BE = 0.28) when the titanium doping ratio was increased from 0.18 to 0.67. This indicates that local environment of titanium in TiOMS-2(0.18) is different from Ti-OMS-2(0.67). The binding energy Ti 2p of samples increases in the order of: Ti-OMS-2(imp) < Ti-OMS-2(0.67) < TiO2-OMS2(0.18) (See Figure 4.25) due to the difference in the local environment of each sample. The lowest binding energy of Ti-OMS-2(0.18) caused by incorporation of titanium in the framework of OMS-2 (Structure B) which has total bond strength lower than Structure C (see Figure 4.24). The highest binding energy calculated for TiO2-OMS-2 (imp) is due to the Ti-O-Ti bond (Structure C) that has the highest total bond strength, thus indicating that extraframework titanium exists in the sample. The middle of binding energy of Ti-OMS-2(0.67) has a significant difference of BE (more than 0.2 eV) indicating that it has both Structures B and C (titanium framework-extraframework exists in the sample) which agrees with the previous statements. The oxidation state of the Ti species was also examined by X-ray photoelectron spectroscopy. Figure 4.25 shows the XPS of the materials in the Ti 2p1/2 and Ti 2p3/2 binding energy (BE) regions. The Ti 2p3/2 peak is centered at 125 457.53-458.17 eV, and the Ti 2p1/2 peak is found at 462.89-464.01 eV, with a spin energy separation of 5.36-5.84 eV. This is characteristic of Ti4+ [177, 178]. In particular, the peak position of Ti4+ is significantly influenced by its coordination environment, and the Ti 2p3/2 XPS peak of the octahedral coordinated Ti4+ generally locates at lower BE (about 457.5-458.3 eV), while that of tetrahedral coordinated Ti4+ at higher BE (about 458.5–463.4 eV) [179]. The observed BE Ti 2p3/2 values approach to those observed for the octahedral coordinated Ti4+, indicating that the Ti species in Ti-OMS-2 materials are located at the octahedral coordinated environments. Configuration Total bond strength (A) 1.27 (B) 1.30 (C) 1.33 Figure 4.24: Bond strength on bridging oxygen atom. 126 2p3/2 2p1/2 Intensity / a.u. (c) (b) (a) 468 466 464 462 460 458 456 454 Binding Energy / eV Figure 4.25: Detailed XPS spectra for the Ti 2p transition for (a) Ti-OMS-2(0.18), (b) Ti-OMS-2(0.67) and (c) TiO2-OMS-2(imp). 127 4.5 Alkylsilylated of OMS-2 and Ti-OMS-2(0.67) The effect of modification by alkylsilylation on the surface of catalysts and their catalytic activity in styrene oxidation studied by attachment of OTS on the surface of OMS-2 (OTS/OMS-2) and Ti-OMS-2(0.67) (OTS/Ti-OMS-2) is depicted in Figure 4.26. The peaks at 2920 and 2850 cm-1 are due to anti-symmetric and symmetric stretching of C-H, respectively. This indicates that OTS was successfully attached on the surface of the catalysts. In order to determine the relative hydrophobic and hydrophilic properties of the samples, water vapour adsorption on the surface of samples was applied. As shown in Figure 4.27, it is clearly observed that the amount of adsorbed water on TiOMS-2(0.67) was higher than OMS-2. This suggests that the surface properties of Ti-OMS-2 is more hydrophilic than OMS-2. Contrary to the creation of the more hydrophobic surface by attachment of OTS on the surface of catalyst [180-182], the modification of OMS-2 and Ti-OMS-2 surfaces by OTS generate the more hydrophilic surface. Actually, attachment of bimodal amphiphilic OTS on the surface of catalyst increased both hydrophilic and hydrophobic groups on the surface of catalysts. However, in this case, adsorption of water vapour on the modified surface was higher than unmodified one (see Figure 4.27). The proposed polymeric model on the surface is demonstrated in Figure 4.28 [183]. As shown in Figure 4.27, the higher absorbed water vapour of Ti-OMS-2 than OMS-2 indicates that there are more OH groups on Ti-OMS-2 than on OMS-2. Higher amount of OH, is preferred in cross-linked polymeric octadecylsiloxane resulted less OH in alkylsilylated TiOMS-2 sample. OMS-2 surface which has less OH group favoured to be in the form of isolated polymeric rather than cross-linked octadecylsiloxane after the alkysililation modification on the OMS-2 surface which could create more OH groups and adsorb more water vapour. 128 Transmittance / a.u. (b) (a) 2850 2920 4000.0 3000 2000 1500 Wavenumber / cm-1 1000 400.0 Figure 4.26: FTIR spectra of modified OTS samples (a) OTS/OMS-2 and (b) OTS/Ti-OMS-2. 14 12 adsorbed water / % (d) 10 (c) 8 6 (b) 4 (a) 2 0 0 2 4 Time / h 6 Figure 4.27: Percentage of adsorbed water on the sample (a) OMS-2, (b) Ti-OMS-2 (0.67), (c) OTS/Ti-OMS-2 (0.67) and (d) OTS/OMS-2. 8 129 Figure 4.28: Proposed polymeric octadecylsiloxane on the surface of OMS-2 and TiOMS-2 samples. 4.6 Sulphated Ti-OMS-2 XRD patterns of Ti-OMS-2 and sulphated Ti-OMS-2 catalysts prepared in water and toluene as solvent are demonstrated in Figure 4.29. Samples prepared with 150 and 200 μL of H2SO4, using toluene as solvent are referred to ST150-Ti-OMS-2 and ST200-Ti-OMS-2. Similarly, SW150-Ti-OMS-2 and SW200-Ti-OMS-2 refer to the samples that were prepared with 150 and 200 μL of H2SO4, respectively, but using water as solvent. From the figure it is observed that ST150-Ti-OMS-2 and SW150-Ti-OMS-2 samples have similar XRD patterns with Ti-OMS-2 sample and no new peaks are observed in both samples. This indicates that the sulfate group was isomorphously attached to cryptomelane structure of Ti-OMS-2 materials due to the low concentration of sulfate. In contrary, ST200-Ti-OMS-2 and SW200-Ti-OMS-2 samples also have cryptomelane peaks but there are new peaks observed in those samples. The new peaks matched those of TiOSO4.H2O, MnSO4 7H2O, MnSO4 peaks by XRD database patterns. The peaks intensity of sulphated sample decreased with the increasing amount of H2SO4 added. It indicates that at higher sulphate content caused a decrease in crystallinity of the sample. It could be due to phase transition of OMS-2 at high sulphate content to the other phases such as TiOSO4.H2O, MnSO4 7H2O and MnSO4 which was observed in Figure 4.29. 130 * # * SW200-Ti-OMS-2 # ¤ ¤ ST200-Ti-OMS-2 Intensity / a.u. # SW150-Ti-OMS-2 ST150-Ti-OMS-2 Ti-OMS-2 5 10 20 30 40 50 60 70 2T / º Figure 4.29: XRD pattern of Ti-OMS-2(0.04) and sulphated Ti-OMS-2 (0.04). # = TiOSO4.H2O,* = MnSO4 7H2O ¤ = MnSO4. Ti-OMS-2 and sulphated Ti-OMS-2 were studied by IR spectroscopy (Figure 4.30). Sulphated Ti-OMS-2 shows characteristic adsorption bands at 980–1230 cm-1, which are assigned to the bidentate sulfate coordinated to metal elements [113, 184, 185]. The details of the bands of sulphated samples are summarized in Table 4.16. The broad bands around 2800–3600 cm-1 may be attributed to surface silanols and adsorbed water molecules, while deformational vibrations of adsorbed molecules caused the adsorption bands at 1620 cm-1 [186]. A weak band at 1230 cm-1 corresponding to symmetric vibration of S=O was observed, but the asymmetric Transmittance / a.u. 131 986 1230 1180 1060 1120 700 4000.0 3000 2000 1500 1000 400.0 Wavenumber / cm-1 Figure 4.30: FTIR spectra of Ti-OMS-2 and sulphated Ti-OMS-2. Table 4.16: Assignments of as-observed IR bands on sulphated samples [187]. IR bands (cm-1) assignment species 2800-3600, br OH H-bonded OH and H2O 1620, rw HOH H-bonded H2O 1400, s (-) asS=O (SO3)ads 1370, s (-) asS=O (H2SO4)ads 1322, sh (-) asS=O (HSO4)ads 1230, w s S=O (H2SO4)ads 1205, sh S-OH (HSO4)ads 1186, sh S-OH (HSO4)ads 1060, s sS-O (H2SO4)ads 980, s asS-O (SO3)ads br = broad, rw = rather weak, s = strong, sh = shoulder, w = weak, (-) the peaks was not observed in our samples, as = asymmetric, s = symmetric. 132 vibration of S=O between 1320-1400 cm-1 [187-190] was absent in our sulphated samples. The S-OH bending vibration expected near 1180 cm-1 on sulphated samples was observed due to the strong hydrogen bonding to the surface [187]. The bands appearing at 1060 and 980 cm-1 are characteristic of S–O symmetric and asymmetric vibrations, respectively [187, 191, 192]. All bands are clearly observed in all sulphated samples except SW150-Ti-OMS-2 which may be caused by the lower sulfate content in this sample. As seen in Figure 4.30 the bands of ST samples are more intense than SW sample which indicates that the amount of sulfate anchored in sample ST is higher than those of SW. The bands around 700 cm-1 which correspond to the tunnel structures of OMS-2 materials are decrease after sulfation with 200 mL H2SO4, due to interference of the other phases in the tunnel. Based on the FTIR data, the proposed structure of sulphated titanium in TiOMS-2 present in the bridging of bidentated structure is represented in Figure 4.31 [113]. According to this model, the strength of Lewis acid sites Ti4+ is increased by the inductive effect of the sulfate groups in the complex. The model also demonstrates that it is possible to convert Lewis acid sites into Brønsted acid sites by water adsorption [193]. Figure 4.31: The bridging of bidentated structure of sulphated Ti-OMS-2. In Figures 4.32 and 4.33 the thermograms of TG-DTA analysis and the first derivative curves, respectively can be seen for Ti-OMS-2 and sulphated Ti-OMS-2. Below 200 ºC a loss of mass can be seen due to the removal of water (in hydration or structural). In the temperature range of 280–340 °C, the oxidative decomposition of residual organic compounds occured. A weight decrease at 625–900 ºC was caused by the decomposition of sulfate on the surface to form SO2 [192]. The loss of sulfate in SW sample occurred at a lower temperature (a peak in the first derivative curve at 133 660 ºC) than in ST sample (peak at >800 ºC). Peaks at 600 and 820 ºC are due to decomposition of Ti-OMS-2 materials, also observed for Ti-OMS-2 sample. However, no peak at around 820 was observed for SW sample may be caused the decomposition of Ti-OMS-2 in SW sample occurs at higher temperature. 95 Ti-OMS-2 Weight / % SW150-Ti-OMS-2 ST150-Ti-OMS-2 90 85 80 100 200 300 400 500 600 700 800 Temperature / ºC Figure 4.32: Thermograms (TGA) of samples. Ti-OMS-2 SW150-Ti-OMS-2 ST150-Ti-OMS-2 mgmin-1 0.2 0.0 -0.2 -0.2 10 20 30 40 50 60 500 600 80 -0.4 100 200 300 400 700 Temperature / ºC Figure 4.33: First derivative curves (DTA) of samples. 800 min 134 The effect of sulfation on morphology of sample was also observed by FESEM. The micrograph of FESEM of SW150-Ti-OMS-2(0.67) is shown Figure 4.34. It shows that agglomeration occurred on the sample after sulphation. The morphology of sample became globular with particle size of about 100 nm. The morphology is similar to the image observed for Ti-OMS-2 (0.67) after ion exchange by H+ {H-TiOMS-2(0.67) sample}. Figure 4.34: FESEM micrograph of SW150-Ti-OMS-2(0.67). CHAPTER 5 CATALYTIC ACTIVITY OF OMS-2 AND MODIFIED OMS-2 SAMPLE IN OXIDATION AND ACID REACTIONS 5.1 Introduction This chapter describes the relationship between catalytic behaviour and their physicochemical properties of OMS-2 and modified sample OMS-2 in oxidation and consecutive oxidative-acidic reactions. The reactions were carried out in oxidation of alcohol, alkanes and alkenes and also consecutive oxidation and acid reactions for direct transformation of alkenes to diols. 5.2 Catalytic Activity and Selectivity of OMS-2 and Modified OMS-2 Samples in Oxidation Reactions 5.2.1 Oxidation of Benzyl Alcohol over OMS-2 Prepared by Different Method Oxidation of benzyl alcohol was used to study the effect of various preparation of OMS-2 to their catalytic activity on oxidation of benzyl alcohol. In this work, OMS-2 samples were prepared in acidic conditions. However, the different synthetic methods in acidic condition i.e. with and without buffer solution was used to study the effect of different methods in the oxidation of benzyl alcohol in order to obtained the best performance of OMS-2 material. The reaction of benzyl 136 alcohol to benzaldehyde was carried out at 110 oC with toluene and air as the solvent and oxidant, respectively, as shown in Figure 5.1. The selection of acidic method to preparation of OMS-2 based on report by Makwana et al. [23]. They studied the catalytic activity of prepared OMS-2 in oxidation of alcohol in different preparation conditions namely acidic, neutral and basic. They found that prepared OMS-2 material in acidic conditions is the most active in oxidation of alcohol. OH O Catalyst Toluene, 110 oC in air benzyl alcohol benzaldehyde Figure 5.1: The schematic reaction of benzyl alcohol to benzaldehyde. Table 5.1 shows conversion and selectivity of OMS-2a, OMS-2b catalysts and blank reaction. The blank reaction is reaction that was carried out without catalyst where all reaction conditions are keeping the same as by using catalyst. No product was observed using blank reaction as expected. The table also shows that OMS-2a gave a significantly higher conversion (90%) compared to only 50% for OMS-2b. Both catalysts gave 100% selectivity, indicating that the catalysts were selective to benzaldehyde. There is no further oxidized product such as benzoic acid was observed in the reaction, as normally observed in the other systems via radical Table 5.1: Conversion of benzyl alcohol and selectivity to benzaldehyde by different catalysts. Catalysts Conversion / % Selectivity / % Blank 0 0 OMS-2a 90 100 OMS-2b 50 100 OTS/OMS-2b 34 100 chain reaction. The selective product may be due to the Mars van Krevelen mechanism pathway of the reaction of benzyl alcohol with molecular oxygen on OMS-2 catalyst. 137 Most of oxidation processes with oxygen as the oxidant were based on the radical chain autooxidation reaction. Unstable peroxide by the reaction of dioxygen with free radicals results in low selectivity of the products was formed. Another catalytic oxygen activation procedure is to use a metal catalyst and an oxygen donor to form an oxometal or a peroxometal intermediate. In the peroxometal route, the oxidation state of the metal ion remains unchanged and the metal ion simply acts as Lewis acid, increasing the oxidizing power of the peroxo group. The oxometal pathway, on the other hand, involves a two-electron reduction of the metal ion, which is subsequently reoxidized by the oxygen donor. The latter pathway is known as the Mars–van Krevelen mechanism [194] and is more commonly observed in gasphase oxidations than in liquid phase. Investigation on oxidation of alcohol with molecular oxygen on OMS-2 as catalyst by kinetic and isotope labelling study proved that reaction occurred via Mars van Krevelen mechanism [24]. The Mars van Krevelen mechanism involves a two-step mechanism; firstly, oxidation of benzyl alcohol and the metal undergoes a two-electron reduction. Mn4+ which exists in OMS-2 materials acts as an oxidant to oxidize benzyl alcohol to benzaldehyde via abstraction of the secondary H atom point to an electron-deficient carbon centre in the intermediate formed in the rate-determining step and itself was reduced to Mn2+. TPR studies on OMS-2 [195] show a much higher susceptibility to reduction, as a reduction of Mn4+ is involved. This leads to easy release of oxygen species from the lattice. The next step is the fast step that is the delocalization of hydrogen atom in the intermediate formed to produce benzaldehyde. This process is summarized as: PhCH2OH + Mn4+ slow fast + PhCHOH + H+ + Mn2+ PhCHO + H+ The second step is the regeneration of catalyst; Mn2+ reoxidation by oxygen from air. There are three possible ways for an oxygen molecule to exchange its atoms on an oxide which can be investigated with the oxygen isotopic exchange reaction [196]: 138 1. R0-mechanism: The oxygen molecule exchanges one of its atoms with another oxygen molecule from the gas phase, without the participation of the oxygen of the metal oxides. 18 O2(g) + 16O2(g) 2 18O16O(g) 2. R1-mechanism: The oxygen molecule exchanges one of its atoms with the surface oxygen of the metal oxide. 18 O2(g) + 16O(s) 18O16O(g) + 18O(s) 3. R2-mechanism: the oxygen molecule exchanges both its atoms with the surface oxygen of the metal oxide. 18 O2(g) + 2 16O(s) 16O2(g) + 2 18O(s) The R0 mechanism is a pseudo-Langmuir-Hinshelwood-type mechanism in which the lattice oxygen is not involved and the exchange takes place between two adsorbed molecules. The R1 and R2 mechanisms are Mars–van Krevelen-type mechanisms, in which one or two oxygens, respectively, from the air are exchanged with lattice oxygen. The oxygen exchange occurs as: where denotes a lattice vacancy and denotes within the lattice. The molecular oxygen undergoes a two-electron reduction to form H2O2, which is known to decompose easily in another catalytic cycle over OMS-2 to form water as the final product [45]. There is no possibility of H2O2 to act as oxidant for this reaction. The data of Doornkamp et al. [196] suggests that much higher temperatures were required to achieve a state of the catalyst such that the R0 mechanism can prevail. The relatively low temperature employed to carry out the oxidation (110 ºC) and the oxygen-labelling experiment by Makwana et al. [24] conclusively pointed to a Mars–van Krevelen mechanism. Since an oxygen exchange was involved, the activity of the catalyst was inversely dependent on the Mn–O bond strength in the OMS-2. The proposed reaction scheme incorporating Mars–van Krevelen-mechanism is shown in Figure 5.2. 139 TGA data confirmed that oxygen species from the lattice OMS-2a were easier to release from the lattice than OMS-2b which produced more active OMS-2a than OMS-2b in the oxidation of benzyl alcohol using air as oxidant. This also suggests that the strength of Mn-O bond in OMS-2a is weaker than OMS-2b. Figure 5.2: Overall alcohol oxidation mechanism [24]. The effect of crystallinity on the oxygen release in oxidation of benzyl alcohol is described by Figure 5.3. The crystalline sample implies the maximising of the resonance energy [197]. As a consequence, the Mn-O bonds are strengthened which caused the difficulty of oxygen release from the lattice. This situation is depicted as Structure A. In less crystalline structure, the resonance effects are weaker which induces the easy breakage of Mn-O bonds to release oxygen from the lattice as shown in Structure B. From the XRD pattern it shows that OMS-2b is more crystalline than OMS-2a. This supports the fact that OMS-2a is more active than OMS- 2b due to the easier oxygen lattice release in OMS-2a than OMS-2b caused by the lower crystallinity of OMS-2a. The interaction of substrate and catalyst is important in catalysis process. The easier access of substrate to the surface/active site of the catalyst can affect the catalytic activity of the catalyst. From FTIR analysis it is proven that OH group of the surface of OMS-2a is greater than OMS-2b. Since the functional group of benzyl alcohol is OH as a consequent this site is easier to approach the surface/active site of 140 OMS-2a than OMS-2b which could enhance the catalytic activity of OMS-2a sample. The effect of more accessibility of benzyl alcohol to the surface is also observed after attachment of OTS on the surface of OMS-2a material. The conversion of benzyl alcohol decreased dramatically from 90% to 34% after alkylsilylation (see Table 5.1). This could be due to the blocking of active sites after modification. Figure 5.3: Resonance model of Mn-O-Mn bond structure: Resonance structure in crystalline OMS-2 (Structure A); and non-resonance structure in amorphous materials (Structure B). 5.3.2 Oxidation of Cyclohexane over Metals Substituted and Ion Exchanged OMS-2 Oxidation of cyclohexane with TBHP under solvent free condition was carried out by using OMS-2, M-OMS-2 H-OMS-2 and H-M-OMS-2 as catalysts. The most active OMS-2 which was tested in the oxidation of benzyl alcohol that is OMS-2a (without buffer) and further labelled for that catalyst was OMS-2 only. Those catalysts were studied for the effect of metal and H-exchanged on the catalytic activity of modified OMS-2 in oxidation of cyclohexane. The schematic of the reaction and its products was detected on those catalysts as depicted in Figure 5.4. Thus, the use of solventless condition during the reaction is the best choice to minimize the step and cost of the process. 141 O OH + Catalyst TBHP, 60 oC cyclohexane cyclohexanone + cyclohexanol OH O cyclohexyl hydroperoxide Figure 5.4: Schematic reaction of cyclohexane. Figure 5.5 demonstrates the conversion and selectivity of OMS-2, M-OMS-2, HOMS-2 and H-M-OMS catalysts in oxidation of cyclohexane. The main products observed after the reaction; detected by GC and GC-MS are cyclohexanol and cyclohexanone. Beside those products, the cyclohexyl hydroperoxide as by product was also observed with low selectivity in the range of 8-11%. Cyclohexyl hydroperoxide was an intermediate which was converted to cyclohexanol and cyclohexanone [160, 198, 199]. Figure 5.5 also shows that in blank reaction condition only cyclohexyl hydroperoxide was detected as product. The absence of cyclohexanol and cyclohexanone in blank reaction indicates that catalyst was needed to decompose cyclohexyl hydroperoxide to cyclohexanol and cyclohexanone. As shown in Figure 5.5, no overoxidation products such as n-hexanal and adipic acid or formation of cyclohexene as reported were observed. This indicates that the catalysts were selective to cyclohexanol and cyclohexanone. The conversion and selectivity to cyclohexanol and cyclohexanone are in the range of 4.4 - 7.8 and 89 - 92%, respectively which are higher than that of industrial process of about 4% and 7585%, respectively [154]. Conversion / % 0 1 2 3 4 5 6 7 8 34 54 11 35 55 11 34 57 9 36 56 9 32 61 8 32 57 11 34 57 9 32 59 9 37 55 8 32 60 8 aqueous TBHP (10 mmol), and catalyst (50 mg) under reflux condition. OMS-2, M-OMS-2 and H-M-OMS-2. All reactions were carried out at 60 ºC for 24 h with cyclohexane (26 mmol), 70% Figure 5.5: The conversion and product selectivity of oxidation of cyclohexane with tert-butyl hydroperoxide (TBHP) using 100 40 50 10 cyclohexanone cyclohexanol cyclohexyl hydroperoxide Product selectivity (%): 142 143 As shown in Figure 5.5 conversion of cyclohexane on OMS-2 catalyst was lower than both metal substituted and ion-exchanged catalysts. The figure shows that the modification of OMS-2 by metal substitution (metals = Ti, Fe, Cu, Co) and H-exchanged increased the activity of that material in oxidation of cyclohexane. The conversion of cyclohexane on OMS-2, Ti-OMS-2(0.09), Fe-OMS-2, Cu-OMS-2, CoOMS-2 and Ti-OMS-2(0.67) are 4.4; 5.1; 5.9; 6.1 and 7.0 %, respectively. The highest conversion of cyclohexane was obtained by Ti-OMS-2(0.67). It suggested that the presence of non-framework TiO2 in the sample (as observed in XRD analysis) increases conversion of cyclohexane. In a series of metal incorporated framework of OMS-2 catalysts, the conversion of cyclohexane increased in the following order: Ti- < Fe- <Cu- <CoOMS-2. The increase of ionic radii of metal substituted OMS-2 increased the conversion of cyclohexane (Figure 5.6). It was expected that larger ionic radii of metals resulted in turn strain of M-O bond. The strain on the Mn–O bond caused positively charged metal, increased Lewis acid site which was proven by pyridine adsorption analysis on Ti-OMS-2 sample. This Lewis active site plays a role in the oxidation reaction through three general ways as discussed in Section 5.3. In a series of H-M-OMS-2 catalysts the conversion of cyclohexane increases by the following order: H-OMS-2 < H-Ti-OMS-2(0.05) < H-Fe-OMS-2(0.09) < HCu-OMS-2(0.02) < H-Co-OMS-2(0.04) < H-Ti-OMS-2(0.67). The similar trend demonstrated by M-OMS-2 catalysts, confirmed that H exchanged M-OMS-2 materials enhanced the catalytic activity of these catalysts in the oxidation of cyclohexane. The increase in conversion of cyclohexane after ion exchange is in the range of 0.5 – 1.1% (see Table 5.2). 144 Conversion of cyclohexane / % 7 Co Cu 6 Fe Ti 5 4 0.60 0.65 0.70 0.75 0.80 Ionic radii of metals in framework OMS-2/ Å Figure 5.6: The relationship of ionic radii of metals substituted OMS-2 to conversion of cyclohexane. Table 5.2: The relation of amount of potassium exchanged by H+ with enhancement of % conversion of cyclohexane on H-M-OMS-2 catalyst. Catalysts % K exchanged by H+ Enhancement of % conversion after ion exchange OMS-2 44 1.03 Fe-OMS-2(0.09) 12 1.11 Cu-OMS-2(0.02) 62 0.71 Co-OMS-2(0.04) 37 0.64 Ti-OMS-2(0.67) 93 0.56 There are two possibilities of the effect of H-exchanged on OMS-2 and MOMS-2 catalysts. The first possibility is the presence of Brönsted acid in H-OMS-2 catalysts as reported by Kumar et al. [70]. It was reported that Brönsted acid on HOMS-2 played a role in acid-catalyzed condensation of phenylhydroxylamine with aniline to produce 2-aminodiphenylamine which did not occur on OMS-2 sample. However, the existence of Brönsted acid site at H-OMS-2 sample was not verified by 145 suitable characterization method such as pyridine adsorption or H-NMR. H-NMR study on H-OMS-2 and H-Ti-OMS-2 sample did not show the existence of Brönsted acid site. It was also supported by the fact that there was no correlation between the amounts of H+ in H-M-OMS-2 with the enhancement of conversion of cyclohexane after ion exchange as shown in Table 5.2. This indicates that Brönsted acids did not play a role in the enhancement of conversion of cycylohexane. Secondly, substitution of potassium by H+ may affect the Mn-O bond strength in the framework structure of OMS-2 and M-OMS-2 catalysts. Post et al. [200] reported that tetragonal cryptomelane may distorted to a monoclinic geometry when the ratio of the lattice Mn ionic radius over the tunnel cationic radius exceeded a value of 0.48. Since the ionic radius of H+ is much smaller than that of K+ the cryptomelane lattice is distorted, this in turn strains the Mn–O bonds. The tetragonal to monoclinic transition was also observed in sample OMS-2 with the lower average oxidation state (AOS), which indicated that the number of Mn3+ sites was higher [23]. The larger ionic radius of Mn3+ compared with Mn4+, fit better in a monoclinic environment. Kumar et al. [70] reported that increasing the amount %H+ in the tunnel decreased the average oxidation state (AOS) of Mn which indicated increasing number of Mn3+ sites on H-OMS-2 samples. Based on that, we concluded that the effect of H-exchanged catalysts on the enhancement of cyclohexane conversion was due to the reduction of bond strength of Mn-O. The strain on the Mn–O bond made the oxygen more negatively charged while the metal more positively charged, causing an increase to the strength of Lewis acid site; thus playing a role to enhance the catalytic activity of catalysts. In order to study the distribution of the product during the reaction process, the sample was analyzed regularly in certain time interval. Figure 5.7 shows the distribution of products vs time on Ti-OMS-2(0.67) catalyst. It was observed that at the first hour of reaction the amount of cyclohexyl hydroperoxide was higher compared to cyclohexanol and cyclohexanone. The amount of cyclohexyl hydroperoxide was relatively the same with cyclohexanol and cyclohexanone at 2 h reaction. After that time the amount of cyclohexyl hydroperoxide was observed to be lower than others and its amount was relatively constant during reaction. This 146 confirmed that cyclohexyl cyclohexanol/cyclohexanone. hydroperoxide was the intermediate for The amount of cyclohexanol and cyclohexanone gradually increased during reaction. From Figure 5.7 it is also observed that the amount of cyclohexanone at the first hour until 7 h of reaction is higher than cyclohexanol. The amount of cyclohexanone was equalized by cyclohexanol after 8 h of reaction time. At the end of the reaction (24 h), amount of cyclohexanol was higher than cyclohexanone (see Figure 5.5). The mechanism of the reaction of cyclohexane with TBHP on Ti-OMS-2 catalysts was proposed based on the distribution of the products. 2.0 Amount of Product / mmol cyclohexanone 1.5 cyclohexanol cyclohexyl hydroperoxide 1.0 0.5 0.0 0 2 4 6 8 Time / h Figure 5.7: Yield of products vs time on Ti-OMS-2 in the reaction of cyclohexane with TBHP as oxidant. Metal-catalyzed oxidation involving alkyl peroxides may proceed either through a homolytic or heterolytic mechanism. Transition metal salts of Co, Mn, Fe, Cu, or the metal oxides are normally involved in homolytic cleavage [86]. Sheldon et al. [48] noted that strong (one-electron) oxidants, e.g., later and/or first row transition elements such as Cr(VI), Mn(III), Co(III) and Fe(III), favour oxo-metal pathways and/or homolytic decomposition of TBHP. Due to the variable oxidation states of Mn (+2, +3, and +4) in OMS-2, it may be possible for the reaction to follow both homolytic and heterolytic pathway. However, the possibility of reaction 147 mechanism of oxidation of cyclohexane was studied. Since the ratio of cyclohexanone to cyclohexanol in all the catalysts was relatively the same and less than 1, implying the mechanism reaction on all catalysts was the same. The distribution of the product vs time was studied for Ti-OMS-2(0.67). The existence of cyclohexyl hydroperoxide indicated that autooxidation process occurred in the oxidation of cyclohexane on OMS-2 and modified OMS-2 samples using TBHP as oxidant [201]. It was also observed in oxidation of cyclohexane on OMS-1 [202] and oxidation of olefins on OMS-2 [28] with TBHP as oxidant. The first step of this reaction was homolytic reaction of TBHP with catalyst to form radical as shown in Figure 5.8. The formation of cyclohexyl hydroperoxide was from the reaction of cyclohexyl radical with oxygen is shown in Figure 5.9. The formation of cyclohexanone from cyclohexyl hydroperoxide is via heterolytic pathway while the formation of cyclohexanol is via homolytic pathway [203]. The heterolytic and homolytic pathway in decomposition of cyclohexyl hydroperoxide are shown in Figures 5.10 and 5.11, respectively. However, there was a possibility of reaction of 2 cyclohexyl hydroperoxide radical to form 1,4-dicyclohexyltetraoxidane and then decompose to form cyclohexanol, cyclohexanone and O2 as shown in Figure 5.12. Based on the product distribution at first 8 h reaction and after 24 h reaction on Ti-OMS-2(0.67), the mechanism of the reaction is proposed. Cyclohexanone was higher at beginning of the reaction. However, the higher amount of cyclohexanol than cyclohexanone after 24 h reaction indicated that the reaction occurred given via mechanism in Figures 5.9 and Figure 5.10. After 24 h reaction, the amount of cyclohexanol was higher than cyclohexanone; strongly suggest that the reaction of cyclohexyl radical with OH radical to form cyclohexanol occurred as shown in Figure 5.11. The mechanism in Figure 5.12 could not be proven by the distribution of the products. Since the homolytic and heterolytic pathways occur in the reaction of cyclohexane on OMS-2 and modified OMS-2, therefore mechanism in Figure 5.12 is possible. 148 OH O• O catalyst + •OH CH• Figure 5.8: Homolytic pathway to form radical from TBHP over catalyst. O• O2 O CH• OH O + CH• Figure 5.9: Formation of cyclohexyl hydroperoxide. O O H H catalyst O + H2O Figure 5.10: The heterolytic pathway of the formation of cyclohexanone from cyclohexyl hydroperoxide. 149 OH catalyst O •OH + O• OH + C• H •OH OH Figure 5.11: The homolytic pathway of the formation of cyclohexanol from cyclohexyl hydroperoxide. O• O O• O + H O O O2 + O + O O OH Figure 5.12: The heterolytic pathway of the formation of cyclohexanol and cyclohexanone from 1,4-dicyclohexyltetraoxidane. 150 5.4.3 Oxidation of Cyclohexene over Ti-OMS-2 Catalyst Oxidation of cyclohexene with TBHP as oxidant in acetonitrile as solvent was done over OMS-2, Ti-OMS-2(0.18; 0.67) and TiO2 rutile. The reaction performed in order to study the effect of titanium on OMS-2 to catalytic oxidation of cyclohexene. TiO2 rutile in the rutile phase was used since this phase was observed as impurity in Ti-OMS-2(0.67). Ti-OMS-2 was chosen for further study of oxidation reaction because of its performance in oxidation of cyclohexane. Beside that, TiOMS-2 (0.67) material could retain the cryptomelane structure at higher molar ratio of metal doping to manganese compared to other metal doped. The reaction was carried out at 70 oC for 2h and the products observed were epoxy cyclohexane, 2 cyclohexen-1-one and 2 cyclohexen-1-ol as shown in Figure 5.13. O O Catalyst, TBHP + Acetonitrile, 80 oC cyclohexene epoxy cyclohexane + 2 cyclohexen-1-one OH 2 cyclohexen-1-ol Figure 5.13: Reaction condition of cyclohexene and its products. The conversion of cyclohexene and selectivity towards 2 cyclohexen-1-one, 2 cyclohexen-1-ol and epoxycyclohexane as the reaction products are shown in Figure 5.14. As demonstrated, the reaction catalyzed by all the catalysts produced the highest yield of cyclohexanone and their selectivity towards the formation of products are almost similar to each other except for TiO2 rutile sample. There was no epoxy cyclohexane observed as product by using TiO2 rutile as the catalyst. From the figure it is also seen that the conversion of cyclohexene on Ti-OMS-2(0.18) is 151 significantly higher than that of OMS-2 catalyst. It was well evidenced that titanium incorporated OMS-2 enhanced the catalytic performance of OMS-2 catalyst which indicates that there is a synergetic effect of Ti and OMS-2 in Ti-OMS-2 catalyst. However, the most active was Ti-OMS-2(0.67) catalyst, where the effect nonframework titanium active sites on OMS-2 material in the enhancement of its catalytic activity in the oxidation of cyclohexene were significant. 70 60 Product selectivity (%) 2-cyclohexen-1-one 2 cyclohexen-1-ol Conversion / % 50 epoxycyclohexane 40 77 68 30 76 20 25 10 18 20 66 34 5 4 7 0 TiO2 OMS-2 Ti-OMS-2 (0.18) Ti-OMS-2 (0.67) Figure 5.14: The conversion and product selectivity of oxidation of cyclohexene with tert-butyl hydroperoxide (TBHP) using TiO2, OMS-2, Ti-OMS-2(0.18), and TiOMS-2(0.67). {All reactions were carried out at 70 ºC for 2 h with cyclohexene (5 mmol), 70% aqueous TBHP (10 mmol), acetonitrile (15 ml) and catalyst (50 mg). The conversion and the amount of product obtained in blank experimental have been subtracted}. The formation of the allylic oxidation products 2-cyclohexene-1-one and 2cyclohexene-1-ol which was more dominant than epoxy cyclohexane shows the preferential attack of the activated C-H bond over the C=C bond [204]. TBHP as oxidant promoted the allylic oxidation pathway and epoxidation was minimized. 152 This was also observed under alumina-supported with divalent and trivalent transition metal ions and complexes [204-206]. 5.4.2 Oxidation of Styrene over Different Location of Titanium sites on TiOMS-2 Catalysts Oxidation of styrene was carried out over different locations of titanium on OMS-2 material in order to relate it to the catalytic activity of that material. The catalysts used in this reaction were titanium silicate-1 (TS-1), manganese oxide octahedral molecular sieve (OMS-2), titanium incorporated OMS-2 material {TiOMS-2(0.18, 0.43, 0.67)}, titanium impregnated OMS-2 material {TiO2-OMS2(imp)} and physical mixture of TiO2 rutile and OMS-2 material {TiO2-OMS2(mix)}. TS-1 was supplied by Nur et al., [207] and used as the reference catalyst. The location of titanium was incorporated in the framework and/or located on the surface of Ti-OMS-2 and elucidated by spectroscopic techniques as discussed in Chapter 4. For each Ti-OMS-2 materials are, the following assumptions are made: Ti-OMS-2(0.18; 0.43): All titanium are incorporated well in the framework structure of OMS-2. Ti-OMS-2(0.67): Titanium exists in framework and non-framework sites. TiO2-OMS-2(imp): The only sample where non-framework titanium exists in OMS-2, there is chemical interaction between titanium and OMS-2. TiO2-OMS-2(mix): A mixture of rutile TiO2 and OMS-2 exist; there is no chemical interaction between both materials. The reaction condition of styrene oxidation was similar to oxidation of cyclohexene. The reaction was carried out at 70 oC in oil bath using acetonitrile and TBHP as solvent and oxidant, respectively. The products of this reaction are styrene oxide, phenyl acetaldehyde and benzaldehyde as shown in Figure 5.15. 153 O Catalyst, TBHP + o acetonitrile, 70 C styrene oxide styrene O + O phenyl acetaldehyde benzaldehyde Figure 5.15: Oxidation of styrene and its product on catalysts using TBHP as oxidant. Reaction products of oxidation of styrene using TBHP as the oxidant catalyzed by TiO2 (rutile phase), OMS-2, TiO2-OMS-2 and Ti-OMS-2 were analyzed by GC. In this study, TS-1 was also used as a reference catalyst. TS-1 (2% of titanium, mol %) was prepared according to a procedure described earlier [207]. The selectivity towards benzaldehyde, styrene oxide and phenylacetaldehyde as the reaction products is shown in Figure 5.16. The reaction catalyzed by all the catalysts produced the highest yield of benzaldehyde and their selectivity towards the formation of products are almost similar to each other. The high selectivity (52 %) towards benzaldehyde over TS-1 was surprising since phenylacetaldehyde and benzaldehyde were the major products from the oxidation of styrene catalyzed by the TS-1 zeolite [175], while the speculated product, styrene oxide, was not detected. Brönsted acid sites originating from framework titanium species catalyzed the rearrangement of the intermediate, leading to the formation of phenylacetaldehyde [208]. This argument is in agreement with our oxidation of styrene over TS-1, since no Brönsted acid sites was detected in our TS-1 [207]. The acidity study by pyridine adsorption shows that Lewis acid sites were formed in Ti-OMS-2 (0.67). In contrast, no Lewis acids were observed for OMS-2 sample. The absence of peaks at 1540 cm-1 confirmed that there were no Brönsted acid sites in both OMS-2 and Ti-OMS-2 (0.67) samples (see Section 4.4.2). This is the possible reason why TiO2-OMS-2, TiOMS-2, TiO2 and OMS-2 catalysts are not selective towards phenylacetaldehyde. A 0 10 20 30 40 50 60 Blank 5 12 83 TiO2 76 8 16 Benzaldehyde Styrene oxide TS-1 Phenylacetaldehyde Product selectivity (%): 52 15 33 OMS-2 74 8 18 76 8 Ti -OMS-2 71 9 20 16 Ti/Mn = 0.43 Ti/Mn = 0.18 72 10 18 Ti/Mn = 0.67 68 14 18 Ti/Mn = 0.18 73 OTS/ Ti-OMS-2 70 20 10 Ti/Mn = 0.67 18 9 TiO2 -OMS-2 TiO2 -OMS-2 OTS/ OMS-2 imp mix 69 16 15 Ti/Mn = 0.67 (15 ml) and catalyst (50 mg) with vigorous stirring. Ti-OMS-2, OMS-2 and TS-1. All reactions were carried out at 70 C with styrene (5 mmol), 70% aqueous TBHP (10 mmol), acetonitrile Figure 5.16: The conversion and product selectivity of oxidation styrene with tert-butyl hydroperoxide (TBHP) using TiO2, TiO2-OMS-2, Conversion / % 70 154 155 high selectivity towards benzaldehyde which may be due to OMS-2 and TiO2, promoted the carbon–carbon bond cleavage, thus resulting in the formation of benzaldehyde. As shown in Figure 5.16, a considerable increase in conversion of styrene over Ti-OMS-2, OMS-2, TiO2-OMS-2, TiO2 and TS-1 after 3 h of the reactions was clearly observed when Ti-OMS-2 (0.67) and TiO2–OMS-2 (imp) were used as catalysts. The increase in oxidation activity of Ti-OMS-2 (0.67) and TiO2–OMS-2 (imp) can be explained on the basis of the presence of non-framework titanium species. The superior performance of Ti-OMS-2 (imp) and Ti-OMS-2 (0.67) strongly suggests the occurrence of synergetic effect of non-framework Ti with OMS-2, since Ti-OMS-2 (mix), a mechanical mixture of TiO2 and OMS-2, gave relatively lower conversion of styrene (see Figure 5.16). Leaching was a particular problem of solid catalysts in liquid phase reaction. The catalysts were recycled three times. The activities of the recovered and dried Ti-OMS-2, OMS-2 and TiO2-OMS2 showed insignificant change (ca. 3-5%) which correspond to experimental observations within experimental error in their catalytic activity in the recycling test. This suggests a good regenerability of the catalysts in the oxidation of styrene with tert-butylhydroperoxide. As shown in Figure 5.16 the catalytic activity of OMS-2 increased after attachment of OTS. It may be caused by access of styrene and TBHP to reach the surface of that catalyst after modification due to the increasing hydrophobic and hydrophilic properties of modified OMS-2. However, appropriate amount of OTS attached to the sample was necessary to enhance the catalytic properties of sample. In case of OTS/Ti-OMS-2 the decreased activity which may be caused by blocking the Ti active site after alkylsylilation modification since the amount of OTS attached to its surface was higher than OMS-2 surface. 156 5.3 The Effect of Lewis Acidity in Catalytic Oxidations The analysis of acidity of Ti-OMS-2(0.67) proved that the titanium site on OMS-2 caused increasing Lewis acidity of the catalyst. Table 5.3 shows the correlation between Lewis acidity and conversion of substrates in oxidation of cyclohexane, cyclohexene and styrene with TBHP as oxidant. Table 5.3: The correlation of Lewis acidity of samples to conversion of cyclohexane, cyclohexene and styrene. Catalysts Lewis acidity strengtha Conversion of cyclohexane (%) Conversion of cyclohexene (%) Conversion of styrene (%) Ti-OMS-2 (0.67) strong 7.4 65 70 OMS-2 weak 4.4 21 52 a The strength of Lewis acidity is relative to each other. It shows that Ti-OMS-2(0.67) which have strong Lewis acidity than OMS-2 gave higher conversion in oxidation of cyclohexane, cyclohexene and styrene. It indicated that Lewis acid played a role in the enhancement of catalytic activity of TiOMS-2 material. This was in agreement with Corma and García [108]; that Lewis acids can catalyze oxidation reactions. According to them, there are two possible ways for the Lewis acids to catalyze oxidation reactions which depend on the substrates i.e abstraction of electron and acid-base adduct. The possible role of the Lewis acid sites in enhancing the activity of Ti-OMS-2 in oxidation reactions are summarized in Table 5.4. It shows that the main products for oxidation of cyclohexane are C-H bond activation in which Lewis acidity enhances the catalytic activity via abstraction of electron to form radical in homolytic pathway. In cyclohexene there are two ways in which Lewis acid sites play a role in oxidation reaction that is, electron abstraction and acid-base adduct. However, the selectivity of cyclohexene as shown in Section 5.4.3 indicates that the catalyst is more selective toward 2 cyclohexen-1-one and 2 cyclohexe-1-ol compared to epoxycyclohexane and C-H bond activation is preferred than C=C bond 157 activation. In the oxidation of styrene, Lewis acid sites can only promote acid-base adduct since only C=C bond can be activated and it has no allylic C-H bond. Table 5.4: The possible role of Lewis acids of catalyst in oxidation of cyclohexane, cyclohexene and styrene. Oxidation reactions Products cyclohexane Cyclohexanone and C-H cyclohexanol Electron abstraction to form radical Cyclohexene 2 cyclohexen-1one, 2 cyclohexe1-ol and epoxycyclohexane C-H and C=C Electron abstraction to form radical and acid-base adduct Styrene Benzaldehyde, phenyl acetaldehyde and styrene oxide C=C acid-base adduct 5.4 Site activation Role of Lewis acid sites The Role of Different Location of Ti Sites in Ti-OMS-2 in Oxidation Reactions Table 5.5 summarizes the role of location of Ti sites to catalytic activity of Ti-OMS-2 in oxidation of cyclohexane, cyclohexene and styrene with TBHP as oxidant. Although OMS-2 itself is active for all reactions, incorporation of titanium in the framework and non-framework of OMS-2 catalyst evidently affect the catalytic activity of OMS-2 materials. For oxidation of cyclohexane and cyclohexene, both Ti sites in framework and non-framework enhanced the catalytic activity of OMS-2 catalysts. However, for oxidation of styrene, only titanium nonframework played a role in the enhancement of catalytic activity of Ti-OMS-2. As mentioned before, the main products of oxidation of cyclohexane and cyclohexene was C-H bond activation whereas in oxidation of styrene was C=C activation. This suggested that incorporation of Ti in the framework only enhances the catalytic 158 activity of Ti-OMS-2 for C-H bond activation and not C=C bond activation. In contrast, Ti sites in non-framework were active for both bond activations. Table 5.5: The role of Ti sites location in oxidation reaction. Location of Ti in OMS-2a Oxidation reactions OMS-2 Ti in framework Ti in Ti in nonframework and framework non-framework cyclohexane + ++ +++ not available cyclohexene + ++ +++ not available styrene + not active ++ ++ a 5.5 The location of Ti in OMS-2 has been elucidated by several spectroscopy techniques (see Section 4.4) The relative catalytic activity of: + < ++ < +++ Catalytic Study on Consecutive Reaction of 1-octene to 1,2-octanediol A model reaction for consecutive oxidation and acid reaction was the synthesis of 1,2 octanediol from 1-octene. Actually, there are two steps in this reaction; first is oxidation of 1-octene to 1,2 epoxyoctane using TBHP as oxidant, second is ring opening of epoxide by hydrolysis due the existence of Brönsted acid in the reaction mixture to form 1,2 octane diol (see Figure 5.17). Sulphated Ti-OMS2(0.67) samples were synthesized by using different solvent as catalysts. The effect of the different solvents used in preparation of sample and the amount of sulfate loading was evaluated in this reaction. Figure 5.18 shows the yield of epoxyoctane and 1,2-octanediol on Ti-OMS2 (0.67) and sulphated Ti-OMS-2(0.67). Sulphated samples were prepared in water and toluene in order to study the effect of solvent in preparation of sulfate sample. In order to study the effect of sulfate content the different amount of concentrated H2SO4 i.e.150 μL and 200 μL was used. The sample labelled as SW150-Ti-OMS-2, indicates that sulphated sample was prepared in water with an amount of 159 concentrated H2SO4 (150 μL). Sulphated sample prepared in toluene was labelled using abbreviation ST. There was no diol detected on Ti-OMS-2(0.67) catalyst. Diol was observed in all sulphated catalysts. It is known that sulphated metal oxide increases both Brönsted and Lewis acid sites of the samples. In contrary, in the case Figure 5.17: Consecutive oxidation and acid reaction to form of 1,2 octane diol from 1-octene on sulphated Ti-OMS-2 catalyst. 1,2-epoxyoctane 1,2-octanediol Yield of products / μmol 450 400 350 300 250 200 150 100 50 ~0 0 Figure 5.18: Yield of epoxyoctane and 1,2 octane diol after 24 h reaction. 160 of sulphated Ti-OMS-2 sample, there was no Brönsted acid site detected by H1-NMR analysis. Since diol was formed in the reaction, it suggested that Brönsted acid may be in the reaction mixture playing the role in hydrolysis of epoxide to diols. It was also possible that the existence of Brönsted acid in the reaction mixture was due to Brönsted acid formed from Lewis acid sites through hydrolysis in the presence of water [193]. Water existed in the reaction mixture since the reaction used aqueous TBHP 70% in water as oxidant. The generation of Brönsted acid by Lewis acid (LA) sites is described in the following equation: LA + H2O H2O+ ---LA H+ + LAOH-, where = vacant sites of Lewis acid. However, although Ti-OMS-2 itself has Lewis acid site but no diol observed in that sample could be due to the Lewis acidity of Ti-OMS-2 was low. The Lewis acidity increases after attachment of sulphate due to the strain in bonding of Ti-O in the presence of sulphate as described in 4.31. From Figure 5.18, it is observed that both sulphated Ti-OMS-2 catalysts in toluene and water with the same amount of sulfate loaded gave relatively the same amount of both epoxide and diols. This indicates that either water or toluene as solvent in synthesis of sulphated Ti-OMS-2 gave the same effect to the catalytic activity of the catalysts. The effect of sulphate content was also shown in Figure 5.13. The amount of 1,2-octanediol produced by sulphated sample in toluene (ST150-Ti-OMS-2 compared to ST200-Ti-OMS-2) and in water (SW150-Ti-OMS-2 compared to SW200-Ti-OMS-2) was higher at higher sulphate loading. This could be due to the increasing amount of Brönsted acids in the reaction mixture; indicated by the increasing Lewis acid sites on the samples. CHAPTER 6 SUMMARY AND CONCLUSION 6.1 Summary Manganese oxide octahedral molecular sieve (OMS-2) and modified OMS2 materials have been successfully synthesized. OMS-2 synthesized in buffer solution is higher in crystallinity and hydrophobicity compared to OMS-2 prepared without buffer solution. TG analysis shows that oxygen lattice of OMS-2 prepared without buffer is easier to release than OMS-2 prepared with buffer solution. Metal substituted OMS-2 prepared by oxidation of the mixture of metal and manganese ion in acidic condition was successfully incorporated to OMS-2. The capability of metal ion incorporated decreased in the order of Ti>Fe>Co>Cu. Incorporation of metals did not change the morphology of OMS-2 except for FeOMS-2. It is also observed that potassium exchanged by H+ retained the cryptomelane structure of OMS-2 and has similar morphology to their parents. A new method for synthesis of metal substituted OMS-2 material without addition of any manganese ion solution is reported. The method gave higher M/Mn ratio compared to the previous method. Only titanium substituted sample and only Ti2(SO4)3 as titanium source retained the cryptomelane structure of OMS-2 material. Incorporation of titanium changes the fibrous morphology of OMS-2 to spherical with the particle size of about 50-80 nm. Detailed study of Ti-OMS-2 showed that 162 the sample with molar ratio of Ti:Mn of about 0.43 has the pure cryptomelane structure. This is the highest level of framework substituted OMS-2 material ever reported. However, with the higher Ti:Mn molar ratio than ca. 0.5, produced rutile phases of TiO2 growth with cryptomelane phase. Lewis acid sites were observed in Ti-OMS-2 sample, may be due to the defect of structure after incorporation of titanium. The location of titanium in the framework was confirmed by XPS. The alkylsilylation on OMS-2 and Ti-OMS-2 materials did not successfully make the surface of those materials more hydrophobic. Lastly, sulfation on Ti-OMS-2 was also not successful in creating the Brönsted acid site in Ti-OMS-2 material. The physicochemical properties-catalytic activity of modified OMS-2 catalysts in model oxidation reactions and consecutive oxidation-acid reaction are summarized in Table 6.1. The method of synthesis affects the physicochemical properties OMS-2 catalyst which also changes the catalytic activity of the catalyst. OMS-2 prepared in buffer is less active than OMS-2 prepared without buffer. Modification of OMS-2 material by metals substitution and ion exchange by H+ increase the catalytic activity of OMS-2 material in oxidation of cyclohexane. Both modifications gave similar effect that is to increase the Lewis acidic sites in OMS-2 which plays a role in abstraction of electron in homolytic pathway mechanism of oxidation of cyclohexane. The activity of metal substituted sample (Ti-OMS-2) is maximum at the highest Ti/Mn ratio. Ti sites in such catalyst evidently exist in both framework and non-framework location. The role of framework and non-framework Ti sites was evaluated in catalytic oxidation of cyclohexane, cyclohexene and styrene. Titanium framework plays a role in oxidation reaction by C-H bond activation whereas nonframework titanium is active for both C-H and C=C bond activations. Sulphated Ti-OMS-2 catalyst is active in the synthesis of diols from alkenes. However, Brönsted acid sites were not detected. In the system, Brönsted acids which play an important role in hydrolysis of epoxides to diols could have been generated from the interaction of Lewis acid sites on the samples with water. o H-exchanged of OMS-2 and MOMS-2 C-H bond oxidation: Homolytic and Oxidation of o OMS-2 cyclohexane with o Metal substituted heterolytic TBHP as oxidant M-OMS-2 (M=Fe, Ti, Co and Cu) o OMS-2 prepared without buffer o OMS-2 prepared Abstraction of secondary H+ : Mars van Krevelen Mechanism. with buffer Oxidation of benzyl alcohol with air as oxidant The possible reaction mechanism from the main product Catalyst Reaction Continued in Page 166 ¾ The Lewis active site plays an important role in enhancing catalytic oxidation of the catalyst in the case of cyclohexane via abstraction of electron to form radical chains. ¾ There is a correlation of ionic radii of metal substituted with conversion of cyclohexane. Increasing ionic radii increases conversion of cyclohexane which could be due to weaker M-O bond resulting more positive metal centre and increases Lewis active site. ¾ Catalytic activity of M-OMS-2 increases in the following order Ti-<Fe-<Cu-<Co-OMS-2, however Ti-OMS-2 at highest titanium content is the most active catalyst. ¾ Both metal substituted and H-exchanged enhance the catalytic activity of OMS-2 ¾ OMS-2 prepared without buffer is more active than prepared with buffer solution which is due to effective oxygen release from the catalyst via Mars van Krevelen mechanism. Physicochemical properties-catalytic activity relationship Table 6.1: The physicochemical properties-catalytic activity relationship of the catalysts. 163 o TiO2 rutile o Ti-OMS-2 (0.18; 0.67) o OMS-2 o OTS/Ti-OMS-2 o OTS/OMS-2 o TiO2-OMS-2(imp) o TiO2-OMS-2(mix) o Ti-OMS-2 o TS-1 Oxidation of styrene with TBHP as oxidant o OMS-2 Oxidation of cyclohexene with TBHP as oxidant C=C bond oxidation: heterolytic metal pathway C-H and C=C bond oxidation: Homolytic and heterolytic. Continued in Page 167 ¾ Alkylsylilation enhance the activity of OMS-2 but decrese the ativity of Ti-OMS-2 which may be caused by blocking of Ti active site. ¾ The catalytic activity of OMS-2 is higher than TS-1. ¾ The framework of titanium species in Ti-OMS-2 has no effect in enhancement of catalytic activity. The only Ti in non-framework structure of OMS-2 induces a synergetic effect that enhances the catalytic activity. ¾ Ti-OMS-2(0.67) has a higher conversion to styrene compared to OMS-2. ¾ Ti-OMS-2 which has a higher Lewis acid site than OMS-2 gave similar effect with oxidation of cyclohexane since the main product is from C-H bond activation. ¾ Both titanium on the framework and non-framework enhance the conversion of cyclohexene. ¾ The main products of 2 cyclohexen-1-one and 2 cyclohexen-1-ol indicates that C-H bond oxidation is dominant. 164 Consecutive oxidation and acid reaction of 1-octene to 1,2 octane diol with TBHP as oxidant. Sulphated Ti-OMS-2(0.67) in different solvent and different sulphate content. ¾ Solvent has no effect in the synthesis of sulphated samples with the amount of diol produced. ¾ The higher amount of sulphate content forms more diol. ¾ No diol is observed in non sulphated sample which indicate that there is a need for more strength in Lewis acid site in order to interact with water to form Brönsted acid. ¾ Diol was observed using sulphated TiOMS-2 as catalyst due to the generation of Brönsted acid through hydrolysis by interaction of Lewis acid site and water. 165 166 6.2 Conclusion The results show that different methods of preparation resulted in different properties of OMS-2. The physicochemical properties-catalytic activity relationship of modified OMS-2 in oxidation and in consecutive oxidation and acid reactions were studied and presented in Figure 6.1. The results from the study provedöö that metals substitution and H-exchanged enhance the catalytic activity of OMS-2 in oxidation of cyclohexane with TBHP as oxidant. Sulphated Ti-OMS-2 successfully catalyzed the consecutive oxidation and acid reactions for synthesis of diols from alkenes; despite the presence of Brönsted acid sites. The reaction was successful due to the generation of Brönsted acid during the reaction that resulted from interaction of Lewis acid sites and water. Synthesis of OMS-2 with and without buffer solution in acidic condition gave different physicochemical properties and catalytic activity in oxidation of benzyl alcohol using air as the oxidant. It showed that the catalytic activity of OMS-2 prepared without buffer is better than with buffer solution. The catalytic activity is closely related to the variety of physicochemical properties of the catalysts. The study showed that the properties include availability of oxygen lattice to be released, crystallinity and hydrophilicity-hydrophobicity of the samples. The synthesis of metal substitution is either with or without manganese ion, added to metal solution in acidic condition before being oxidized by potassium permanganate. Only Ti substituted OMS-2 was successfully synthesized by the later method (without addition of manganese ion). Characterization of the samples shows that high amount titanium Ti:Mn ratio of 0.43 retained the cryptomelane structure of OMS-2, while other metals caused collapse of the cryptomelane structure. It is the highest amount of metal incorporated OMS-2 ever been reported; achievable by the first method which is without addition of manganese (II) ion solution. Among the metal substituted OMS-2 catalysts by addition of manganese ion, Ti substituted has the less enhancement effect to the catalytic activity. The enhancement of the catalytic activity of metal substituted OMS-2 in oxidation of Increased followed: Ti4+- <Fe2+- <Co2+- <Cu2+- Framework Tunnel H-exchanged Weaken Mn-O bond caused Mn more positively charged 1,2 octanediol Brönsted acid generated during reaction Reaction of 1-octene with TBHP as oxidant No Brönsted acid site Sulfated Ti-OMS-2 Figure 6.1: Assignments of modified OMS-2 in oxidation of cyclohexane and consecutive reaction of 1-octene to 1,2 octanediol Theactivityinoxidationofcyclohexane withTBHPasoxidantincreases Play role in electron abstraction Strengthen Lewis acidity Increasing of ionic radii decreases strength of M-O caused M more positively charged Ti-OMS-2 at high Ti/Mn ratio (most active) Framework + non-framework Metal incorporated Manganese oxide Octahedral Molecular Sieve (OMS-2) 167 168 cyclohexane with TBHP as oxidant without using any solvent in increasing order is: Ti- < Fe- < Cu- < Co-OMS-2. The results also show that all H-exchanged samples of either OMS-2 or M-OMS-2 enhance the conversion of cyclohexane compared to the unchanged one. The enhancement in catalytic performance of metals substituted and H-exchanged in oxidation of cyclohexane with TBHP as oxidant is presented in Figure 6.1. There is a correlation of ionic radii of M-OMS-2 to conversion of cyclohexane, which is related to the strength Lewis acid sites. Similar effect was also observed in H-exchanged catalyst. The much smaller ionic radii of H+ compared to K+ causes the strain in the cryptomelane lattice, which in turn strains the Mn–O bonds and makes Mn more positively charged, consequently increases Lewis acidity. Lewis acid sites may play a role in the oxidation reaction via abstraction of electron for activation of C-H bond. The catalyst with high titanium content [Ti-OMS-2(0.67)] with Ti sites located in framework and non-framework position is found to be the most active catalyst for the reaction of cyclohexane. The activity of the catalyst was also + increased after ion exchanged with H . This suggests that the superior catalytic activity of the catalyst is also due to of non-framework Ti site, besides the increasing Lewis acidity by H-exchange and incorporation of Ti in framework. The catalytic activity of Ti substituted OMS-2 in the oxidation of cyclohexene and styrene with TBHP as oxidant, demonstrates that framework and non-framework Ti sites play an important role (see Figure 6.2). However, in oxidation of styrene, only non-framework titanium contributes to the increasing of conversion of styrene. From the main products of the oxidation of cyclohexane, cyclohexene and styrene, it can be concluded that the framework Ti sites enable the abstraction of electron and have no effect to acid-base adduct pathway, whereas nonframework Ti sites play a role for both pathways as shown in Figure 6.2. It is also observed that the activity of OMS-2 and modified OMS-2 catalysts are higher than TS-1 which makes those modified OMS-2 catalysts as promising catalyst for oxidation reactions. Main products: C-H bond activation Cyclohexene Ti non-framework only enhance the catalytic activity of OMS-2 Acid-base adduct Main products: C=C bond activation Styrene Figure 6.2 : The role of the location of Ti sites in Ti-OMS-2 in oxidation of cyclohexane, cyclohexene and styrene. Ti framework and non-framework enhance the catalytic activity of OMS-2 Electron abstraction Main products: C-H bond activation Cyclohexane On the location of different Ti sites in TiOMS-2 in oxidation reactions 169 170 Lastly, it is concluded that every reaction requires the specific active sites, hydrophobicity-hydrophylicity and also depends on the crystallinity of the catalyst. The selection of suitable synthetic methods, type, amount and location of metal substituted which gave significant effect to the physicochemical properties of OMS-2 materials is needed to find the best catalyst for a selected reaction. As a result of the research, a guideline in the selection of the active site for suitable reaction is proposed. 171 REFERENCES 1. Chorkendorff, I. and Niemantsverdriet, J. W. Concepts of Modern Catalysis and Kinetics. Wiley-VCH, Weinheim. 2003. 2. van Santen, R. A. and Niemantsverdriet, J. W. Chemical Kinetics and Catalysis. Plenum, New York. 1995. 3. Basic Energy Sciences Advisory Committee. Opportunities for Catalysis in the 21st Century. Subpanel Workshop Report. May 14-16 2002. Gaithersburg, United States. 4. Chen, B., Dingerdissen, U., Krauter, J. G. E., Lansink Rotgerink, H. G. J., Moíbus, K., Ostgard, D. J., Panster, P., Riermeier, T. H., Seebald, S., Tacke, T. and Trauthwein, H. New Developments in Hydrogenation Catalysis Particularly in Synthesis of Fine and Intermediate Chemicals. Applied Catalysis A: General. 2005. 280(1): 17-46. 5. Clark, J. H. and Rhodes, C. N. Clean Synthesis Using Porous Inorganic Solid Catalysts and Supported Reagents. RSC, Cambridge. 2000. 6. The Freedonia Group. World Catalyst: Industry study with forecasts to 2010 & 2015. January 2007: 4. 7. Coulier. Hydrotreating Model Catalysts: From Characterization to Kinetics. Ph.D Thesis. Eindhoven University of Technology, the Netherlands. 2001. 8. Clark, J. H. Catalysis for Green Chemistry. Pure and Applied Chemistry. 2001. 73(1): 103-111. 9. Linssen, T., Cassiers, K., Cool, P. and Vansant, E. F. Mesoporous Templated Silicates: An Overview of Their Synthesis, Catalytic Activation and Evaluation of the Stability. Advances in Colloid and Interface Science. 2003. 103(2): 121-147. 10. Selvam, P., Bhatia, S. K. and Sonwane, C. G. Recent Advances in Processing and Characterization of Periodic Mesoporous MCM-41 Silicate Molecular Sieves. Industrial and Engineering Chemistry Research. 2001. 40(15): 32373261. 172 11. Suib, S. L. Synthesis, Characterization and Catalysis with Microporous Ferrierites, Octahedral Molecular Sieves, and Layered Materials. Studies in Surface Science and Catalysis, Vol. 102, 1996. pp. 47-74. 12. Suib, S. L. Microporous Manganese Oxides. Current Opinion in Solid State and Materials Science. 1998. 3(1): 63-70. 13. Feng, Q., Kanoh, H. and Ooi, K. Manganese Oxide Porous Crystals. Journal of Materials Chemistry. 1999. 9(2):319-333. 14. Post, J. E. Manganese Oxide Minerals: Crystal Structures and Economic and Environmental Significance. Proceedings of the National Academy of Sciences of the United States of America. 1999. 96(7):3447-3454. 15. O'Young, C.-L., Sawicki, R. A., Shen, Y.-F. and Suib, S. L. Manganese Oxide Octahedral Molecular Sieve as Acid-Base Catalyst. U.S. Patent: 5, 523, 509. 1996 16. Shen, Y. F., Zerger, R. P., DeGuzman, R. N., Suib, S. L., McCurdy, L., Potter, D. I. and O'Young, C. L. Manganese Oxide Octahedral Molecular Sieves: Preparation, Characterization, and Applications. Science. 1993. 260(5107):511-515. 17. Tian, Z. R., Tong, W., Wang, J. Y., Duan, N. G., Krishnan, V. V. and Suib, S. L. Manganese Oxide Mesoporous Structures: Mixed-Valent Semiconducting Catalysts. Science. 1997. 276(5314):926-930. 18. Giraldo, O., Brock, S. L., Marquez, M., Suib, S. L., Hillhouse, H. and Tsapatsis, M. Materials: Spontaneous Formation of Inorganic Helices. Nature. 2000. 405(6782): 38. 19. Ching, S., Roark, J. L., Duan, N. and Suib, S. L. Sol-Gel Route to the Tunneled Manganese Oxide Cryptomelane. Chemistry of Materials. 1997. 9(3): 750-754. 20. Suib, S. L. Sorption, Catalysis and Separation by Design. Chemical Innovation. 2000. 30(3): 27-33. 21. Suib, S. L. Structure, Porosity, and Redox in Porous Manganese Oxide Octahedral Layer and Molecular Sieve Materials. Journal of Materials Chemistry. 2008. 18(14):1623-1631. 22. Suib, S. L. Porous Manganese Oxide Octahedral Molecular Sieves and Octahedral Layered Materials. Accounts of Chemical Research. 2008. 41(4):479-487. 23 Makwana, V. D., Garces, L. J., Liu, J., Cai, J., Son, Y.-C. and Suib, S. L. Selective Oxidation of Alcohols Using Octahedral Molecular Sieves: Influence of Synthesis Method and Property-Activity Relations. Catalysis Today. 2003. 85(2-4): 225-233. 173 24. Makwana, V. D., Son, Y.-C., Howell, A. R. and Suib, S. L. The Role of Lattice Oxygen in Selective Benzyl Alcohol Oxidation Using OMS-2 Catalyst: A Kinetic and Isotope-Labeling Study. Journal of Catalysis. 2002. 210(1): 46-52. 25. Rebello, J. S., Samant, P. V., Figueiredo, J. L. and Fernandes, J. B. Enhanced Electrocatalytic Activity of Carbon-Supported Mnox/Ru Catalysts for Methanol Oxidation in Fuel Cells. Journal of Power Sources. 2006. 153(1): 36-40. 26. Luo, J., Zhang, Q., Huang, A. and Suib, S. L. Total Oxidation of Volatile Organic Compounds with Hydrophobic Cryptomelane-Type Octahedral Molecular Sieves. Microporous and Mesoporous Materials. 2000. 35-36: 209-217. 27. Ghosh, R., Shen, X., Villegas, J. C., Ding, Y., Malinger, K. and Suib, S. L. Role of Manganese Oxide Octahedral Molecular Sieves in Styrene Epoxidation. Journal of Physical Chemistry B. 2006. 110(14): 7592-7599. 28. Ghosh, R., Son, Y.-C., Makwana, V. D. and Suib, S. L. Liquid-Phase Epoxidation of Olefins by Manganese Oxide Octahedral Molecular Sieves. Journal of Catalysis. 2004. 224(2): 288-296. 29. Khenkin, A. M., Weiner, L., Wang, Y. and Neumann, R. Electron and Oxygen Transfer in Polyoxometalate, H5PV2Mo10O40, Catalyzed Oxidation of Aromatic and Alkyl Aromatic Compounds: Evidence for Aerobic Mars-Van Krevelen-Type Reactions in the Liquid Homogeneous Phase. Journal of the American Chemical Society. 2001. 123(35):8531-8542. 30. Chen, X., Shen, Y. F., Suib, S. L. and O'Young, C. L. Characterization of Manganese Oxide Octahedral Molecular Sieve (M-OMS-2) Materials with Different Metal Cation Dopants. Chemistry of Materials. 2002. 14(2): 940948. 31. Xia, G. G., Yin, Y. G., Willis, W. S., Wang, J. Y. and Suib, S. L. Efficient Stable Catalysts for Low Temperature Carbon Monoxide Oxidation. Journal of Catalysis. 1999. 185(1): 91-105. 32. Liu, J., Makwana, V., Cai, J., Suib, S. L. and Aindow, M. Effects of Alkali Metal and Ammonium Cation Templates on Nanofibrous Cryptomelane-Type Manganese Oxide Octahedral Molecular Sieves (OMS-2). Journal of Physical Chemistry B. 2003. 107(35): 9185-9194. 33. Ching, S., Welch, E. J., Hughes, S. M., Bahadoor, A. B. F. and Suib, S. L. Nonaqueous Sol-Gel Syntheses of Microporous Manganese Oxides. Chemistry of Materials. 2002. 14(3):1292-1299. 34. Jothiramalingam, R., Viswanathan, B. and Varadarajan, T. K. Preparation, Characterization and Catalytic Properties of Cerium Incorporated Porous 174 Manganese Oxide OMS-2 Catalysts. Catalysis Communications. 2005. 6(1): 41-45. 35. Brock, S. L., Sanabria, M., Nair, J., Suib, S. L. and Ressler, T. Tetraalkylammonium Manganese Oxide Gels: Preparation, Structure, and Ion-Exchange Properties. J. Phys. Chem. B. 2001. 105(23): 5404-5410. 36. Cai, J., Liu, J., Willis, W. S. and Suib, S. L. Framework Doping of Iron in Tunnel Structure Cryptomelane. Chemistry of Materials. 2001. 13(7): 24132422. 37. Ching, S., Driscoll, P. F., Kieltyka, K. S., Marvel, M. R. and Suib, S. L. Synthesis of a New Hollandite-Type Manganese Oxide with Framework and Interstitial Cr(III). Chemical Communications. 2001(23): 2486-2487. 38. Jothiramalingam, R., Viswanathan, B. and Varadarajan, T. K. Synthesis and Structural Characterization of Copper Incorporated Manganese Oxide OMS-2 Materials Synthesized via Potassium Birnessite. Materials Chemistry and Physics. 2006. 100(2-3): 257-261. 39. Jothiramalingam, R., Viswanathan, B. and Varadarajan, T. K. Synthesis, Characterization and Catalytic Oxidation Activity of Zirconium Doped KOMS-2 Type Manganese Oxide Materials. Journal of Molecular Catalysis A: Chemical. 2006. 252(1-2): 49-55. 40. Krishnan, V. V. and Suib, S. L. Oxidative Dehydrogenation of 1-Butene over Manganese Oxide Octahedral Molecular Sieves. Journal of Catalysis. 1999. 184(2): 305-315. 41. Liu, J., Son, Y. C., Cai, J., Shen, X., Suib, S. L. and Aindow, M. Size Control, Metal Substitution, and Catalytic Application of Cryptomelane Nanomaterials Prepared Using Cross-Linking Reagents. Chemistry of Materials. 2004. 16(2): 276-285. 42. Polverejan, M., Villegas, J. C. and Suib, S. L. Higher Valency Ion Substitution into the Manganese Oxide Framework. Journal of the American Chemical Society. 2004. 126(25): 7774-7775. 43. Ramstad, A. L. and Mikkelsen, Ø. Structural Characterisation of CopperContaining Manganese Oxide Octahedral Molecular Sieve (Cu-OMS-2) Materials by X-Ray Absorption Spectroscopy and Cyclic Voltammetry. Journal of Molecular Structure. 2004. 697(1-3):109-117. 44. Segal, S. R., Suib, S. L. and Foland, L. Decomposition of Pinacyanol Chloride Dye Using Several Manganese Oxide Catalysts. Chemistry of Materials. 1997. 9(11): 2526-2532. 45. Zhou, H., Shen, Y. F., Wang, J. Y., Chen, X., O'Young, C.-L. and Suib, S. L. Studies of Decomposition of H2O2 over Manganese Oxide Octahedral Molecular Sieve Materials. Journal of Catalysis. 1998. 176(2): 321-328. 175 46. Zhou, H., Wang, J. Y., Chen, X., O'Young, C. L. and Suib, S. L. Studies of Oxidative Dehydrogenation of Ethanol over Manganese Oxide Octahedral Molecular Sieve Catalysts. Microporous and Mesoporous Materials. 1998. 21(4-6): 315-324. 47. Arends, I. W. C. E. and Sheldon, R. A. Activities and Stabilities of Heterogeneous Catalysts in Selective Liquid Phase Oxidations: Recent Developments. Applied Catalysis A: General. 2001. 212(1-2): 175-187. 48. Sheldon, R. A., Arends, I. W. C. E. and Lempers, H. E. B. Liquid Phase Oxidation at Metal Ions and Complexes in Constrained Environments. Catalysis Today. 1998. 41(4): 387-407. 49. He, J., -ping Xu, W., Evans, D. G., Duan, X. and Li, C.-y. Role of Pore Size and Surface Properties of Ti-MCM-41 Catalysts in the Hydroxylation of Aromatics in the Liquid Phase. Microporous and Mesoporous Materials. 2001. 44-45: 581-586. 50. Taramasso, M., Perego, G. and Notari, B. Preparation of Porous Crystalline Synthetic Material Comprised of Silicon and Titanium Oxides. U.S. Patent: 4,410,501. 1983 51. Taylor, R. J. K., Reid, M., Foot, J. and Raw, S. A. Tandem Oxidation Processes Using Manganese Dioxide: Discovery, Applications, and Current Studies. Accounts of Chemical Research. 2005. 38(11):851-869. 52. Prasetyoko, D., Ramli, Z., Endud, S. and Nur, H. TS-1 Loaded with Sulfated Zirconia as Bifunctional Oxidative and Acidic Catalyst for Transformation of 1-Octene to 1,2-Octanediol. Journal of Molecular Catalysis A: Chemical. 2005. 241(1-2): 118-125. 53. Prasetyoko, D., Ramli, Z., Endud, S. and Nur, H. Preparation and Characterization of Bifunctional Oxidative and Acidic Catalysts Nb2O5/Ts-1 for Synthesis of Diols. Materials Chemistry and Physics. 2005. 93(2-3): 443449. 54. Prasetyoko, D., Ramli, Z., Endud, S. and Nur, H. Enhancement of Catalytic Activity of Titanosilicalite-1 - Sulfated Zirconia Combination Towards Epoxidation of 1-Octene with Aqueous Hydrogen Peroxide. Reaction Kinetics and Catalysis Letters. 2005. 86(1): 83-89. 55. Arata, K., Matsuhashi, H., Hino, M. and Nakamura, H. Synthesis of Solid Superacids and Their Activities for Reactions of Alkanes. Catalysis Today. 2003. 81(1):17-30. 56. Chen, X. Decomposition of 2-Propanol over M-OMS-2 Materials and Dissociation of Stable Molecules in Plasma Reactors. Ph.D thesis. University of Connecticut, Storrs. 2000. 176 57. O'Young, C. L., Sawicki, R. A. and Suib, S. L. Micropore Size Distribution of Octahedral Molecular Sieves (OMS). Microporous Materials. 1997. 11(12): 1-8. 58. Tanaka, Y. and Tsuji, M. New Synthetic Method of Producing -Manganese Oxide for Potassium Selective Adsorbent. Materials Research Bulletin. 1994. 29(11): 1183-1191. 59. Rebello, J. S. and Fernandes, J. B. Ion Induced Synthesis of Manganese Oxide Octahedral Molecular Sieves (Oms-2). Microporous and Mesoporous Materials. 2005. 77(2-3):189-192. 60 De Guzman, R. N., Shen, Y. F., Shaw, B. R., Suib, S. L. and O'Young, C. L. Role of Cyclic Voltammetry in Characterizing Solids: Natural and Synthetic Manganese Oxide Octahedral Molecular Sieves. Chemistry of Materials. 1993. 5(10):1395-1400. 61 Villegas, J. C., Garces, L. J., Gomez, S., Durand, J. P. and Suib, S. L. Particle Size Control of Cryptomelane Nanomaterials by Use of H2O2 in Acidic Conditions. Chemistry of Materials. 2005. 17(7):1910-1918. 62. DeGuzman, R. N., Shen, Y. F., Neth, E. J., Suib, S. L., O'Young, C. L., Levine, S. and Newsam, J. M. Synthesis and Characterization of Octahedral Molecular Sieves (Oms-2) Having the Hollandite Structure. Chemistry of Materials. 1994. 6(6):815-821. 63 Ohzuku, T., Kitagawa, M., Sawai, K. and Hirai, T. Topotactic Reduction of -Manganese(II) Oxide in Nonaqueous Lithium Cells. Journal of the Electrochemical Society. 1991. 138(2): 360-365. 64. Rossouw, M. H., Liles, D. C., Thackeray, M., David, W. I. F. and Hull, S. Manganese Dioxide for Lithium Batteries: A Structural and Electrochemical Study. Materials Research Bulletin. 1992. 27(2): 221-230. 65. Yamamoto, N. and Tamada, O. Hydrothermal Crystal Growth of a New Manganese Dioxide. Journal of Crystal Growth. 1985. 73(1): 199-200. 66. Zhang, Q., Luo, J., Vileno, E. and Suib, S. L. Synthesis of CryptomelaneType Manganese Oxides by Microwave Heating. Chemistry of Materials. 1997. 9(10): 2090-2095. 67. Vileno, E., Ma, Y., Zhou, H. and Suib, S. L. Facile Synthesis of Synthetic Todorokite (OMS-1), Co-Precipitation Reactions in the Presence of a Microwave Field. Microporous and Mesoporous Materials. 1998. 20(1-3): 315. 68. Chen, X., Shen, Y.-F., Suib, S. L. and O'Young, C. L. Catalytic Decomposition of 2-Propanol over Different Metal-Cation-Doped OMS-2 Materials. Journal of Catalysis. 2001. 197(2): 292-302. 177 69. Choudary, B. M., Kantam, M. L. and Santhi, P. L. New and Ecofriendly Options for the Production of Speciality and Fine Chemicals. Catalysis Today. 2000. 57(1-2):17-32. 70 Nicolas-Tolentino, E., Tian, Z. R., Zhou, H., Xia, G. and Suib, S. L. Effects of Cu2+ Ions on the Structure and Reactivity of Todorokite- and Cryptomelane-Type Manganese Oxide Octahedral Molecular Sieves. Chemistry of Materials. 1999. 11(7): 1733-1741. 71. Kumar, R., Garces, L. J., Son, Y.-C., Suib, S. L. and Malz, J. R. E. Manganese Oxide Octahedral Molecular Sieve Catalysts for Synthesis of 2Aminodiphenylamine. Journal of Catalysis. 2005. 236(2): 387-391. 72. Abecassis-Wolfovich, M., Jothiramalingam, R., Landau, M. V., Herskowitz, M., Viswanathan, B. and Varadarajan, T. K. Cerium Incorporated Ordered Manganese Oxide OMS-2 Materials: Improved Catalysts for Wet Oxidation of Phenol Compounds. Applied Catalysis B: Environmental. 2005. 59(1-2): 91-98. 73. Li, W. N., Yuan, J., Gomez-Mower, S., Sithambaram, S. and Suib, S. L. Synthesis of Single Crystal Manganese Oxide Octahedral Molecular Sieve (OMS) Nanostructures with Tunable Tunnels and Shapes. Journal of Physical Chemistry B. 2006. 110(7): 3066-3070. 74. Nur, H., Ng Yun, H., Misnon, I. I., Hamdan, H. and Muhid, M. N. M. Hydrophobic Fluorinated TiO2-ZrO2 as Catalyst in Epoxidation of 1-Octene with Aqueous Hydrogen Peroxide. Materials Letters. 2006. 60(17-18):22742277. 76. Nur, H., Ikeda, S. and Ohtani, B. Phase-Boundary Catalysis: A New Approach in Alkene Epoxidation with Hydrogen Peroxide by Zeolite Loaded with Alkylsilane-Covered Titanium Oxide. Chemical Communications. 2000:2235-2236. 77. Ikeda, S., Nur, H., Sawadaishi, T., Ijiro, K., Shimomura, M. and Ohtani, B. Direct Observation of Bimodal Amphiphilic Surface Structures of Zeolite Particles for a Novel Liquid-Liquid Phase Boundary Catalysis. Langmuir. 2001. 17:7976-7979. 79. Arata, K. and Hino, M. Preparation of Superacids by Metal Oxides and Their Catalytic Action. Materials Chemistry and Physics. 1990. 26(3-4):213-237. 80. Ding, Y. S., Shen, X. F., Sithambaram, S., Gomez, S., Kumar, R., Crisostomo, V. M. B., Suib, S. L. and Aindow, M. Synthesis and Catalytic Activity of Cryptomelane-Type Manganese Dioxide Nanomaterials Produced by a Novel Solvent-Free Method. Chemistry of Materials. 2005. 17(21): 5382-5389. 81. Malinger, K. A., Ding, Y.-S., Sithambaram, S., Espinal, L., Gomez, S. and Suib, S. L. Microwave Frequency Effects on Synthesis of Cryptomelane- 178 Type Manganese Oxide and Catalytic Activity of Cryptomelane Precursor. Journal of Catalysis. 2006. 239(2): 290-298. 82. Xingfu, T., Xiumin, H., Jianjun, S., Junlong, L., Yonggang, L., Yide, X. and Wenjie, S. Synthesis and Catalytic Performance of Manganese Oxide Octahedral Molecular Sieve Nanorods for Formaldehyde Oxidation at Low Temperature. Chinese Journal of Catalysis. 2006. 27(2): 97-99. 83. Frías, D., Nousir, S., Barrio, I., Montes, M., López, T., Centeno, M. A. and Odriozola, J. A. Synthesis and Characterization of Cryptomelane- and Birnessite-Type Oxides: Precursor Effect. Materials Characterization. 2007. 58(8-9): 776-781. 84. Gandhe, A. R., Rebello, J. S., Figueiredo, J. L. and Fernandes, J. B. Manganese Oxide OMS-2 as an Effective Catalyst for Total Oxidation of Ethyl Acetate. Applied Catalysis B: Environmental. 2007. 72(1-2): 129-135. 85. Punniyamurthy, T., Velusamy, S. and Iqbal, J. Recent Advances in Transition Metal Catalyzed Oxidation of Organic Substrates with Molecular Oxygen. Chemical Reviews. 2005. 105(6):2329-2364. 86 Sheldon, R. A. and Kochi, J. K. Metal-Catalyzed Oxidation of Organic Compounds. Academic Press, New York. 1981. 87 Blatter, F., Sun, H., Vasenkov, S. and Frei, H. Photocatalyzed Oxidation in Zeolite Cages. Catalysis Today. 1998. 41(4): 297-309. 88. Lahtinen, P., Lankinen, E., Leskela, M. and Repo, T. Insight into Copper Oxidation Catalysts: Kinetics, Catalytic Active Species and Their Deactivation. Applied Catalysis A: General. 2005. 295(2): 177-184. 89. Langhendries, G., De Vos, D. E., Sels, B. F., Vankelecom, I., Jacobs, P. A. and Baron, G. V. Clean Catalytic Technology for Liquid Phase Hydrocarbon Oxidation. Clean Technologies and Environmental Policy. 1998. 1(1): 21-29. 90. Sheldon, R. A. and Downing, R. S. Heterogeneous Catalytic Transformations for Environmentally Friendly Production. Applied Catalysis A: General. 1999. 189(2): 163-183. 91 Corma, A., Esteve, P. and Martínez, A. Solvent Effects During the Oxidation of Olefins and Alcohols with Hydrogen Peroxide on Ti-Beta Catalyst: The Influence of the Hydrophilicity-Hydrophobicity of the Zeolite. Journal of Catalysis. 1996. 161(1): 11-19. 92. Adam, W., Saha-Moller, C. R. and Ganeshpure, P. A. Synthetic Applications of Nonmetal Catalysts for Homogeneous Oxidations. Chemical Reviews. 2001. 101(11): 3499-3548. 93. De Vos, D. E., Dams, M., Sels, B. F. and Jacobs, P. A. Ordered Mesoporous and Microporous Molecular Sieves Functionalized with Transition Metal 179 Complexes as Catalysts for Selective Organic Transformations. Chemical Reviews. 2002. 102(10): 3615-3640. 94. Hashmi, A. S. K. Gold-Catalyzed Organic Reactions. Chemical Reviews. 2007. 107(7): 3180-3211. 95. Hirao, T. Vanadium in Modern Organic Synthesis. Chemical Reviews. 1997. 97(8): 2707-2724. 96 Mizuno, N. and Misono, M. Heterogeneous Catalysis. Chemical Reviews. 1998. 98(1): 199-218. 97. Sheldon, R. A. and van Bekkum, H. Fine Chemicals through Heterogeneous Catalyst. Wiley-VCH, Weinheim. 2001. 98. Besson, M., Bonnet, M. C., Gallezot, P., Tkatchenko, I. and Tuel, A. Catalysis for Fine Chemicals: Towards Specificity with Polyphasic Media. Catalysis Today. 1999. 51(3-4): 547-560. 99. Raja, R., Sankar, G. and Thomas, J. M. New Catalysts for the Aerobic Selective Oxidation of Hydrocarbons: Mn(III)- and Co(III)-Containing Molecular Sieves for the Epoxidation of Alkenes. Chemical Communications. 1999(9): 829-830. 100. Thomas, J. M., Raja, R., Sankar, G., Bell, R. G. and Lewis, D. W. Benign by Design: New Catalysts for an Environmentally Conscious Age. Pure and Applied Chemistry. 2001. 73(7): 1087-1101. 101. Tatsumi, T., Nakamura, M., Negishi, S. and Tominaga, H. Shape-Selective Oxidation of Alkanes with H2O2 Catalysed by Titanosilicate. Journal of the Chemical Society, Chemical Communications. 1990(6): 476-477. 102. Camblor, M. A., Corma, A., Martínez, A. and Pérez-Pariente, J. Synthesis of a Titaniumsilicoaluminate Isomorphous to Zeolite Beta and Its Application as a Catalyst for the Selective Oxidation of Large Organic Molecules. Journal of the Chemical Society, Chemical Communications. 1992(8): 589-590. 103. Perego, C., Carati, A., Ingallina, P., Mantegazza, M. A. and Bellussi, G. Production of Titanium Containing Molecular Sieves and Their Application in Catalysis. Applied Catalysis A: General. 2001. 221: 63-72. 104. Clerici, M. G. Zeolites for Fine Chemicals Production. Topics in Catalysis. 2000. 13(4): 373-386. 105. Clerici, M. G. and Ingallina, P. Epoxidation of Lower Olefins with Hydrogen Peroxide and Titanium Silicalite. Journal of Catalysis. 1993. 140(1): 71-83. 106. Kholdeeva, O. A. and Trukhan, N. N. Mesoporous Titanium Silicates as Catalysts for the Liquid-Phase Selective Oxidation of Organic Compounds. Russian Chemical Reviews. 2006. 75(5):411-432. 180 107. Gao, H., Lu, G., Suo, J. and Li, S. Epoxidation of Allyl Chloride with Hydrogen Peroxide Catalyzed by Titanium Silicalite 1. Applied Catalysis A: General. 1996. 138(1): 27-38. 108. Corma, A. and García, H. Lewis Acids as Catalysts in Oxidation Reactions: From Homogeneous to Heterogeneous Systems. Chemical Reviews. 2002. 102(10): 3837-3892. 109. Chen, L. Y., Chuah, G. K. and Jaenicke, S. Propylene Epoxidation with Hydrogen Peroxide Catalyzed by Molecular Sieves Containing Framework Titanium. Journal of Molecular Catalysis A: Chemical. 1998. 132(2-3): 281292. 110. Hutchings, G. J., Firth, P. G., Lee, D. F., McMorn, P., Bethell, D., Page, P. C. B., King, F. and Hancock, F. Shape Selective Epoxidation of Crotyl Alcohol with H2O2 in the Presence of TS-1. Studies in Surface Science and Catalysis, Vol. 110, 1997. pp. 535-544. 111. Tatsumi, T. Metallozeolites and Applications in Catalysis. Current Opinion in Solid State and Materials Science. 1997. 2(1): 76-83. 112. Lane, B. S. and Burgess, K. Metal-Catalyzed Epoxidations of Alkenes with Hydrogen Peroxide. Chemical Reviews. 2003. 103(7): 2457-2473. 113. Arata, K. Preparation of Superacids by Metal Oxides for Reactions of Butanes and Pentanes. Applied Catalysis A: General. 1996. 146(1):3-32. 114. Corma, A. Solid Acid Catalysts. Current Opinion in Solid State and Materials Science. 1997. 2(1): 63-75. 115. Chu, W., Hu, J., Xie, Z. and Chen, Q. Design and Elaboration of New Solid Acids for the Synthesis of Butylacetate. Catalysis Today. 2004. 90(3-4): 349353. 116. Hino, M. and Arata, K. Catalytic Activity of Iron Oxide Treated with Sulfated Ion for Dehydration of 2-Propanol and Ethanol and Polymerization of Isobutyl Vinyl Ether. Chemistry Letters. 1979: 477. 117. Farcasiu, D., Ghenciu, A. and Li, J. Q. The Mechanism of Conversion of Saturated Hydrocarbons Catalyzed by Sulfated Metal Oxides: Reaction of Adamantane on Sulfated Zirconia. Journal of Catalysis. 1996. 158(1): 116127. 118. Wissler, M. C., Jagusch, U. P., Sundermann, B. and Hoelderich, W. F. OnePot Synthesis of a New Potential Analgesic over Bifunctional Palladium/Amberlyst Catalysts. Catalysis Today. 2007. 121(1-2): 6-12. 119. Thomas, J. M. and Raja, R. Design of a "Green" One-Step Catalytic Production Of -Caprolactam (Precursor of Nylon-6). Proceedings of the 181 National Academy of Sciences of the United States of America. 2005. 102(39): 13732-13736. 120. Moore, T., Adhikari, R. and Gunatillake, P. Chemosynthesis of Bioresorbable Poly(-Butyrolactone) by Ring-Opening Polymerisation: A Review. Biomaterials. 2005. 26(18): 3771-3782. 121. Pawlowski, P. and Rokicki, G. Synthesis of Oligocarbonate Diols from Ethylene Carbonate and Aliphatic Diols Catalyzed by Alkali Metal Salts. Polymer. 2004. 45(10): 3125-3137. 122. Burkhardt, E. R. and Matos, K. Boron Reagents in Process Chemistry: Excellent Tools for Selective Reductions. Chemical Reviews. 2006. 106(7):2617-2650. 123. Lorbach, V., Franke, D., Nieger, M. and Múller, M. Cyclohexadiene-TransDiols as Versatile Starting Material in Natural Product Synthesis: Short and Efficient Synthesis of Iso-Crotepoxide and Ent-Senepoxide. Chemical Communications. 2002(5): 494-495. 124. Ono, M. O.,Takasago, S. T., Kakogawa, H. T. and Akashi, J. H. Manufacturing Method of Optically-Active 1,2 Diols. U.S. Patent: 4,981,796. 1991 125. Döbler, C., Mehltretter, G. M., Sundermeier, U. and Beller, M. Dihydroxylation of Olefins Using Air as the Terminal Oxidant. Journal of Organometallic Chemistry. 2001. 621(1-2): 70-76. 126. Gelbard, G. Epoxidation of Alkenes with Tungsten Catalysts Immobilised on Organophosphoryl Macroligands. Comptes Rendus de l'Academie des Sciences - Series IIc: Chemistry. 2000. 3(10): 757-764 127. Gelbard, G., Breton, F., Quenard, M. and Sherrington, D. C. Epoxidation of Cyclohexene with Polymethacrylate-Based Peroxotungstic Catalysts. Journal of Molecular Catalysis A: Chemical. 2000. 153(1-2): 7-18. 128. Gelbard, G., Gauducheau, T., Vidal, E., Parvulescu, V. I., Crosman, A. and Pop, V. M. Epoxidation with Peroxotungstic Acid Immobilised onto SilicaGrafted Phosphoramides. Journal of Molecular Catalysis A: Chemical. 2002. 182-183: 257-266. 129. Nakamura, K., Yamanaka, R., Matsuda, T. and Harada, T. Recent Developments in Asymmetric Reduction of Ketones with Biocatalysts. Tetrahedron: Asymmetry. 2003. 14(18): 2659-2681. 130. Misawa, N., Nakamura, R., Kagiyama, Y., Ikenaga, H., Furukawa, K. and Shindo, K. Synthesis of Vicinal Diols from Various Arenes with a Heterocyclic, Amino or Carboxyl Group by Using Recombinant Escherichia Coli Cells Expressing Evolved Biphenyl Dioxygenase and Dihydrodiol Dehydrogenase Genes. Tetrahedron. 2005. 61(1): 195-204. 182 131. Van Der Waal, J. C., Creyghton, E. J., Kunkeler, P. J., Tan, K. and Van Bekkum, H. Beta-Type Zeolites as Selective and Regenerable Catalysts in the Meerwein-Ponndorf-Verley Reduction of Carbonyl Compounds. Topics in Catalysis. 1997. 4(3-4):261-268. 132. Krijnen, S., Sánchez, P., Jakobs, B. T. F. and van Hooff, J. H. C. A Controlled Post-Synthesis Route to Well-Defined and Active Titanium Beta Epoxidation Catalysts. Microporous and Mesoporous Materials. 1999. 31(12): 163-173. 133. Sasidharan, M., Hegde, S. G. and Kumar, R. Surface Acidity of Al-, Ga- and Fe-Silicate Analogues of Zeolite NCL-1 Characterized by FTIR, TPD (NH3) and Catalytic Methods. Microporous and Mesoporous Materials. 1998. 24(13): 59-67. 134. Trong On, D., Nguyen, S. V., Hulea, V., Dumitriu, E. and Kaliaguine, S. Mono- and Bifunctional MFI, BEA and MCM-41 Titanium-Molecular Sieves. Part 1. Synthesis and Characterization. Microporous and Mesoporous Materials. 2003. 57(2): 169-180. 135. Niemantsverdriet, J. W. Spectroscopy in Catalysis: An Introduction. VCH, Weinheim 1995. 136. Richards, R. Surface and Nanomolecular Catalysis. Taylor & Francis, Florida, 2006. 137. Lagalante, A. F. Atomic Absorption Spectroscopy: A Tutorial Review. Applied Spetroscopy Reviews 1999. 34(3):173–189. 138. Chatwal, P. and Anand, G. Spectroscopy (Atomic and Molecular). Himalaya, Bombay. 1985. 139. Williams, D. H. and Fleming, I. Spectroscopic Methods in Organic Chemistry. McGraw-Hill, London 1995. 140. Leofanti, G., Padovan, M., Tozzola, G. and Venturelli, B. Surface Area and Pore Texture of Catalysts. Catalysis Today. 1998. 41(1-3): 207-219. 141 Watt. The Principles and Practice of Electron Miscroscopy. Cambridge University Pr, Cambridge 1985. 142. Gfroerer. Photoluminescence in Analysis of Surfaces and Interfaces. John Wiley & Sons Ltd, Chichester 2000. 143. Endud, S. Secondary Aluminated Mesoporous Molecular Sieves Al-MCM-41: Synthesis, Characterization and Catalytic Activity. PhD Thesis. Universiti Teknologi Malaysia. 2000. 144. ten Brink, G.-J., Arends, I. W. C. E., Papadogianakis, G. and Sheldon, R. A. Catalytic Conversions in Water: Part 13. Aerobic Oxidation of Olefins to 183 Methyl Ketones Catalysed by a Water-Soluble Palladium Complex Mechanistic Investigations. Applied Catalysis A: General. 2000. 194195:435-442. 145. Marko, I. E., Giles, P. R., Tsukazaki, M., Brown, S. M. and Urch, C. J. Copper-Catalyzed Oxidation of Alcohols to Aldehydes and Ketones: An Efficient, Aerobic Alternative. Science. 1996. 274(5295): 2044-2046. 146. Cicco, S. R., Latronico, M., Mastrorilli, P., Suranna, G. P. and Nobile, C. F. Homogeneous and Heterogeneous Catalytic Oxidation of Benzylic and Secondary Alcohols with a Metal Dioxygenato Complex in the Presence of 2Methylpropanal and Dioxygen. Journal of Molecular Catalysis A: Chemical. 2001. 165(1-2): 135-140. 147. Lee, A. F., Gee, J. J. and Theyers, H. J. Aspects of Allylic Alcohol Oxidation: A Bimetallic Heterogeneous Selective Oxidation Catalyst. Green Chemistry. 2000. 2(6): 279-282. 148. Chen, K., Khodakov, A., Yang, J., Bell, A. T. and Iglesia, E. Isotopic Tracer and Kinetic Studies of Oxidative Dehydrogenation Pathways on Vanadium Oxide Catalysts. Journal of Catalysis. 1999. 186(2): 325-333. 149. Soenen, V., Herrmann, J. M. and Volta, J. C. In Situ Electrical Characterization of Magnesium Vanadate Reference Phases (Meta-MgV2O6, Pyro-Mg2V2O7, and Ortho-Mg3V2O8) Used in Oxidative Dehydrogenation of Propane to Propene. Journal of Catalysis. 1996. 159(2): 410-417. 150. Sedmak, G., Hocévar, S. and Levec, J. Kinetics of Selective Co Oxidation in Excess of H2 over the Nanostructured Cu0.1Ce0.9O2-Y Catalyst. Journal of Catalysis. 2003. 213(2): 135-150. 151. Mazumder, B. and Védrine, J. C. Oxidative Dehydrodimerisation and Aromatisation of Isobutene on Bi2O3-SnO2 Catalysts. Applied Catalysis A: General. 2003. 245(1): 87-102. 152. Mazumder, B. and Védrine, J. C. Kinetics Study of the Oxidative Dehydrodimerisation and Aromatisation of Isobutene on Bi2O3-SnO2. Journal of Catalysis. 2003. 214(2): 284-291. 153. Karasevich, E.I., Kulikova, V.S., Shilov, A.E., Shteinman, A.A. Biomimetic Alkane Oxidation Involving Metal Complexes. Russian Chemical Reviews. 1998. 67(4): 335. 154. Sawatari, N., Yokota, T., Sakaguchi, S., Ishii, Y. Alkane Oxidation with Air Catalyzed by Lipophilic N-Hydroxyphthalimides without Any Solvent. Journal of Organic Chemistry. 2001. 66(23): 7889-7891. 155. Yao, W., Chen, Y., Min, L., Fang, H., Yan, Z., Wang, H., and Wang, J. Liquid Oxidation of Cyclohexane to Cyclohexanol over Cerium-Doped Mcm-41. Journal of Molecular Catalysis A: Chemical. 2005. 246: 161–165. 184 156. Wang, Z., Martell ,A.E., Motekaitis, R.J., and Reibenspies, J.H. Dinuclear Fe(II) Macrocyclic Complexes as Catalysts for the Oxidation of Hydrocarbons by Molecular Oxygen. Inorganica Chimica Acta. 2000. 300– 302: 378–383. 157. Schuchardt U., P., R. and Rufo, M. Iron III and Copper II Catalyzed Cyclohexane Oxidation by Molecular Oxygen in the Presence of tert-Butyl Hydroperoxide. Journal of Molecular Catalysis A: Chemical. 1998. 135: 257–262. 158. Pires, E. L., Wallau, M. and Schuchardt, U. Cyclohexane Oxidation over Rare Earth Exchanged Zeolite Y. Journal of Molecular Catalysis A: Chemical. 1998. 136: 69–74. 159. Luna, F. J., Ukawa, S. E. Wallau, M. and Schuchardt, U. Cyclohexane Oxidation Using Transition Metal-Containing Aluminophosphate (MAPOVFI). Journal of Molecular Catalysis A: Chemical. 1997. 117:405-411. 160. Schuchardt, U., Cardoso, D., Sercheli, R., Pereira, R., da Cruz, R. S., Guerreiro M. C., Mandelli D., Spinacé E.V., and Pires E.L. Cyclohexane Oxidation Continues to Be a Challenge. Applied Catalysis A: General. 2001. 211: 1–17. 161. Mayhew, M. P., Reipa, V., Holden, M.J. and Vilker, V.L. Improving the Cytochrome P450 Enzyme System for Electrode-Driven Biocatalysis of Styrene Epoxidation. Biotechnology Progress. 2000. 16: 610-616. 162. Tong, X., Xu, J., Miao, H., Yang, G., Ma, H., Zhang, Q. Highly Efficient and Metal-Free Oxidation of Olefins by Molecular Oxygen under Mild Conditions. Tetrahedron. 2007. 63(32): 7634-7639. 163. Kumar S. B., M. S. P., Pais G. C. G., Kumar P. and Kumar R. Epoxidation of Styrene over a Titanium Silicate Molecular Sieve TS-1 Using Dilute H2O2 as Oxidizing Agent. Journal of Catalysis. 1995. 156(1):163-166. 164. Uguina, M. A., Serrano, D.P., Sanz, R., Fierro, J.L.G., López-Granados, M., Mariscal, R. . Preliminary Study on the TS-1 Deactivation During Styrene Oxidation with H2o2. Catalysis Today. 2000. 61(1): 263-270. 165. Hulea, V. a. D., E. Styrene Oxidation with H2O2 over Ti-Containing Molecular Sieves with MFI, BEA and MCM-41 Topologies. Applied Catalysis A: General. 2004. 277(1-2):99-106. 166. Nie, L., Xin, K.K., Li, W.S., Zhou, X.P. Benzaldehyde Synthesis via Styrene Oxidation by O2 over TiO2 and TiO2/SiO2. Catalysis Communications. 2007. 8: 488–492. 167. Lorbach, V., Franke, D., Nieger, M. and Müller, M. Cyclohexadiene-TransDiols as Versatile Starting Material in Natural Product Synthesis: Short and 185 Efficient Synthesis of Iso-Crotepoxide and Ent-Senepoxide. Chemical Communications. 2002(5): 494-495. 168. Database of Joint Committee for Powder Diffraction Studies (JPCDS) PDF# 29,102. 169. Kim, H. S. and Stair, P. C. Bacterially Produced Manganese Oxide and Todorokite: UV Raman Spectroscopic Comparison. Journal of Physical Chemistry B. 2004. 108(44): 17019-17026. 170. Potter, R. M. and Rossman, G. R. The Tetravalent Manganese Oxides: Identification, Hydration, and Structural Relationships by Infrared Spectroscopy. American Mineralogist. 1979. 64: 1199-1218. 171. Luo, J., Huang, A., Park, S. H., Suib, S. L. and O'Young, C. L. Crystallization of Sodium-Birnessite and Accompanied Phase Transformation. Chemistry of Materials. 1998. 10(6): 1561-1568. 172. O'Young, C.-L., Sawicki, R. A. and Suib, S. L. Micropore Size Distribution of Octahedral Molecular Sieves (OMS). Microporous Materials. 1997. 11(12): 1-8. 173. Pirson, A., Mohsine, A., Marchot, P., Michaux, B., van Cantfort, O., Pirard, J. P. and Lecloux, A. J. Synthesis of SiO2-TiO2 Xerogels by Sol-Gel Process. Journal of Sol-Gel Science and Technology. 1995. 4(3): 179-185. 174. Kang, L., Zhang, M., Liu, Z. H. and Ooi, K. IR Spectra of Manganese Oxides with Either Layered or Tunnel Structures. Spectrochimica Acta - Part A: Molecular and Biomolecular Spectroscopy. 2007. 67(3-4): 864-869. 175. Fujihara, K., Izumi, S., Ohno, T. and Matsumura, M. Time-Resolved Photoluminescence of Particulate TiO2 Photocatalysts Suspended in Aqueous Solutions. Journal of Photochemistry and Photobiology A: Chemistry. 2000. 132(1-2): 99-104. 176. Gallardo-Amores, J. M., Armaroli, T., Ramis, G., Finocchio, E. and Busca, G. A Study of Anatase-Supported Mn Oxide as Catalysts for 2-Propanol Oxidation. Applied Catalysis B: Environmental. 1999. 22(4):249-259. 177. Mayer, J. T., Diebold, U., Madey, T. E. and Garfunkel, E. Titanium and Reduced Titania Overlayers on Titanium Dioxide(110). Journal of Electron Spectroscopy and Related Phenomena. 1995. 73(1):1-11. 178. Zou, Z.-Q., Meng, M., Guo, L.-H. and Zha, Y.-Q. Synthesis and Characterization of CuO/Ce1-XTixO2 Catalysts Used for Low-Temperature Co Oxidation. Journal of Hazardous Materials. 2009. 163(2-3):835-842. 179. Arillo, M. A., López, M. L., Pico, C., Veiga, M. L., Jiménez-López, A. and Rodríguez-Castellón, E. Surface Characterisation of Spinels with Ti(Iv) 186 Distributed in Tetrahedral and Octahedral Sites. Journal of Alloys and Compounds. 2001. 317-318:160-163. 180 Ikeda, S., Nur, H., Sawadaishi, T., Ijiro, K., Shimomura, M. and Ohtani, B. Direct Observation of Bimodal Amphiphilic Surface Structures of Zeolite Particles for a Novel Liquid-Liquid Phase Boundary Catalysis. Langmuir. 2001. 17(26): 7976-7979. 181. Nur, H., Ikeda, S. and Ohtani, B. Phase-Boundary Catalysis: A New Approach in Alkene Epoxidation with Hydrogen Peroxide by Zeolite Loaded with Alkylsilane-Covered Titanium Oxide. Chemical Communications. 2000(22): 2235-2236. 182. Nur, H., Ikeda, S. and Ohtani, B. Phase-Boundary Catalysis of Alkene Epoxidation with Aqueous Hydrogen Peroxide Using Amphiphilic Zeolite Particles Loaded with Titanium Oxide. Journal of Catalysis. 2001. 204(2):402-408. 183. Nur, H., Hau, N. Y., Muhid, M. N. M. and Hamdan, H. Surface Structure of Alkylsilylated HZSM-5 as Phase-Boundary Catalyst. Physics Journal of the Indonesia Physical Society. 2004. A7(0218):1-5. 184. Tatsumi, T., Matsuhashi, H. and Arata, K. A Study of the Preparation Procedures of Sulfated Zirconia Prepared from Zirconia Gel: The Effect of the pH of the Mother Solution on the Isomerization Activity of n-Pentane. Bulletin of the Chemical Society of Japan. 1996. 69(5): 1191-1194. 185. Arata, K., Hino, M. and Yamagata, N. Acidity and Catalytic Activity of Zirconium and Titanium Sulfates Heat-Treated at High Temperature: Solid Superacid Catalysts. Bulletin of the Chemical Society of Japan. 1990. 63(1): 244-246. 186. Selvaraj, M., Sinha, P. K. and Pandurangan, A. Synthesis of Dypnone Using SO42-/Al-MCM-41 Mesoporous Molecular Sieves. Microporous and Mesoporous Materials. 2004. 70(1-3): 81-91. 187. Wang, P., Yang, S., Kondo, J. N., Domen, K., Yamada, T. and Hattori, H. Spectroscopic Study of H2 and Co Adsorption on Platinum-Promoted Sulfated Zirconia Catalysts. Journal of Physical Chemistry B. 2003. 107(43): 11951-11959. 188. Jung, S. M. and Grange, P. Characterization and Reactivity of Pure TiO2SO42- SCR Catalyst: Influence of SO42- Content. Catalysis Today. 2000. 59(3): 305-312. 189. Spielbauer, D., Mekhemer, G. A. H., Bosch, E. and Knözinger, H. n-Butane Isomerization on Sulfated Zirconia. Deactivation and Regeneration as Studied by Raman, Uv-Vis Diffuse Reflectance and ESR Spectroscopy. Catalysis Letters. 1996. 36(1-2):59-68. 187 190. Babou, F., Coudurier, G. and Vedrine, J. C. Acidic Properties of Sulfated Zirconia: An Infrared Spectroscopic Study. Journal of Catalysis. 1995. 152(2): 341-349. 191. Laperdrix, E., Sahibed-Dine, A., Costentin, G., Saur, O., Bensitel, M., Nédez, C., Mohamed Saad, A. B. and Lavalley, J. C. Reduction of Sulfate Species by H2S on Different Metal Oxides and Promoted Aluminas. Applied Catalysis B: Environmental. 2000. 26(2):71-80. 192. Kurosaka, T., Matsuhashi, H. and Arata, K. Dehydrogenative Coupling of Methane Catalyzed by Platinum-Added Sulfated Zirconia and Characterization of the Catalyst Surface. Journal of Catalysis. 1998. 179(1): 28-35. 193. Corma, A. and García, H. Lewis Acids: From Conventional Homogeneous to Green Homogeneous and Heterogeneous Catalysis. Chemical Reviews. 2003. 103(11): 4307-4365. 194. Grootendorst, E. J., Verbeek Y., Ponec V. . The Role of the Mars and Van Krevelen Mechanism in the Selective Oxidation of Nitrosobenzene and the Deoxygenation of Nitrobenzene on Oxidic Catalysts. Journal of Catalysis. 1995. 157(2):706-712. 195. Yin, Y. G., Xu, W. Q., DeGuzman, R., Suib, S. L. and O'Young, C. L. Studies of Stability and Reactivity of Synthetic Cryptomelane-Like Manganese Oxide Octahedral Molecular Sieves. Inorganic Chemistry. 1994. 33(19): 4384-4389. 196. Doornkamp, C., Clement, M., and Ponec, V. The Isotopic Exchange Reaction of Oxygen on Metal Oxides. Journal of Catalysis. 1999. 182(2): 390-399. 197. Nur, H. Tailoring of Novel Metal Substituted AlPO4-5 Molecular Sieves as Potential Catalysts for Conversion of Alcohols. PhD Thesis. Universiti Teknologi Malaysia. 1998. 198. Larsen, R. G., Saladino, A.C., Hunt, T.A., Mann, J.E., Xu, M., Grassian V.H. and Larsen, S.C. A Kinetic Study of the Thermal and Photochemical Partial Oxidation of Cyclohexane with Molecular Oxygen in Zeolite Y. Journal of Catalysis. 2001. 204: 440 - 449. 199. Tian, P., Liu, Z., Wu, Z., Xu, L. and He, Y. Characterization of MetalContaining Molecular Sieves and Their Catalytic Properties in the Selective Oxidation of Cyclohexane. Catalysis Today. 2004. 93-95: 735–742. 200. Post, J. E., Von Dreele, R. B. and Buseck, P. R. Symmetry and Cation Displacements in Hollandites: Structure Refinements of Hollandite, Cryptomelane and Priderite. Acta Crystallographica B. 1982. 38(4): 10561065. 188 201. Sun, Z., Xu, J., Du, Z., Zhang, W. Decomposition of Cyclohexyl Hydroperoxide over Transition Metal-Free Zeolite H-Beta. Applied Catalysis A: General. 2007. 323: 119-125. 202. Wang, J.-Y., Xia, G.-G., Yin, Y.-G., Suib, S. L. and O'Young, C. L. Cyclohexane Functionalization Catalyzed by Octahedral Molecular Sieve (OMS-1) Materials. Journal of Catalysis. 1998. 176(2): 275-284. 203. Tian, P., Liu, Z., Wu, Z., Xu, L., He, Y. Metal-Containing Molecular Sieves as Catalysts for the Selective Oxidation of Cyclohexane by Oxygen. In Studies in Surface Science and Catalysis. 2004. 154C: 2674-2680. 204. Salavati-Niasari, S. M., and Banitaba, S.H. Alumina-Supported Mn(II), Co(II), Ni(II) and Cu(II) Bis(2-Hydroxyanil)Acetylacetone Complexes as Catalysts for the Oxidation of Cyclohexene with tert-Butylhydroperoxide. Journal of Molecular Catalysis A: Chemical. 2003. 201(1-2): 43-54. 205. Salavati-Niasari, M. Nanoscale Microreactor-Encapsulation 14-Membered Nickel(II) Hexamethyl Tetraaza: Synthesis, Characterization and Catalytic Activity. Journal of Molecular Catalysis A: Chemical. 2005. 229(1-2): 159164. 206. Salavati-Niasari, M., Salemi, P. and Davar, F. Oxidation of Cyclohexene with Tert-Butylhydroperoxide and Hydrogen Peroxide Catalysted by Cu(II), Ni(II), Co(II) and Mn(II) Complexes of N,Nಿ-Bis-(-Methylsalicylidene)-2,2Dimethylpropane-1,3-Diamine, Supported on Alumina. Journal of Molecular Catalysis A: Chemical. 2005. 238(1-2):215-222. 207. Nur, H., Prasetyoko, D., Ramli, Z. and Endud, S. Sulfation: A Simple Method to Enhance the Catalytic Activity of TS-1 in Epoxidation of 1-Octene with Aqueous Hydrogen Peroxide. Catalysis Communications. 2004. 5(12): 725728. 208. Zhuang, J., Ma, D., Yan, Z., Liu, X., Han, X., Bao, X., Zhang, Y., Guo, X. and Wang, X. Effect of Acidity in TS-1 Zeolites on Product Distribution of the Styrene Oxidation Reaction. Applied Catalysis A: General. 2004. 258(1): 1-6. Available online at www.sciencedirect.com Catalysis Communications 8 (2007) 2007–2011 www.elsevier.com/locate/catcom On the location of different titanium sites in Ti–OMS-2 and their catalytic role in oxidation of styrene Hadi Nur *, Fitri Hayati, Halimaton Hamdan Ibnu Sina Institute for Fundamental Science Studies, Universiti Teknologi Malaysia, 81310 UTM Skudai, Johor, Malaysia Received 20 December 2006; received in revised form 30 March 2007; accepted 2 April 2007 Available online 10 April 2007 Abstract Octahedral manganese oxide molecular sieves (OMS-2) modified by impregnation of TiO2 exhibit a higher catalytic activity for oxidation of styrene with tert-butylhydroperoxide in comparison to titanium-incorporated OMS-2, where the styrene conversions were ca. 70% and 45–50%, respectively. The framework of titanium species has no effect on the enhancement of catalytic activity, while the nonframework of titanium species induces a synergetic effect that enhances the oxidation of styrene with tert-butylhydroperoxide. 2007 Elsevier B.V. All rights reserved. Keywords: Titanium-incorporated octahedral manganese oxide molecular sieves; Non-framework titanium; Oxidation of styrene 1. Introduction Octahedral manganese oxide molecular sieves (OMS-2) are currently considered as one of the potential catalysts in oxidation of alcohols and olefins [1–3]. OMS-2 materials are manganese oxide (MnOx; x = 1.85–2.00) with a framework structure consisting of 2 · 2 type tunnels, built up of MnO6 octahedra with a pore diameter of 4.6 Å [4]. Although, recently, the oxidation of styrene has been reported by using Me–OMS-2 (Me = Fe, Cu, Ni and Co) [5], it is desirable to study the catalytic performance of OMS-2 material with the addition or incorporation of titanium. An understanding of the synthetic methods of catalyst preparation is needed for the precise control of the structure and location of catalytically active sites on OMS-2. Here, we report on the study of the effect of the location of titanium sites, either in the framework or non-framework of OMS-2, and their catalytic role in the oxidation of styrene with tert-butyl hydroperoxide (TBHP) as the oxidant. X-ray powder diffraction (XRD), photolu* Corresponding author. Tel.: +60 7 5536077; fax: +60 7 5536080. E-mail address: hadi@kimia.fs.utm.my (H. Nur). URL: http://www.ibnusina.utm.my/~hadi (H. Nur). 1566-7367/$ - see front matter 2007 Elsevier B.V. All rights reserved. doi:10.1016/j.catcom.2007.04.002 minescence, surface area and pyridine adsorption measurements were used to characterize these samples. The catalytic activity of Ti containing OMS-2 was also compared with TS-1 as a reference catalyst. 2. Experimental 2.1. Synthesis OMS-2 was prepared by a precipitation method according to [6]. A 0.4 M solution of KMnO4 (13.3 g in 225 ml of deionized water) was added to a mixture of a 1.75 M solution of MnSO4 Æ H2O (19.8 g in 67.5 ml deionized water) and 6.8 ml of concentrated HNO3. The resulting black precipitate was stirred vigorously and refluxed at 373 K for 24 h. The precipitate was filtered and washed with deionized water until neutral pH and dried at 393 K. This gave OMS-2. Titanium incorporated OMS-2 (Ti–OMS-2) was prepared by the stepwise addition of a KMnO4 solution (13.3 g in 225 ml of deionized water) to different amounts of Ti2SO4 (25–75 ml) (15% v/v in H2SO4) in order to produce Ti–OMS-2 in the Ti/Mn ratio of 0.18, 0.43 and 0.67 (as analyzed by using an atomic absorption spectrometer). Upon completion, the mixture was stirred, refluxed, 2008 H. Nur et al. / Catalysis Communications 8 (2007) 2007–2011 filtered, washed, and dried following the above procedure and it was labeled as Ti–OMS-2 (0.18), Ti–OMS-2 (0.43) and Ti–OMS-2 (0.67), where the number in parentheses is the molar ratio of Ti/Mn. Titanium(IV) tetra-2-propoxide ðTiðOPri4 Þ was impregnated from its toluene solution into OMS-2 powder and calcined at 773 K for 3 h. Here, this modified OMS-2 is called Ti–OMS-2 (imp). The molar amount of Ti calculated to give the molar ratio of Ti/Mn was 0.18. For comparison, a mechanical mixture of TiO2 and OMS-2 was prepared by the addition of a calculated amount of TiO2 (rutile) powder to OMS-2 with a Ti/Mn molar ratio of 0.67. This catalyst was labeled as TiO2– OMS-2 (mix). Table 1 summarizes the preparation method, the chemical composition and labeling of titanium containing OMS-2. In this study, TS-1 was also used as a reference catalyst. TS-1 (2 mol% of titanium) was prepared according to a procedure described earlier [7,8]. 2.2. Characterization All samples were characterized by powder XRD for the crystallinity and phase content of the solid materials using a Shimadzu XRD 6000 diffractometer with the CuKa k = 1.5405 Å) radiation as the diffracted monochromatic beam at 30 kV and 30 mA. Atomic absorption analysis (AAS) was employed for elemental composition analyses of manganese and titanium in the sample. A Perkin–Elmer model Analyst 400 spectrophotometer was used to carry out the analyses. The acidity of the solids was characterized by an absorbed base probe molecule. A wafer of the sample (10–12 mg) was locked in the cell equipped with CaF2 windows, and evacuated at 400 C under vacuum condition for 4 h. Pyridine as a probe molecule was introduced into the evacuated sample at room temperature. The IR spectra of the sample were monitored at room temperature after desorption of pyridine at 150 C for 1 h. Photoluminescence spectra were recorded in air at room temperature on a Perkin–Elmer LS 55 spectrometer. The emission spectra observed at an excitation wavelength was 430 nm. 2.3. Catalytic testing Oxidation of styrene was carried out using the above catalysts. Styrene (5 mmol), 70% aqueous ter-butyl hydroperoxide (TBHP) (10 mmol), catalyst (50 mg) and acetonitrile (15 ml) as solvent were placed in a round-bottomed flask with a reflux condenser and the reaction was performed with stirring at 70 C in an oil bath. The products were collected in a period of time and analyzed by GC and GC–MS. 3. Results and discussion XRD patterns of OMS-2, Ti–OMS-2 (0.18) and Ti– OMS-2 (0.43) show that the samples are pure and highly crystalline and matched those of cryptomelane Q [9]; the natural counterpart of OMS-2 material (see Fig. 1). The results confirmed that OMS-2, Ti–OMS-2 (0.18) and Ti– OMS-2 (0.43) materials consist of the cryptomelane structure: 2 · 2 tunnels with a pore size of 4.6 Å, composed of double chains of edge sharing and corner sharing MnO6 octahedra [4]. The absence of other peaks in the XRD patterns except the cryptomelane peaks suggested that Ti was successfully incorporated in the framework of Ti–OMS-2. In order to confirm the successful incorporation of titanium, the XRD patterns of Ti–OMS-2 were compared with the mechanical mixture of TiO2 (rutile) and OMS-2 where the mixture (TiO2–OMS-2 (mix)) showed the presence of the rutile phase (see Fig. 1 and Table 1). If Ti is successfully incorporated in the framework of OMS-2, one expects that the bigger the substituted atom is, the bigger is the unit cell volume. Calculation of the unit cell volume of OMS-2 Table 1 Chemical composition and physicochemical properties of OMS-2, Ti–OMS-2 and TiO2–OMS-2 Samples Molar ratio of Ti/Mna Molar amount of Mnb OMS-2 Ti–OMS-2 (0.18) Ti–OMS-2 (0.43) Ti–OMS-2 (0.67) TiO2–OMS2 (imp)c TiO2–OMS2 (mix)d 0.00 0.18 552.2 412.5 0.43 a b c d e Molar amount of Tib Methods of introduction of Ti species Location of Ti species Structure of nonframework Ti species Surface area (m2 g1) 0.0 74.4 – Direct synthesis – Framework – – 155 152 420.4 181.4 Direct synthesis Framework – 149 0.67 342.0 228.2 Direct synthesis Rutile 152 0.18 552.0 99.4 Impregnation Framework and non-framework Non-framework Amorphous N.d.e 0.67 552.2 369.9 Mechanical mixing Non-framework Rutile N.d.e Analysis was carried out by atomic absorption spectrometer. The amount of OMS-2, Ti–OMS-2 and TiO2–OMS-2 were 50 mg. Titanium(IV) tetra-2-propoxide ðTiðOPri4 Þ was impregnated from its toluene solution into OMS-2 powder and calcined at 773 K for 3 h. Catalyst was prepared by addition of calculated amount of Ti from TiO2 powder to OMS-2. Not determined. H. Nur et al. / Catalysis Communications 8 (2007) 2007–2011 * (g) * * (f) (e) * * * Relative intensity / a.u. * (d) (c) (b) (a) 10 20 30 40 2θ / o 50 60 70 Fig. 1. X-ray diffractograms of (a) cryptomelane (JCPDS 29, 102), (b) OMS-2, (c) Ti–OMS-2 (0.18), (d) Ti–OMS-2 (0.43), (e) Ti–OMS-2 (0.67), (f) TiO2–OMS-2 (imp) and (g) Ti–OMS-2 (mix). (271 Å3) and Ti–OMS-2 (0.43) (277 Å3) shows that the unit cell volumes increase on incorporation of Ti in the framework of OMS-2. The lattice enlargement originates from the replacement of the smaller ionic Mn4+ (ionic radius is 0.53 Å) by the relatively larger ionic Ti4+ (ionic radius is 0.61 Å). This shows a strong evidence of isomorphous substitution of Mn atoms by Ti into the framework of Ti–OMS-2. Moreover, it can be found that with the incorporation of Ti, the full width at half maximum decreases indicating an increase in the grain size. It is observed that the peaks of the rutile phase of TiO2 appear at a Ti/Mn ratio higher than ca. 0.5 and are not observed at a Ti/Mn ratio less than ca. 0.5 (see Fig. 1). By considering the upper limit of the titanium that can be incorporated into the framework, one would expect non-framework titanium species to be formed when the Ti/Mn ratio reached ca. 0.5. The amount of Ti located in the non-framework is ca. 25% in Ti–OMS (0.67). This argument is supported by the presence of the rutile phase of TiO2 in Ti–OMS-2 (0.67) (see Fig. 1). However, no reflection for the rutile phase of TiO2 is observed in TiO2–OMS-2 (imp) where the catalyst is prepared by the impregnation method. This result implies that the structure of TiO2 in TiO2–OMS-2 (imp) is in the amorphous form. The photoluminescence (PL) was also used to confirm the absence of non-framework TiO2, because non-framework TiOx with a very small crystallite size cannot be detected by XRD. The PL spectra are useful to disclose the efficiency of charge carrier trapping, migration and transfer, and to understand the nature of electron–hole pairs in TiO2 semiconductor particles since PL emission results from the recombination of photo-excited free carriers [10]. In this study, the 430 nm excited PL spectra of all pressed-powder samples at room temperature were examined in the range of 560–680 nm. The PL spectra of OMS-2, Ti–OMS-2 (0.43) and TiO2–OMS-2 (mix) are shown in Fig. 2. The results indicated that the photoluminescence intensity of TiO2–OMS-2 (mix) (kem(max) 600 ± 10 nm; FWHM 40 nm) was substantially higher than that of OMS-2 and Ti–OMS-2 (0.43) (see Fig. 2). The relative intensity of Ti–OMS-2 (0.43) and TiO2–OMS-2 (mix) is similar suggesting that there are no TiOx particles exist in Ti–OMS-2 (0.43) sample. This result reinforces our suggestion that Ti is incorporated in the framework of Ti– OMS-2. The incorporation of Ti in the framework of OMS-2 was also further supported by the surface area analysis (see Table 1). It is revealed that the surface area of OMS-2 and Ti–OMS-2 is almost the same. Reaction products of oxidation of styrene using TBHP as the oxidant catalyzed by TiO2 (rutile phase), OMS-2, TiO2–OMS-2 and Ti–OMS-2 were analyzed by GC. The major products in this reaction proved to be benzaldehyde, styrene oxide and phenylacetaldehyde. The selectivities towards benzaldehyde, styrene oxide and phenylacetaldehyde as the reaction products are shown in Fig. 3. As Ti-OMS-2 (0.43) TiO2-OMS-2 (mix) Intensity / a.u. * = TiO2 (rutile) * 2009 OMS-2 560 580 600 620 640 Wavelength / nm 660 680 Fig. 2. Photoluminescence spectra of OMS-2, Ti–OMS-2 (0.43) and TiO2– OMS-2 (mix). The excitation wavelength is 430 nm. 2010 H. Nur et al. / Catalysis Communications 8 (2007) 2007–2011 70 Ti/Mn = 0.67 Product selectivity (%): Ti/Mn = 0.18 Phenylacetaldehyde 60 Styrene oxide Benzaldehyde Conversion / % 15 18 16 14 Ti/Mn = 0.43 Ti/Mn = 0.18 50 18 8 40 16 8 20 Ti/Mn = 0.67 9 18 30 10 69 74 20 76 71 8 16 10 5 12 83 0 Blank 72 15 33 76 TiO2 68 52 TS-1 OMS-2 Ti-OMS-2 TiO 2 -OMS-2 TiO2 -OMS-2 Fig. 3. The conversion and product selectivity of oxidation styrene with tert-butyl hydroperoxide (TBHP) using TiO2, TiO2–OMS-2, Ti–OMS-2, OMS-2 and TS-1. All reactions were carried out at 70 C with styrene (5 mmol), 70% aqueous TBHP (10 mmol), acetonitrile (15 ml) and catalyst (50 mg) with vigorous stirring. (b) Absorbance / a.u. indicated in Fig. 3, the reaction catalyzed by all the catalysts produced the highest yield of benzaldehyde and their selectivities towards the formation of products are almost similar to each other. The high selectivity (52%) towards benzaldehyde over TS-1 was surprising since that phenylacetaldehyde and benzaldehyde were the major products from the oxidation of styrene catalyzed by the TS-1 zeolite [10], while a speculated product, styrene oxide, was not detected. Brønsted acid sites originating from framework titanium species catalyze the rearrangement of the intermediate, leading to the formation of phenylacetaldehyde [11]. This argument is in agreement with our oxidation of styrene over TS-1, since no Brønsted acid sites have been detected in our TS-1 which has been analyzed by the pyridine adsorption method [7]. IR spectrum of acidity study by pyridine adsorption after evacuation under a vacuum at 400 C and 150 C revealed in Fig. 4a shows that Lewis acid sites are formed in Ti–OMS-2 (0.43) as indicated by the appearance of peaks at 1447 cm1, 1489 cm1 and 1604 cm1. In contrast, no peaks are observed for the OMS-2 sample in Fig. 4b. The absence of peaks at 1540 cm1 confirms that there are no Brønsted acid sites in both samples. This is the possible reason why TiO2– OMS-2, Ti–OMS-2, TiO2 and OMS-2 catalysts are not selective towards phenylacetaldehyde. A high selectivity towards benzaldehyde may be due to OMS-2 and TiO2 promoting the carbon–carbon bond cleavage, thus resulting in the formation of benzaldehyde. As shown in Fig. 3, a considerable increase in the conversion of styrene over Ti–OMS-2, OMS-2, TiO2–OMS-2, TiO2 and TS-1 after 3 h of the reactions is clearly observed when Ti–OMS-2 (0.67) and TiO2–OMS-2 (imp) are used as catalysts. The increase in oxidation activity of Ti–OMS-2 1447 1604 1489 (a) 1640 1600 1560 1520 1480 1440 1400 -1 Wavenumber / cm Fig. 4. FTIR spectra of (a) Ti–OMS-2 (0.43) and (b) OMS-2 after evacuation under vacuum at 400 C for 4 h followed by pyridine adsorption at room temperature and evacuation at 150 C for 1 h. (0.67) and TiO2–OMS-2 (imp) can be explained on the basis of the presence of non-framework titanium species. The superior performance of Ti–OMS-2 (imp) and Ti– OMS-2 (0.67) strongly suggests the occurrence of a synergetic effect of non-framework Ti with OMS-2, since Ti– OMS-2 (mix), a mechanical mixture of TiO2 and OMS-2, gives a relatively lower conversion of styrene (see Fig. 3). Leaching is a particular problem of solid catalysts in liquid phase reactions. The catalysts were recycled three times. The activities of the recovered and dried Ti–OMS-2, OMS-2 and TiO2–OMS-2 showed an insignificant change (ca. 3–5%) that corresponds to experimental observations within experimental error in their catalytic activity in the H. Nur et al. / Catalysis Communications 8 (2007) 2007–2011 recycling test. This suggests a good regenerability of the catalysts in the oxidation of styrene with tertbutylhydroperoxide. The results of this investigation lead to the conclusion that the framework of titanium species in Ti–OMS-2 has no effect on the enhancement of catalytic activity, while the existence of Ti in the non-framework structure of OMS-2 induces a synergetic effect that enhances the catalytic activity in the oxidation of styrene with tert-butylhydroperoxide. Ti–OMS-2 containing non-framework Ti is highly active (ca. 70% conversion of styrene) and selective towards the oxidation of styrene to give benzaldehyde (68%) as the main product. Further detailed studies are, however, necessary to understand the interactions between TiO2 and OMS-2 support and the reaction mechanism. Acknowledgements This research was supported by the Ministry of Science Technology and Innovation Malaysia (MOSTI), under IRPA grant and The Academy of Sciences for the Developing 2011 World, under the TWAS Grants in Basic Sciences no. 04462 RG/CHE/AS. References [1] V.V. Krishnan, S.L. Suib, J. Catal. 184 (1999) 305. [2] Y.-C. Son, V.D. Makwana, A.R. Howell, S.L. Suib, Angew. Chem., Int. Ed. 40 (2002) 4280. [3] R. Ghosh, Y.-C. Son, S.L. Suib, J. Catal. 224 (2004) 288. [4] C.L. O’Young, Y.F. Shen., R.N. DeGuzman, S.L. Suib., United States Patent, 5,702,674 (1997). [5] R. Ghosh, X. Shen, J.C. Villegas, Y. Ding, K. Malinger, S.L. Suib, J. Phys. Chem. B 110 (2006) 7592. [6] R.N. DeGuzman, Y.F. Shen, S.L. Suib, B.R. Shaw, C.L. O’Young, Chem. Mater. 5 (1993) 1359. [7] H. Nur, D. Prasetyoko, Z. Ramli, S. Endud, Catal. Commun. 5 (2004) 725. [8] D. Prasetyoko, Z. Ramli, S. Endud, H. Nur, J. Mol. Catal. A 241 (2005) 118–125. [9] Database of Joint Committee for Powder Diffraction Studies (JCPDS) PDF# 29, 102. [10] K. Fujihara, S. Izumi, T. Ohno, M. Matsumura, J. Photochem. Photobiol. A 132 (2000) 99. [11] J. Zhuang, D. Maa, Z. Yan, X. Liu, X. Hana, X. Bao, Y. Zhang, X. Guo, X. Wang, Appl. Catal. A 258 (2004) 1.