INTERMOLECULAR PHOTOREACTIONS AND SELECTIVITY STUDIES IN CONFINED SPACE OF Y ZEOLITES YEOH KAR KHENG UNIVERSITI TEKNOLOGI MALAYSIA PSZ 19:16 (Pind. 1/97) UNIVERSITI TEKNOLOGI MALAYSIA BORANG PENGESAHAN STATUS TESIS♦ JUDUL : INTERMOLECULAR PHOTOREACTIONS AND SELECTIVITY STUDIES IN CONFINED SPACE OF Y ZEOLITES SESI PENGAJIAN: 2004/2005 Saya : YEOH KAR KHENG (HURUF BESAR) mengaku membenarkan tesis ( PSM / Sarjana / Doktor Falsafah )* ini disimpan di Perpustakaan Universiti Teknologi Malaysia dengan syarat-syarat kegunaan seperti berikut: 1. Tesis adalah hakmilik Universiti Teknologi Malaysia. 2. Perpustakaan Universiti Teknologi Malaysia dibenarkan membuat salinan untuk tujuan pengajian sahaja. 3. Perpustakaan dibenarkan membuat salinan tesis ini sebagai bahan pertukaran antara institusi pengajian tinggi. 4. **Sila tandakan ( √ ) √ SULIT (Mengandungi maklumat yang berdarjah keselamatan atau kepentingan Malaysia seperti yang termaktub dalam AKTA RAHSIA RASMI 1972) TERHAD (Mengandungi maklumat TERHAD yang telah ditentukan oleh organisasi/badan di mana penyelidikan dijalankan) TIDAK TERHAD Disahkan oleh ______________________________ (TANDATANGAN PENULIS) ________________________________ (TANDATANGAN PENYELIA) Alamat Tetap: 429, LORONG 9, TAMAN KAYA, 34000 TAIPING, PERAK. ASSOC. PROF. DR. ABDUL RAHIM YACOB Tarikh:___27 MAY 2005 Tarikh: Nama Penyelia 27 MAY 2005 CATATAN: * Potong yang tidak berkenaan. ** Jika tesis ini SULIT atau TERHAD, sila lampirkan surat daripada pihak berkuasa/organisasi berkenaan dengan menyatakan sekali sebab dan tempoh tesis ini perlu dikelaskan sebagai SULIT atau TERHAD. ♦ Tesis dimaksudkan sebagai tesis bagi Ijazah Doktor Falsafah dan Sarjana secara penyelidikan, atau disertasi bagi pengajian secara kerja kursus dan penyelidikan, atau Laporan Projek Sarjana Muda (PSM). “We hereby declare that we have read this thesis and in our opinion this thesis is sufficient in terms of scope and quality for the award of the degree of Master of Science (Chemistry).” Signature : ……………………………………………… Name of Supervisor I : Assoc. Prof. Dr. Abdul Rahim Yacob Date : 27 May 2005 Signature : ……………………………………………… Name of Supersivor II : Assoc. Prof. Dr. Farediah Ahmad Date 27 May 2005 : BAHAGIAN A ⎯ Pengesahan Kerjasama* Adalah disahkan bahawa projek penyelidikan tesis ini telah dilaksanakan melalui kerjasama antara _______________________ dengan ________________________ Disahkan oleh: Tandatangan : _________________________________ Tarikh: ______________ Nama : _________________________________ Jawatan : _________________________________ (Cop rasmi) * Jika penyediaan tesis/projek melibatkan kerjasama. BAHAGIAN B ⎯ Untuk Kegunaan Pejabat Sekolah Pengajian Siswazah Tesis ini telah diperiksa dan diakui oleh: Nama dan Alamat Pemeriksa Luar : Prof. Dr. Zakaria bin Mohd Amin School of Chemical Science, Universiti Sains Malaysia 11800, Minden, Pulau Pinang. Nama dan Alamat Pemeriksa Dalam : Assoc. Prof. Dr. Zainab Ramli Dept. of Chemistry, Faculty of Science, Universiti Teknologi Malaysia, 81310 Skudai, Johor. Nama Penyelia Lain (jika ada) : _____________________________________ _____________________________________ _____________________________________ _____________________________________ Disahkan oleh Penolong Pendaftar di SPS: Tandatangan : _________________________________ Tarikh: ______________ Nama : Ganesan A/L Andimuthu INTERMOLECULAR PHOTOREACTIONS AND SELECTIVITY STUDIES IN CONFINED SPACE OF CATION-EXCHANGED Y ZEOLITES YEOH KAR KHENG A thesis submitted in fulfilment of the requirements for the award of the degree of Master of Science (Chemistry) Faculty of Science Universiti Teknologi Malaysia MAY 2005 ii I declare that this thesis entitled “Intermolecular Photoreactions and Selectivity Studies in Confined Space of Cation-Exchanged Y Zeolites” is the result of my own research except as cited in the references. The thesis has not been accepted for any degree and is not concurrently submitted in candidature of any other degree. Signature : ____________________ Name : YEOH KAR KHENG Date : 27 May 2005 iii For my family, teachers and friends. iv ACKNOWLEDGEMENT This thesis could not be completed without the help of many. I am most grateful to my supervisors Assoc. Prof. Dr. Abdul Rahim Yacob and Assoc. Prof. Dr. Farediah Ahmad who always gave me guidance and support. I must also thank all the lectures in Institute Ibnu Sina including Prof. Dr. Halimaton Hamdan, Assoc. Prof. Dr. Zainab Ramli, Assoc. Prof. Dr. Salasiah Endud, Dr. Hadi and others who always provided me good suggestions and solutions whenever I faced with difficulties in this research. My grateful acknowledgement is also due to En. Kadir, En. Fuad, En. Khairul and other staffs in Department of Chemistry who provided me a lot of helps in this research. I am also indebted to my parents, friends and fellow researches. It was from their encouragements, motivations, and supports enable me to complete this work. Lastly, I must record my profound thanks to the National Science Fellowship (NSF) and IRPA vote 74505 for financially supports in this research. v ABSTRACT Photochemistry in organized assemblies has attracted considerable attention because of their potential use in controlling photophysical and photochemical behaviour of organic molecules in a confined space. Conversion of a starting material to product in a photoreaction involves selectivity by the reaction cavity to the specified product. For solid and rigid media like zeolite, the size of the reaction cavity plays an important role in products selectivity. The surface of NaY zeolite was first studied with paramagnetic probe using Electron Spin Resonance spectroscopy (ESR). Two favourable active sites were identified. The study of a confine space reaction was first studied in the photosensitization of triethylamine by acetophenone in NaY zeolite. ESR result showed that radical cation of amine dimer was formed inside zeolite resulted from the confinement effect of the zeolite Y supercage. Ultraviolet (UV) irradiation of acetophenone in toluene solution results in photochemical hydrogen abstraction and yielded a mixture of both symmetric (1,2-diphenylethane and 1,2-diphenylethyl alcohol) and asymmetric (1,2-diphenylpinacol) coupling products. These were identified and characterized by gas chromatography-mass spectrometry (GC-MS) and nuclear magnetic resonance (NMR). With the introduction of NaY zeolite, high yield of asymmetric product, 1,2-diphenylpinacol was observed. It further proved the confinement effect played by the zeolite produced a drastic change in product selectivity compared to homogenous reaction. Photodimerization of 2-cyclohexenone in various cation-exchanged Y zeolites were also studied in solid state and zeolite-solvent slurries. Both the reactions showed a great reversal of head-to-tail (HT) cyclohexenone dimer, to head-to-head (HH) cyclohexenone dimer with increasing pattern from LiY to CsY zeolite. The study of regioselectivity in the photocycloaddition of 2-cyclohexenone to vinyl acetate was also carried out in zeolite slurries, in which the result showed a drastically change of product yield compared to the homogeneous reaction. However, the cationexchanged zeolites failed to control the selectivity. This is explained by the passive cavity effect of zeolite. vi ABSTRAK Fotokimia di dalam media teraturapi telah banyak menarik perhatian kerana potensinya dalam mengawal sifat fotofizik dan fotokimia molekul organik dalam ruang terhad. Pengubahan bahan pemula kepada produk dalam tindak balas fotokimia melibatkan kepilihan kaviti tindak balas terhadap produk tertentu. Untuk pepejal tegar seperti zeolit, saiz kaviti tindak balasnya memainkan peranan dalam kepilihan produk. Permukaan zeolite NaY telah dikaji dengan prob paramagnet menggunakan spektroskopi Resonans Spin Elektron (RSE). Dua tapak aktif telah dikenalpasti. Tindak balas dalam ruang terhad pada mulanya telah dikaji dalam pemfotopekaan trietilamina oleh asetofenon dalam zeolit NaY. Keputusan RSE menunjukkan radikal kation dimer amina terbentuk dalam zeolit disebabkan oleh kesan ruang terhad supersangkar zeolit. Penyinaran ultra-lembayung (UL) ke atas asetofenon dalam pelarut toluena pula menyebabkan pengabstrakan hidrogen dan menghasilkan campuran kedua-dua hasil gandingan simetri (1,2-difeniletana dan 1,2difeniletil alkohol) dan tidak simetri (1,2-difenilpinakol). Pengenalpastian dan pencirian hasil ini seterusnya dilakukan menggunakan kromatografi gas-spektrometri jisim (KG-SJ) dan resonans magnet nukleus (RMN). Penggunaan zeolit NaY pula menghasilkan hasil utama tidak simetri, 1,2-difenilpinakol. Ini membuktikan bahawa ruang terhad pada zeolit telah mengubah kepilihan hasil tersebut berbanding dengan tindak balas homogen. Pemfotodimeran 2-sikloheksenon dalam pelbagai zeolit Y tertukar kation juga dikaji dalam fasa pepejal dan buburan zeolit-pelarut. Kedua-dua tindak balas menunjukkan keterbalikan daripada dimer sikloheksenon kepala-ekor kepada dimer sikloheksenon kepala-kepala dengan penambahan corak daripada zeolit LiY kepada CsY. Seterusnya, keregiopilihan dalam pemfototambahan 2sikloheksenon kepada vinil asetat telah dijalankan dalam buburan zeolit. Keputusan menunjukkan perubahan besar dalam kepilihan produk berbanding dengan tindak balas homogen. Kegagalan zeolit tertukar kation dalam mengawal kepilihan produk adalah disebabkan oleh kesan kaviti pasif zeolit. vii TABLE OF CONTENTS CHAPTER 1 2 TITLE PAGE INTRODUCTION 1 1.1 Objectives of the Research 2 1.2 Scope of the Studies 2 LITERATURE REVIEWS 3 2.1 Supramolecular Photochemistry Versus Zeolite 3 2.2 The Origins of Supramolecular Chemistry 5 2.3 Zeolite 6 2.4 2.5 2.3.1 Faujasite (FAU) Zeolite 8 2.3.2 Ion Exchange Behavior 9 2.3.3 Electrostatic Field 10 2.3.4 Adsorption 11 2.3.5 Diffusion 13 2.3.6 14 Photochemistry 15 2.4.1 Basic Laws of Photochemistry 15 2.4.2 Electronic Transitions 17 2.4.3 Pathways of Excited States 19 2.4.4 Frontier Orbital Approach in Photochemical Reactions 21 Photocycloaddition Reactions 2.5.1 2.6 Confinement Effect 22 Regiochemistry and Stereochemistry of Photocycloaddition in Enones 23 Electron Spin Resonance (ESR) Spectroscopy 27 2.6.1 28 The ESR Spectrometer viii 2.7 2.8 3 2.6.2 Basic Principle of ESR 29 2.6.3 Hyperfine Structure 30 X-Ray Diffraction (XRD) 32 2.7.1 Theory of XRD 33 Flame Emission Spectroscopy (FES) 35 2.8.1 Basic theory and Flame Photometer 35 2.8.2 Quantitative Analysis 36 EXPERIMENTAL 38 3.1 Instrumentations 38 3.2 Chemicals 39 3.3 UV Irradiation of H2 in NaY Zeolite 39 3.4 ESR Study of the Photosensitization of Triethylamine by 3.5 Acetophenone in NaY Zeolite 41 Preparation of Alkali Metal Cation-Exchanged Y Zeolites 42 3.5.1 Quantitative Analysis of the Cation-Exchanged Y Zeolites 3.6 Photochemical Hydrogen Abstraction by Acetophenone in Toluene Solution and NaY Zeolites Slurry 43 3.6.1 Homogeneous Reaction 43 3.6.2 Isolation of Photoproducts 44 3.6.2.1 Thin Layer Chromatography (TLC) 45 3.6.2.2 Gravity Column Chromatography (CC) 45 Photoreaction in NaY Zeolite Slurry 45 3.6.3 3.7 Photodimerizations of 2-Cyclohexenone 46 3.7.1 Homogeneous Reactions 46 3.7.2 Solid State Photoreactions in Cation-Exchanged Y Zeolites 3.7.3 47 Photoreactions in Cation-Exchanged Y Zeolite-Slurries 3.8 42 49 Photocycloaddition of 2-Cyclohexenone to Vinyl Acetate 50 3.8.1 Homogenous Photoreaction 50 3.8.1.1 Acid Test 51 3.8.2 Photoreactions in Cation-Exchanged ix Y Zeolite Slurries 4 RESULTS AND DISCUSSION 4.1 ESR Study of the UV Irradiation of H2 in NaY Zeolite 4.2 51 53 53 An ESR Investigation of Amine Dimers Radical Cation in the Photosensitization of Triethylamine by Acetophenone in NaY Zeolite Supercages 58 4.3 Alkali Metals Cation-Exchanged Y Zeolites 62 4.4 Photochemical Hydrogen Abstraction by Acetophenone in 4.5 Toluene Solution and NaY Zeolites Slurry 65 4.4.1 Homogenous Photoreaction 65 4.4.2 Photoreaction in NaY Zeolite Slurry 68 Regioselective Photodimerizations of 2-Cyclohexenone in Alkali Metal Cation-Exchanged Y Zeolites 4.5.1 Photodimerizations of 2-Cyclohexenone in Homogenous Solution 4.5.2 83 Photocycloaddition of 2-Cyclohexenone to Vinyl Acetate (VA) in Alkali Metal Cation-Exchagned Y Zeolite-Slurries 4.6.1 Homogenous Solution 4.6.2 Photocycloadditions in Alkali Metal Cation-Exchanged Y Zeolite Slurries 5 76 Photodimerizations of 2-Cyclohexenone in Alkali Metal Cation-Exchanged Zeolite Slurries 4.6 72 Solid State Photodimerizations of 2-Cyclohexenone in Alkali Metal Cation-Exchanged Y Zeolites 4.5.3 72 CONCLUSIONS 87 87 89 93 REFERENCES 95 APPENDIXES 1-13 113 x LIST OF TABLES TABLE NO. TITLE PAGE 2.1 Cation dependence of supercage free volume in FAU zeolites 14 3.1 GC-MS analysis of the supernatants in the photochemical hydrogen abstractions in NaY zeolite slurries 46 3.2 GC-MS analysis of the tetrahydrofuran extracts in the photochemical hydrogen abstractions in NaY zeolite slurries 3.3 GC peak ratios of the photoproducts in the photodimerizations of 2-cyclohexenone in homogenous reactions 3.4 46 47 GC peak ratios of the photoproducts in the solid state photodimerizations of 2-cyclohexenone carried on different cation-exchanged Y zeolites 3.5 49 GC peak ratios of the photoproducts obtained in the photodimerizations of 2-cyclohexenone carried in cation-exchanged Y zeolite-slurries 50 GC peak ratios of the photoproducts in the photocycloadditions of 2-cyclohexenone to vinyl Acetate in cation-exchanged Y zeolite-slurries 52 4.1 Ion-Exchanged levels of alkali metal cations-exchanged Y zeolites 65 4.2 Product ratios calculated by GC in the photochemical hydrogen abstraction by acetophenone in toluene solution 68 Product ratios in the tetrahydrofuran extract of the photolysed NaY zeolite 70 3.6 4.3 xi 4.4 4.5 4.6 4.7 4.8 Product ratios of the photodimerization of 2-cyclohexenone in n-hexane 74 Product ratios of the solid state photodimerizations of 2-cyclohexenone in alkali metal cation-exchanged Y zeolites with tetrahydrofuran extractions 77 Product ratios obtained by solid state photodimerizations of 2-cyclohexenone in alkali metal cation-exchanged Y zeolites with HCl treatment and ethyl acetate extractions 79 Product ratios of the photodimerizations of 2-cyclohexenone in alkali metal cation exchanged Y zeolite-hexane- slurries 86 Product ratios obtained in photocycloadditions of 2-cyclohexenone to vinyl acetate in different mediums. 92 xii LIST OF FIGURES FIGURE NO. TITLE PAGE 2.1 Oxygen is shared between two tetrahedra 8 2.2 External surface and supercage of FAU zeolite 9 2.3 Adsorption and desorption isotherm curves of N2 in zeolite NaY at 77 K 12 Pictorial representation of the diffusion of molecules in a zeolite particle 13 2.5 Orbital energy level description of absorption and emission 18 2.6 Jablonski Diagram 20 2.7 Frontier orbital interactions between a photochemically excited molecule and a ground state molecule of 1,3,5-hexatriene 22 2.8 [4 + 2] cycloaddition (a Diels-Alder reaction) 23 2.9 Alkene [2 + 2] photocycloaddition 23 2.10 Head-to head and head-to-tail regioisomers found in photocyloaddition of cyclohexenone to unsymmetrical alkene 23 2.11 Photocycloaddition of cyclohexenone to methoxyethylene 24 2.12 Photocycloaddition of cyclohexenone to electron-rich alkenes 25 2.13 Photocycloaddition of methyl substitution cyclohexenone to alkene 25 2.14 Stereochemical disposition around the cyclobutane ring in the cis-fused photoaddition products 25 2.15 Photocyclodimerization reaction of acenaphtylene 26 2.16 Regioselectivity on photocycloadditon reactions 2.4 xiii of substituted cyclohexenone with cycloalkenylesters 27 2.17 The schematic diagram of an ESR spectrometer 28 2.18 The absorption and first derivative of ESR spectra 29 2.19 Zeeman energy levels of an electron in an applied magnetic field 30 The interaction of an electron with a single nucleus I = ½ and the resulting ESR spectrum 31 2.21 Simplified X-ray diffractometer 33 2.22 Pictorial view of Bragg’s Law 34 2.23 Schematic diagram of a flame photometer 36 2.24 Plot of emission intensity versus concentration 37 3.1 Sample cell for activation and UV irradiation 40 3.2 Vacuum line used for sample activation and sample degassing 41 Experiment set up for UV irradiations in homogenous solutions and zeolite- solvent slurries 44 3.4 Experiment set up of solid state photoreactions 48 4.1 ESR spectrum of H2 in NaY before UV irradiation 54 4.2 ESR spectrum of UV irradiation (after 45 minutes) of H2 in NaY zeolite supercages 55 Stucture of the FAU zeolite with cation position type II and type III in the supercages 56 ESR spectrum of UV photolysis (after1 hour) of Acetophenone in NaY zeolite 57 2.20 3.3 4.3 4.4 4.5 (a) Peak 1 intensity and (b) Peak 2 intensity against UV irradiation time 4.6 ESR spectrum of UV photolysis (after1 hour) of triethylamine in NaY zeolite 4.7 4.8 ESR spectrum of UV photolysis (after1 hour) of acetophenone and triethlyamine in the NaY zeolite supercages X-ray diffractograms of the alkali metal cation-exchanged 59 60 60 xiv Y zeolites compared to parent NaY zeolite 63 4.9 Crystalinity versus cation-exchanged Y zeolites 64 4.10 Emission intensity versus concentration of Na analysis in flame emission photometry 64 GC chromatograms (a) before and (b) after the homogenous photoreaction of acetophenone in toluene solution 66 GC chromatograms of the supernatant and the resulting tetrahydrofuran extract 69 The difference of molecule distributions in homogenous solution and zeolite slurry (spectator approach) 70 GC chromatograms of the homogeneous photoreactions of 2-cyclohexenone compared to solid state photoreactions 73 4.15 Corey’s model 75 4.16 GC chromatograms (b) and (d) show the remained products which trapped in the zeolites after tetrahydrofuran extractions. 78 GC chromatograms of the solid state photodimerizations of 2-cyclohenone in alkali metal cation-exchangedY zeolites (a)-(e) 80 Ratio HT(16)/HH(17) obtained in this research compared to ratio obtained by Lem et al. 82 GC chromatograms of the photodimerizations of 2-cyclohexenone in alkali metal cation-exchanged Y zeolite-hexane slurries (a)-(e) 85 GC analysis on the reaction mixture in the photocycloaddition of vinyl acetate to 2-cyclohexenone in hexane 88 4.21 Photocycloaddition of vinyl acetate to 2-cyclohexenone 88 4.22 GC chromatograms of the photoproducts in photocycloadditons of 2-cyclohexenoen to vinyl acetate in alkali metal cation-exchanged Y zeolite-slurries 91 4.11 4.12 4.13 4.14 4.17 4.18 4.19 4.20 xv LIST OF SCHEMES SCHEME NO. 4.1 TITLE PAGE The proposed mechanism of amine photosensitization by acetophenone inside NaY zeolite supecages 61 The mechanism of photochemical hydrogen abstraction by acetophenone in toluene solution and zeolite NaY slurry 71 4.3 Photodimerization of 2-cyclohexenone (1) 74 4.4 Various intermediates which can lead to cyclohexenone dimers 76 4.5 Photocycloadditon of 2-cyclohexenone to ethoxyethene 89 4.2 xvi LIST OF SYMBOLS/ABBREVIATIONS A - Ampere Å - Meter-10 AcP - Acetophenone cm - Centimeter CH - 2-Cyclohexenone Cps - Count per second Eq. - Equation EtOAc - Ethyl acetate g - Gram HH - Head-to-head HT - Head-to-tail Hz - Hertz (Second-1) 1 - Proton Nuclear magnetic Resonance FAU - Faujasite zeolite K - Kelvin k - Kilo L - Litre Μ - Mol/Litre M+ - Molecular ion H NMR xvii MY - Alkali metals Y zeolite m - multiplet min - Minute mg - Milligram mL - Millimeter mT - Millitesla m/z - mass per charge N - Normality Rf - Retention factor Rt - Retention time s - singlet sec - Second TEA - Triethylamine THF - Tetrahydrofuran V - Volt VA - Vinyl acetate W - Watt xviii LIST OF APPENDICES APPENDIX TITLE PAGE 1. MS spectrum of 1,2-diphenylethane (DPE) (10) 113 2. MS spectrum of 2,3-diphenylpropan-2-ol (DPP) (11) 114 3. MS spectrum of 2,3-diphenylbutan-2,3-diols (DPB) (12) 115 4. 1 116 5. MS spectrum of CH dimer, HT (16) 117 6. MS spectrum of CH dimer, HH (17) 118 7. MS spectrum of CH dimer (18) or (19) (Peak 1 in Figure 4.14) 119 8. MS spectrum of CH dimer (18) or (19) (Peak 3 in Figure 4.14) 120 9. MS spectrum of cyclohexene-cyclobutene adduct (P 1) 121 10. MS spectrum of cyclohexene-cyclobutene adduct (P2) 122 11. MS spectrum of cyclohexene-cyclobutene adduct (P3) 123 12. MS spectrum of cyclohexene-cyclobutene adduct (P4) 124 13. MS spectrum of cyclohexene-cyclobutene adduct (P5) 125 H NMR spectrum of 2,3-diphenylbutan-2,3-diols (DPB) (12) CHAPTER 1 INTRODUCTION Zeolites have been an object of scientific research and a material beneficial to mankind for more than two centuries since its discovery in 1756. However, it was not until 10-15 years ago that zeolites attracted the keen interest of photochemists who wanted to use them in their research. Photochemists are most interested in controlling chemical reactions with the aid of supramolecular assemblies, aimed at constructing artificial photosynthetic systems, controlling chirality and inventing nanoscale advanced materials. Zeolites are found to be particularly useful for such purpose since they can host various organic molecules in their cavities and channels; such inclusions have often been shown to modify the photophysicals and photochemistry of a given species. Besides, photochemical reactions pursued in zeolites also provide product distributions considerably different from those in solution [1]. Zeolite nanospace could be considered as “hard” because of the frameworks of zeolite are rigid, and “active” because of the non-bonding interaction between the walls of the supercage and the included molecules. On top of these, the most desirable property of zeolite is that it is transparent to light in the near-UV and visible regions. Thus it eliminates the possibility of competitive absorption between the medium and the guest molecule presents [2]. The synthetic utility of intermolecular photodimerization of cyclic enones and the cycloadditon to unsymmetrical alkenes can be limited by the formation of the mixtures of the head-to-head (HH) and head-to-tail (HT) regioisomers [3, 4]. HT 2 regioisomers are always formed in much larger amount compared to HH isomer in solution reaction [5, 6]. In this research, cation-exchanged Y zeolites were applied to control the regioselectivity of photoproducts in photoreactions of 2-cyclohexenone. 1.1 Objectives of the Research The objective of this research is to evaluate the feasibility of using zeolite as reaction medium to carry out intermolecular organic photoreactions, i. e. photocycloaddition and photodimerization. This could be further divided to two: (i) To compare the products selectivity of between the conventional homomogenous photoreactions with solid state and/or slurry photoreactions in zeolite supercage (ii) To utilize the cation-exchanged property of zeolite to control the regioselectivity of desired photoproducts. Faujasite-Y zeolite was used as host because it possesses large supercages volume which enable us to study a variety of photochemical reactions. 1.2 Scope of the Studies At the first part of this research, locations of the paramagnetic probe in different adsorption sites of NaY zeolite were studied using Electron Spin Resonance Spectroscopy (ESR). Most of the research in the supramolecular photochemistry within zeolites deal with the intramolecular reaction. In order to study the different approaches used in the intermolecular photoreaction, we have studied the triplet sensitization technique and “spectator” method. The triplet sensitization technique had been applied in the dimerization of triethylamine (TEA) within Y zeolite, while 3 the “spectator” method was used in gaining selective asymmetric coupling products in the hydrogen abstraction of toluene by acetophenone (AcP). After gaining experiences from the first part, we turned to the next part, the utilization of the size constriction effect and cation-guest interactions of the cationexchanged zeolites to modify the selectivity of the regioisomers in photodimerization of 2-cyclohexenone (CH) and photocycloaddition of CH to vinyl acetate (VA). CHAPTER 2 LITERATURE REVIEWS 2.1 Supramolecular Photochemistry Versus Zeolite The last two centuries have witnessed the growth of organic photochemistry from relatively unknown to a more developed discipline. During this period, photochemists have discovered new reactions, established mechanism of photoreactions, laid out the ground rules for the behavior of molecules in excited state and surfaces, and found applications of photochemistry in everyday life. In spite of these achievements, photochemistry is yet to become a sought after tool in industrial synthetic processes. Nowadays, organic photochemistry seemed to have developed into three stages: (a) discovery of reaction; (b) mechanistic pursuit, and (c) gaining control on the outcome of a reaction [7]. Supramolecular or guest@host chemistry (symbol @ represents noncovalent binding of the guest and host) is the chemistry that is dominated by forces resulting from molecular non-bonded, non-covalent electrostatic forces (due to static and oscillating fixed charge interactions) and dispersion force (due to induced transient charge interactions) to control the selectivity of reactions of geminate radical pairs whose molecular chemistry involves random radical-radical reaction [8]. The guests serve as molecular probes of the host structure. The host controls the initial sitting of the guests, the sieving probabilities, the size shape selective diffusional dynamic, and the topology of diffusion pathways available to the adsorbed guest and to the reactive intermediates produces by the absorbed guest organic molecules [9]. 5 Conversion of starting material to product in photoreaction (and any chemical reaction) involves change in the shape of reaction cavity from reactant-like to product like. A number of organized assemblies such as micelles, vesicles, mono and bilayer, liquid crystal, cyclodextrins, silica clay and zeolite surfaces have been examined as media to control the excited state of organic molecules. Each of them is unique in their ability to modify in photoreaction. In the case of photochemical reactions occurring in liquid or liquid-like media like micelles, this change is not obvious since the surrounding media rearranges itself to accommodate this change. However, this becomes different in the case of rigid structures like crystal. Among these, zeolites are the most versatile host system to control the reactions of a large variety of molecules [2, 7]. Zeolites are made-up of silica and alumina. They are not that different from the glassware used in laboratory reaction, except that the size of the “glassware” now is at molecular dimensions [2]. One of the recent trends in photochemical research on zeolite cavities focuses on the way in which the restricted spaces influence the geminatoselectivity [7, 9] regioselectivity [10-12] and stereoselectivity [13, 14] of products and enhanced molecular interaction of guest molecules [15-17]. Recently, attentions have been given to the subjects of photochemical asymmetric synthesis [18-25] with the use of chiral auxiliaries, and the study of photosensitization within zeolites [26-29]. Carrying out the photoreactions in nanospace of zeolite cavity gives us a unique opportunity to understand the role played by factors like confinement and electrostatic interactions with the cation sites on the reaction pathway. 2.2 The Origin of Supramolecular Chemistry The starting point in the history of supramolecular effects on the chemistry may be traced back to 1934 [30], which J. Franck and Rabinowitch coined the term ‘cage effect’ to explain observations comparing the photochemistry of diatomic molecules (e.g. I2) in the gas phase to their photochemistry in the liquid phase [8]. A molecule in solution may happen to dissociate after light absorption, and the radicals or atoms formed in this way separate with a certain amount of kinetic 6 energy will be at once lost in collisions with the solvent. In addition to the ‘normal’ probability of recombination governed by the law of mass action, there will be an additional probability of primary recombination of two particles, which have been parts of the same molecule before dissociation. The recombination effect will probably show wavelength dependence, decreasing with the increasing energy of the absorbed quantum. A greater excess energy will permit the dissociation products to find their way through the surrounding ‘walls’ of the solvent and to put more molecular layers between them before coming to rest [8]. The brilliant insight and imagery of the importance of a radical pair in a ‘solvent cage’ may be considered as setting the stage for supramolecular chemistry, which concerned with how non-covalent, intermolecular interactions can influence the chemistry of ‘bimolecular’ systems. The solvent cage is a primitive but fundamental supramolecular ‘host’ that exerts an influence on the chemistry and reactivity of an incarcerated ‘guest’ molecule or pair of ‘guest’ molecule [8]. 2.3 Zeolite Zeolites are inorganic material with a three-dimensional structure resembling a honeycomb. It is made up of interconnecting channels and cages that extend three dimensionally throughout the structure. Organic guest molecules could be adsorbed and held inside the cavities because of non-bonded interactions and electrostatic forces inside these cavities [31-34]. The history of zeolite began from the discovery of the stilbite by a Swedish mineralogist, Axel Cronstedt, in 1756 [35]. He found that the natural mineral stilbite visibly lost water when heated. Accordingly, he named this class of mineral, zeolite from the classical Greek words “zeo” (to boil) and “lithos” (stone). Zeolites can be simply divided into two categories, natural and synthetic. Natural zeolites are usually found in basaltic areas and volcanic regions as well as in the sedimentary deposits in many part of the world [36]. The pioneering work in synthetic zeolite was carried out by Barrer [37, 38] and Milton [39]. There are approximately 40 naturally 7 occurring and over 100 synthetic forms of zeolites. Some of the examples of the natural zeolites are stilbite, analcime, chabazite, ferrierite, mesolite and clinoptilolite. Predominant types of synthetic zeolites are type A, type X, type Y and ZSM-5 [32, 40]. The natural zeolites have not gained the commercial importance of the synthetic zeolites due to limitation in availability, large variations in the mineral composition, crystal size, porosity, and pore diameter. Application areas of the natural zeolite include building materials, agriculture, water treatment, radioactive waste treatment, and pet litter and odor control. On the other hand the synthetic zeolites have large market volume in detergent builder, petroleum refining and petrochemical processing catalysts, and a variety of uses such as adsorbents/desiccants (molecular sieves). Zeolite A for example was developed specifically as an eco-friendly (environmentally preferable) detergent builder as an alternative to phosphate builder which can cause eutrophication. The current global value market for zeolite is estimated to be $2.15 billion per annum, the components of natural and synthetic zeolites being $450 million and $1.7 billion, respectively [40]. The primary building blocks of zeolite are [SiO4]4- and [AlO4]5- tetrahedra. These tetrahedra are linked through oxygen atoms to form channels and cages of discrete size with no two aluminum atoms sharing the same oxygen atom (Figure 2.1). As a result, the total framework charge is negative, and it must be balanced by cations, typically of an alkali or alkaline earth metal cations. These cation could be exchanged by conventional ion-exchange method. The position, size and the number of cations can significantly alter the properties of the zeolite. Zeolites can be broadly divided into two types based on the pore structure. Zeolites with interconnecting cages (e.g. Faujasite, Zeolite A) and zeolites made up of channels that might be or might not be interconnected (e.g. ZSM-5, Zeolite-Beta). In short, zeolite can be considered as a polymeric porous crystalline hydrated aluminosilicate based on an infinite three-dimensional structure which has a general formula of [31-33]: 8 Mx/n[(SiO4)y(AlO4)x].zH2O where M= exchangeable cations of valency n []= zeolite framework n = valency of cation x = the number of AlO4 tetrahedra y = the number of SiO4 tetrahedra z = the number of moles of zeolitic water Figure 2.1: Oxygen is shared between two tetrahedra [8]. 2.3.1 Faujasite (FAU) Zeolite The structure of faujasite (FAU) zeolites (Zeolite X and Y) is cubic and built from sodalite cage (0.66 nm in diameter, with an entry aperture of 0.21 nm) connected via the double 6-membered ring. The entry aperture of the sodalite cage is too small for oxygen molecule to enter; however, water molecules are known to go into it. Zeolite X and Y have different framework of Si/Al ratios: 1.0 < Si/Al <1.5 for zeolite X and 1.5< Si/Al < 3 for zeolite Y. No faujasite with Si/Al ratio less than 1.0 has been prepared to date due to the unstable framework structure [1]. These zeolites form three-dimensional network of nearly spherical supercages of about 1.3 nm in diameter connected tetrahedrally to four other supercages through 0.74 nm windows. The charge-compensating cations are mobile and distributed among several types of sites. The supercage concentration in zeolite Y with Na+ ions as charge-compensating cation (NaY) is estimated to be approximately 6 x 10-4 9 mol/g on the basis of the crystal structure. Figure 2.2 shows the external surface and the internal supercage of a FAU zeolite [1,8]. Each supercage in FAU zeolites could accommodate up to five molecules of benzene, two molecules of naphthalene, or two molecules of pyrene [41]. The relatively large dimensions supercages can be used as a host for various photoreactions, such as photodimerization [10, 42], photocycloaddition [43] and recently, it was applied in the photochemical asymmetric synthesis [18-25], and the study of photosensitization reactions [26-29]. FAU External Surface Sodalite cage Window opening 0.74 nm FAU Framework Diameter 1.3 nm Internal supercage Figure 2.2: External surface (left) and supercage (right) of FAU zeolite [8]. 2.3.2 Ion Exchange Behavior The cations and water molecules are distributed within the zeolite intracrystalline pore system. Unlike water, the cations are not free to leave the crystals unless they are replaced by their electrochemical equivalent of other cations [31]. The ion exchange behavior (selectivity and degree of exchange) mainly depends on the size and charges of the hydrated cation, the temperature, the concentration, and to some degree the anion species. Cation exchange may produce 10 considerable change in thermal stability, adsorption behavior, and catalytic activity. The ion exchange process is presented by the following equation: ZABZB (z) + ZBAZA (s) ZABZB (s) + ZBAZA (z) where ZA and ZB are the ionic charge of cations A and B, and (z) and (s) represent the zeolite and solution. The ion exchange property of zeolites has been used in commercial applications such as detergent builders, radioisotope separation, and removal of ammonium ions from wastewater streams. Besides these, zeolites are used to replace phosphates as water-softening agents [40]. 2.3.3 Electrostatic Field Zeolite can be regarded as a solvent that dissolves or disperses molecules into pores and channels similar to solvent cage. Despite the similarity of the zeolite pores to solvent shells, zeolite pores are rigid and distinctly shaped in contrast to the soft and featureless solvent shells [1, 44]. The solvation-like interaction can be expected for zeolite host-guest molecule pairs. The negatively charged framework and the mobile cations combine to produce an electrostatic field akin to solvent polarity inside the cavities where the molecules reside. The electrostatic field strength in zeolites has been reported to be extremely high [45, 46]. It is considered due to the fact that the cations exposed at the center of the supercage being only partially shielded. The strength of the electric field is dependent on both the cation size of the charge-compensating cation and Si/Al ratio of the zeolite framework. For example, the smallest alkali metal cation, Li+ ion induces a stronger field in its proximity than the largest Cs+ ion. Also, cations in zeolite Y (which has higher Si/Al ratio) exhibit higher fields than those in zeolite X. 11 The effects of the fields have been explored with a number of fluroscence probes incorporated into zeolite. [47, 48]. 2.3.4 Adsorption Adsorption is one of the fundamental issues in zeolite science. Figure 2.3 represents an adsorption isotherm of N2 in NaY zeolite at 77 K [1]. As shown by the figure, both the adsorption and desorption curves are superimposed indicating the adsorption and desorption processes are completely reversible. The adsorption process occurs in two stages with increasing equilibrium pressure. Initially, the steep rise part indicates that N2 is being absorbed into the internal supercages, while the flat part indicates the process takes place very slow with the increase in pressure. This accounts for N2 molecules covering the external surface of the zeolite at a monoleyer level. In this case, more than 90% of the N2 molecules were adsorbed in the supercages. It can be simply explain by the dominating volume of supercages in the zeolite, and the adsorption takes place first in the cages mainly because of a big gain in entropy [1]. Size exclusion is a basic guideline for adsorption of organic molecules in zeolites. Results showed the maximum quantity of anthracene that could be adsorbed in NaX zeolite from hexane solution at 296 K is 5.3 x 10-4 mol/g, nearly one molecule per supercage. However, an abrupt decrease in adsorption (5 x 10-6 mol/g) was observed for 9,10-dimethylanthracene that apparently has a size larger than the aperture of the supercage [49]. The adsorbed organic molecules are mainly residing inside the zeolites if they can meet the requirement of size limitation imposed by the entry apertures to the cages [1]. A variety of techniques have demonstrated the cation-guest interaction plays an important role during the adsorption of organic molecules in cation-exchanged Y zeolites [50-53]. Cation-π interaction (also known as cation-quadrupolar interaction) [54-56] has been recognized as the main force of binding between aromatic guest molecules and the zeolite supercage. However, when a benzene ring contains a polar 12 group such as nitro (dinitrobenzene), the primary interaction is between the cation and the oxygen of the nitro group (not the π cloud of the phenyl ring) [57]. Such a type of dipolar interaction between the cation and the guest predominate even in nonaromatic such as hyroflurocarbons within NaY [58-60]. It has been shown the emission spectra for larger aromatic species such as naphthalene [61], and pyrene [62] in NaY or NaX zeolite are loading dependent. The contribution of excimer emission increases at the expense of monomer as the loading level increases. This indicates a heterogeneous distribution within the zeolites, because the excimer emission is considered to arise from more than one molecule occupying the same cage. An average occupancy of unity in zeolites does not mean that most of the supercage contain one guest species on the average but rather that most cages are empty, with some being multiply-occupied [63]. Figure 2.3: Adsorption and desorption isotherm curves of N2 in zeolite NaY at 77 K: (o) adsorption curve; (●) desorption curve [1]. 13 2.3.5 Diffusion The diffuse of molecules within the intracrystalline cage network of zeolite is important in influencing the chemical reactions of the guest species and for the design and application of shape-selective catalysts [64,65]. The molecules introduced externally into zeolites are not fixed at particular sites, but rather migrate by executing a hopping motion from one adsorption site to another within the same cage and occasionally to other cages through one of the connecting windows. Typically, intercage hopping assumes a relatively high activation barrier and is a slow process, allowing the guest molecules to stroll around the adsorption sites within a given cage before they jump out. Figure 2.4 shows the pictorial representation of the diffusion of molecules in a zeolite particle. The diffusional motion is classified as two types according to the nature of the activation energy; (1) intercage jump and (2) intracage jump [66, 67]. Figure 2.4: Pictorial representation of the diffusion of molecules in a zeolite particle. The arrows in this picture represent the motion of molecules [1]. The cage-to-cage diffusivity of the guest species is highly possible dependent on the molecular size because of the constraints on the diffusive motion imposed by the windows and walls of the zeolite. It has been pointed out that the intercage hopping dynamics of an organic molecule is largely affected by the adsorption interaction with the host zeolite [68]. 14 2.3.6 Confinement Effect The reactions which take place within a zeolite actually occur in the supercage, which can be considered as “reaction cavity”. This reaction cavity is of molecular scale dimensions hence would influence the reactivity of the substrate and the course of the reaction. In this confined space, the mobility and conformational flexibility will be restricted. The “free volume” indicates the space in which the reactants transform themselves to products. The volume available for an organic molecule within a supercage depends on the number and the nature of the cation [69, 70]. Table 2.1 shows the available volume for a guest decreases as the cation size increases from Li to Cs [70]. Table 2.1: Cation dependence of supercage free volume in FAU zeolites [70]. Cation (M+) Radius of the cation (Å) Y- zeolite X-zeolite Vacant space within the supercage (Å3) Li 0.60 834 873 Na 0.95 827 852 K 1.33 807 800 Rb 1.48 796 770 Cs 1.69 781 732 The spatial confinement in the cavities of zeolite is expected to provide molecules with geometric restrictions. The cavities of zeolite may prevent or restrict the approach of the guest molecules, in particular, reactive intermediate which may due to unusual photophysic and/or photochemistry of the guest species. This idea was tried with a molecule whose size is similar to the dimension of the entry aperture, and thus the host/guest complex was expected to fix in very tightly [1]. Recently, the idea of confinement effect has led to the understanding of electronic confinement. Marquez et al. [71-73] proposed the influence of the cavity dimensions on the electronic structure of some guest molecules incorporated within the zeolites framework can be related to the quantum confinement concept. This explanation was similar to “electronic confinement effect” in which the electron 15 density of the guest is constrained and localized within the zeolite cavity as a result of strong range repulsion with the electrons of the zeolite walls [74]. 2.4 Photochemistry Photochemistry is the study of chemical changes by visible or ultraviolet light. It plays an important role in everyday processes that occur in nature such as photosynthesis in plants and photodissociation of ozone in the atmosphere that prevents harmful ultraviolet radiation of sun reaching the earth’s surface. The process of vision itself involves the photochemical isomerization of the protein, rhodopsin in the retina of eyes. [75, 2] Quantum Chemistry predicts that molecules exist in a variety of electronic states that differ in both electronic energies and wave functions. Classical chemistry involves reactions of molecules in their electronic ground state (R0) (lowest electronic energy state) [76]. The study of reactions that occur through high electronic energy (R*) (electronic excited states) of organic molecules is called organic photochemistry. Reactions that are thermodynamically unfavorable when the reactants are in the ground states may occur from an excited state in photoreaction. In addition, the other advantage of photochemical reactions is specific bonds could be activated depending on the frequency of the radiation used for excitation [2, 77]. 2.4.1 Basic Laws on Photochemistry The First Law of Photochemistry or Grottus-Draper Law states, “Only radiation absorbed in a system can produce a chemical change” [83]. The amount of light absorbed is related to the concentration of the absorbing molecule in the path of the irradiation; this relation is known as Beer-Lambert Law [77, 78] and is represented by Eq.1: A = log I0/I = εcl (Eq. 1) 16 A: Beer-Lambert law absorbance I0: incident radiant flux I: transmitted radiant flux ε: molar absorption coefficient (l mol-1 cm-1) c: concentration of absorbing molecules (mol l-1) l: absorbing path length (cm) The Beer-Lambert law is valid except when very high intensities of radiation are employed, such as laser light [79]. “A molecule which undergoes a photochemical change does so by an absorption of a single quantum of light energy”. This statement is referred to the Stark-Einstein Law of Photochemical Equivalence. In other word, it means that the number of activated molecules equal to the number of quanta of radiation absorbed (Eq. 2). Exceptions of the law have been observed in two-photon absorption processes [77, 80]. 1hv + 1R0 = 1R* (Eq. 2) h : Planck’s constant v : frequency of absorbed radiation R0: reactant in ground state R*: reactant in excited state The energy required for electronic excitation is excitation energy (Eexc) [77], which is defined as Eq. 3: Eexc = E* - E0 = hv = hc/λ E : electronic excitation energy E*: energy of the excited state E0: energy of the ground state h : Planck’s constant v : frequency of absorbed radiation (sec-1) (Eq. 3) 17 c : velocity of light in vacuo (3 x 1010 cm sec-1) λ: wavelength of absorbed light (cm) Generally, electronic excitations that produce photochemical reactions are induced by the absorption of ultraviolet (UV) or visible (VIS) electromagnet by a molecule. The radiation energy corresponds to the excitation energies of organic molecules is 140 – 30 kcal mol-1 (λ = 200 –700 nm) [77]. The efficiency of a photochemical process is variable and is expressed in term of quantum yield Φ which is showed in Eq. 4. Φ = Number of molecules reacting or formed Number of quanta absorbed 2.4.2 (Eq. 4) Electronic Transitions Absorption of ultraviolet or visible light by an organic molecule results in the excitation of an electron from an initial occupied, low energy orbital to a high energy, previously unoccupied orbital [81]. Organic molecules that have the capability of doing so contain chromophores, also defined as functional groups which absorb near ultraviolet or visible radiation. Examples include C=O, Ph, NO2 and -N=N-. Chromophores absorb in the far ultraviolet (λ < 200 nm) are C=C and C≡C [82, 83]. The absorption process can be simplified as in Figure 2.5 [81]. 18 hv + S1 S0 hv + (Spin allowed absorption) electron jump S1 S0 (Spin forbidden absorption) electron jump and spin flip S1 S0 electron jump S1 S0 + hv (fluorescence) + hv (phosphorescence) electron jump and spin flip Figure 2.5: Orbital energy level description of absorption and emission. The arrows intersected by the levels represent electrons. The direction of the arrow represents the orientation of the electron spin [81]. Two excited electronic states derive from the electronic orbital configuration produced by light absorption. In one state, the electron spins are paired (antiparallel) and the other state the electron spins are unpaired (parallel). The state with paired spins has no resultant spin magnetic moment, but the state with unpaired spins possesses a net spin magnetic moment. A state with paired spins interacts remain a single state in the presence of magnetic field, and is termed as a singlet state. A state with unpaired spins interacts with the magnetic field and splits into three quantized states, and is termed a triplet state. These are three states, which are most crucial to an understanding of organic photoreactions [81]: 1. S0 = ground, singlet state 2. S1 = lowest energy excited, singlet state 3. T1 = lowest triplet state The triplet state is slightly lower in energy than the corresponding singlet state because two paired electrons have a greater electronic repulsion Hund’s rule [84]. 19 2.4.3 Pathways of Excited States Excited states are short-lived and lose excess energy by returning to the ground state as rapidly as possible. There are several pathways it can occur; there are [85]: (i) Radiative processes, (ii) Radiationless processes, (iii) Energy transfer, (iv) Chemical reactions. The commonly encountered photophysical radiative and radiationless processes are best shown by Jablonski Diagram (Figure 2.6) [81]. The Jablonski Diagram can be simply summarized as: (A) Radiative processes (shown by arrow in Figure 2.6): (a) “Allowed” or singlet-singlet absorption (S0 + hv → S1) (b) “Forbidden” or singlet-triplet absorption (S0 + hv → T1) (c) “Allowed” or singlet-singlet emission, called fluorescence (S1→ S0 + hv) (d) “Forbidden” or singlet-triplet emission, called phosphorescence (T1→ S0 + hv) (B) Radiationless processes (shown by dotted arrow in Figure 2.6): (e) “Allowed” transitions between states of the same spin, called internal conversion (e.g., S1→ S0 + heat) (f) “Forbidden” transitions between excited states of different spin, called intersystem crossing (e.g., S1→ T1 + heat) (g) “Forbidden” transitions between triplet states and the ground state – also called intersystem crossing (e.g., T1→ S0+ heat) 20 Radiative process Radiationless process S3 T3 S2 S-S absorption T2 T-T absorption (f) S1 T1 (e) (a) (c) (b) S0 (d) (g) S0 Figure 2.6: Jablonski Diagram [81]. The intersystem crossing occurs within 10-8 to 10-10 sec and is slower than internal conversion (10-9 to 10-14 sec). The triplet state T1 is therefore longer-lived than singlet state, S1. Chemical reactions are much more common for excited triplet species because of the longer lifetimes of these states [81]. 21 2.4.4 Frontier Orbital Approach in Photochemical Reactions Knowledge of all the molecular orbitals in a compound is necessary to fully understand its chemistry. However, a great deal can be learned by looking at only two of the orbitals, that are the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO). These two molecular orbitals are known as the frontier orbitals [86]. The first step of bimolecular photoreactions is the excitation of a component with the chromophore which most efficiently absorbs light. If a conjugated system present in the component, promotion of a frontier electron from HOMO to LUMO occurs on the excitation. The excited state produce is ππ* whilst nπ* when the excitation results from the non-bonding orbital, such as lone pair electron of carbonyl group in ketone. The second step of the reaction is the interaction between the excited molecule and the ground state molecule. The two effective frontier orbital interactions are [87]: (1) Interaction between the singly occupied π* of the excited molecule with LUMO of the ground state molecule (shown at top part of Figure 2.7). (2) Interaction of the singly occupied n or π orbital of the excited molecule with HOMO of the ground state molecule (shown at bottom part of Figure 2.7). Therefore, the important frontier orbitals in a photochemical reaction are HOMO/ ‘HOMO’ and LUMO/ ‘LUMO’ of the ground state and the excited state molecules respectively. There are strong bonding since the orbitals are closest in energy [86, 87]. 22 Energy LUMO ‘LUMO’ ‘HOMO’ Excited molecule HOMO Molecular orbital Ground state molecule Figure 2.7: Frontier orbital interactions between a photochemically excited molecule and a ground state molecule of 1,3,5-hexatriene [87]. 2.5 Photocycloaddition Reactions Two different π-bond-containing molecules react to form a cyclic compound in a cycloaddition reaction. Each of the reactants loses a π bond, and the resulting cyclic product has two new σ bonds. These are classified according to the number of π electrons that interact in the reaction. The best-known example of cycloaddition reaction is Diels-Alder [4 + 2] addition and the [2+2] photocycloaddition. The examples of Diels-Alder [4 + 2] addition and [2 + 2] photocycloaddition are shown in Figure 2.8 and Figure 2.9 respectively [86]. 23 O + O O O O O Figure 2.8: [4 + 2] cycloaddition (a Diels-Alder reaction) [86]. hv + Figure 2.9: Alkene [2 + 2] photocycloaddition [86]. 2.5.1 Regiochemistry and Stereochemistry of Photocycloaddition of Enones Head-to-head (HH) and head-to-tail (HT) regioisomers are two possible orientations commonly found in the addition of an enone to an unsymmetrical alkene. This example is given in Figure 2.10 [3-5]. R hv + + R O R O head-to-head (HH) O head-to-tail (HT) Figure 2.10: Head-to-head and head-to-tail regioisomers found in photocyloaddition of cyclohexenone to unsymmetrical alkene [3]. 24 The orientation of the addition is control by the geometry of the intermediate π-complex formed between the enone and the alkene in the photocycloaddition process. This then proceeds via a 1,4-diradical to the cyclobutane photoproduct [5]. The exciplex (π-complex) formed results from dipolar interaction between the excited enone (acceptor) and the ground state alkene (donor). Calculations of charge distribution in the n→ π*excited state of planar α, β-unsaturated ketones show the Cβ is quite negative relative to Cα [87-89]. Thus addition of excited 2-cyclohexenone (CH) (1) to ground state methoxyethylene (2) leads largely to the HT photoadduct (3) (Figure 2.11) [5]. δ− δ+ O (1) + OCH3 δ+ δ− (2) OCH3 OCH3 hv O O (3) Figure 2.11: Photocycloaddition of cyclohexenone (1) to methoxyethylene (2) [5]. The steric effect was also found to influence the regiochemistry of the addition. It can be done by increasing the bulk of the enone β-substituent or alkene substituent [90-92]. The presence of the nitro group on the β-position of cyclohexenone (1) was found to increase the proportion of the HH regioisomers in the adduct mixture [93]. Attempts also have been made to control the regiochemistry by placing removable directing groups on the enone. For example, a trimethylsilyl group at the α-position of the 2-cyclopentenone has been found to increase the proportion of the HT isomer in the photocycloaddition reaction with 2-acetoxyproprene; the silyl group can be than removed from the adducts by treatment with fluoride ion [94]. The stereospecificity of [2 + 2] photocycloaddition reaction is discussed base on the relative stereochemistry of the newly created chiral centers of the cyclobutene ring. Addition of acyclic alkenes to cyclic enones usually will form the cis-fused photoproducts whereas with the electron-rich alkenes, the addition will give transfused adducts as the major products (Figure 2.12) [91, 92]. 25 R R + O H R hv R R = CH3, OCH3 H O Figure 2.12: Photocycloaddition of CH to electron-rich alkenes [91]. It was found that alkyl substitution at Cβ in cyclohexenone (1) [91, 92] also tends to increase the proportion of cis-fused products (Figure 2.13). R R + O R hv R R = CH3, OCH3 O H Figure 2.13: Photocycloaddition of methyl substitution cyclohexenone to alkene [91]. [2 +2] Photocycloaddition of the enones with cyclic alkenes give either cis or trans-fused adducts, the preferred arrangement usually is determined by the ring size. A further point of interest is the stereochemical disposition around the cyclobutane ring in the cis-fused photoaddition products. The possible stereoisomers are cis-anticis and the cis-syn-cis isomers (Figure 2.14) [81]. O O cis-anti-cis (a) cis-syn-cis (b) Figure 2.14: Stereochemical disposition around the cyclobutane ring in the cis-fused photocycloaddition products [81]. 26 One of the examples of this isomeric form is found in the photocyclodimerization reaction of acenaphtylene (4) (Figure 2.15) [95]. hv syn (4) hv triplet sens anti Figure 2.15: Photocyclodimerization reaction of acenaphtylene (4) [95]. Recently, Omar et al. [96] reported the experimental results of the triplet [2 + 2] photocycloaddition reactions of substituted CH with cycloalkenylesters, cyclobutenylester (5), cyclopentenylester (6), and cyclohexenylester (7) gave remarkable change in the regioselectivity of the products (Figure 2.16). The HT/HH product ratio increases with the increment of the cycle-size. The changes in the HT/HH ratio with the enlargement of the alkene ring size may be due to the increment of the repulsion energy between the enone carbonyl and esters in the alkenylesters, and large changes in the deformation energy of the reactants. 27 R h t R H (5) hv O H R > 95% R (6) t R R H hv h O O (4) H R R R O H H 40% 60% R R = CO2Me hv R R R H (7) O H R O < 5% head-to-head (HH) H H > 95% head-to-tail (HT) Figure 2.16: Regioselectivity on photocycloadditon reactions of substituted CH with cycloalkenylesters (5), (6), and (7) [96]. 2.6 Electron Spin Resonance (ESR) Spectroscopy Electron Spin Resonance (ESR) or Electron Paramagnetic Resonance (EPR) spectroscopy is a physical method of observing resonance absorption of microwave power by unpaired electron spins in magnetic field. It has developed into a most direct, sensitive and powerful non-destructive method for the characterization and measurement of species with unpaired electron [97]. ESR is a specific technique for systems with net electron spin angular momentum [97]. These systems include: (i) free radicals in the solid, liquid, or gaseous states; (ii) some point defects (localized crystal imperfections) in solids; (iii) 28 biradicals; (iv) systems in the triplet state; (v) systems with three or more electrons and (iv) most transition-metal ions and rare-earth ions. 2.6.1 The ESR Spectrometer The schematic diagram of the more common continuous-wave (CW) ESR spectrometer is shown in Figure 2.17. It consists of a microwave source (klystron oscillator), cavity in which the sample is inserted in a quartz container, a microwave detector, and electromagnet with a field that can varied in the region of 0.3 T. The ESR spectrum is obtained by monitoring the microwave absorption as the magnetic field is varied. To minimize the noise from the diode in steady state measurements, a magnetic field modulation scheme with phase sensitive detection is usually employed. As a result, the detected signal appears as a first derivative of the absorption intensity [98]. Figure 2.18 shows the absorption and first derivative of ESR spectra. Figure 2.17: The schematic diagram of an ESR spectrometer [98]. 29 Figure 2.18: The absorption and first derivative of ESR spectra [98]. 2.6.2 Basic Principle of ESR In the presence of magnetic field, H, an interaction between the magnetic moment of an unpaired electron and the applied field will occur and these energy which yields different spin states known as “Zeeman Energy”. The Zeeman energy is given by: Ez = gβMsH (Eq.4) where Ez is the Zeeman energy, Ms represent the magnetic quantum number, β is the electronic Bohr magneton with a value 9.2733 x 10-28 J/Gauss and g is the spectroscopic splitting factor which has a value of 2.0023 for a free electron. The possible values of Ms are Ms = + ½ and Ms = – ½ for an electron. Hence, the two possible values of the energy levels (Zeeman levels) are Ez = + ½ gβH (α state) and Ez = - ½ gβH (β state) which is represented in Figure 2.19. 30 ∆E = gβH Ez = - ½ gβH H=0 H Figure 2.19: Zeeman energy levels of an electron in an applied magnetic field [97]. The direction of the spin is changed by the absorption of microwaves when the energy different (∆E = gβH) is equal to the quantum energy of an electromagnetic wave, hv, where h is the Planck’s constant and v the frequency of an electromagnetic radiation. This absorption of the electromagnetic wave (microwave) by the unpaired electron is called “Electron Spin Resonance”. The resonance condition is represented by ∆E = gβHr = hv (Eq. 5) where Hr is the resonance magnetic field. This is the fundamental equation of ESR spectroscopy and to obtain an absorption by paramagnetic species, we either fix the magnetic field and vary the frequency, or fix the frequency and vary the magnetic field however, the later is chosen with the frequency being in the microwave magnetic region (λ = 3 cm and v ≈ 9 GHz) and the magnetic field being centered around 3000 gauss (300 mT) [97-100]. 2.6.3 Hyperfine Structure If the only effect observed in ESR were the interaction of an unpaired electron with an external field, then the spectrum would consist of only one line. However, one of the most important features of ESR spectra are their hyperfine 31 structure, the splitting of the individual electron resonance lines into components with respect to nucleus with a spin. In spectroscopy, the term “hyperfine structure” means the structure of the spectrum that can be traced to interactions of the electrons with other nuclei as a result of the latter’s point electric charge [98]. The “hyperfine coupling” is the term used to describe the magnetic coupling that occur between the spin of the unpaired electron and those of the nearest magnetic nuclei in the molecule [101]. Figure 2.20 shows the interaction of an unpaired electron with a single nucleus with nuclear quantum number, I = ½ (upper part in figure) and the resulting ESR spectrum (bottom part in figure); “A” represents the hyperfine coupling constant (Hz) while gN and βN represent the spectroscopic splitting factor and Bohr magneton of the nucleus. hA measures the interaction between the electron and nucleus. Dashed line in the figure correspond to the allowed transition according to selection rule ∆Ms = ± 1 and ∆MI = 0. The total number of lines is given by, N = 2nI + 1 where “n” is the number of equivalent nuclei which interact with the electron. MI = - 1/2 - ¼ hA Ms = + 1/2 H=0 geβH E2 - gNβNH MI = + 1/2 + ¼ hA MI = - 1/2 - ¼ hA E1 E3 - gNβNH Ms = - 1/2 MI = + 1/2 + ¼ hA Degenerate Levels Electronic Zeeman Splitting Nuclear Zeeman Splitting E4 Hyperfine Interaction A Figure 2.20: The interaction of an electron with a single nucleus I = ½ (upper) and the resulting ESR spectrum (bottom) [101]. 32 2.7 X-Ray Powder Diffraction (XRD) X-Rays were discovered by Wilhelm Röntgen in 1895. They are electromagnetic radiation with wavelengths of the order of 10-10 m and are typically generated by bombarding a metal with high-energy electrons. While the phenomena of diffraction is the interference caused by an object in the path of waves. It occurs when the dimensions of the diffracting objects are comparable to the wavelength of the radiation. The pattern of varying intensity that results from the phenomena is called the diffraction pattern [98]. XRD is an instrumental technique that is used to identified minerals, as well as other crystalline materials. XRD provides the researcher with a fast and reliable tool for routine mineral identification. Other information obtained can include the degree of cystallinity, the structural state, possible deviations of the minerals from their ideal compositions, and degree of hydration for minerals that contain water in their structure. In X-ray powder diffractometry, X rays are generated within a sealed tube that is under vacuum. A current is applied that heats a filament within the tube. A high voltage typically 15-60 kilovolts is applied within the tube. This high voltage accelerates the electrons which then hit a target, commonly made of copper and Xrays are produced. These X-rays are collimated and directed onto the sample, which has been ground to a fine powder. A detector detects the X-ray signal, which then processed either by a microprocessor or electronically, converting the signal to a count rate. Changing the angle between the X-ray source, the sample, and the detector at a controlled rate between preset limits is an X-ray scan. Figure 2.21 shows a simplified X-ray diffractometer which contain the X-ray source (X-ray tube), X-ray detector, and the sample during and X-ray scan. In this configuration, the X-ray tube and the detector both move through the angle (θ), and the sample remains stationary [102]. 33 Figure 2.21: Simplified X-ray diffractometer [102]. 2.7.1 Theory of XRD When X-ray radiation passes through matter, it interacts with the electrons in the atoms resulting in scattering of the radiation. If the atoms are organized in planes (i.e. the matter is crystalline) and the distances between the atoms are of the same magnitude as the wavelength of the X-rays, constructive and destructive interference will occur. This result in diffraction where X-rays are emitted at characteristic angles based on the spaces between the atoms organized in crystalline structures called planes. There are many different sets of planes in crystal. Each set of planes has a specific interplanar distance that will give rise to a characteristic angle of diffracted X-rays [98]. The relationship between wavelength (λ), atomic spacing (d) and angle was solved as the Bragg’s Law in Eq. 6. Figure 2.22 shows the pictorial representative of the equation [102]. 34 n λ = 2 d sin θ (Eq. 6) Where, n = the order of the diffracted beam λ= wavelength of the incident X-ray beam d= the distance between adjacent planes of atoms (d spacing) θ= angle of the incidence X ray beam. X-ray Plane of atoms Figure 2.22: Pictorial view of Bragg’s Law [102]. Since λ is known and θ can be measured, then d-spacing can be calculated. The characteristic set of d-spacings generated in a typical X-ray scan provides a unique “fingerprint” of the material. When properly interpreted by comparison with the standard reference patterns and measurements, this “fingerprint” allows for identification of the material [102]. 35 2.8 Flame Emission Spectroscopy (FES) 2.8.1 Basic theory and Flame Photometer The early use of a flame as an excitation source for analytical emission dates back to Herschel [103] and Talbolt [104], who identified alkali metals by flame excitation. Flame emission spectroscopy (FES) is so named because of the use of a flame to provide the energy of excitation to atoms introduced into the flame [105]. Flame spectrophotometry has been widely used in clinical application such as analysis of cations in biological fluids and tissues, and also diagnosis and treatment of many diseases. Besides, it also been used in the analysis of soils, plant materials, plant nutrients, cement and glass [106]. The high stability of the flame source was recognized as the key to the construction of simple instruments for the determination of easily excited elements such as alkali metals, sodium and potassium [105]. This relies on the principle that an alkali metal salt drawn into a non-luminous flame will ionize, absorb energy from the flame and then emit light of a characteristic wavelength as the excited atoms decay to the unexcited ground state. The intensity of emission is proportional to the concentration of the element in the solution. A photocell detects the emitted light and converts it to a voltage, which can be recorded. Since Na+ and K+ emit light of different wavelengths, by using appropriate colored filters the emission due to Na+ and K+ (and hence their concentrations) can be specifically measured in the same sample. Besides alkali metals, FES also can be used to analyze other elements such as calcium, magnesium, iron, nickel and platinum [105,106]. Figure 2.23 shows the instrumental set up of a flame photometer. The sequence processes occur in a flame photometer can be simply summarized as below [105]: 1. Sample solution sprayed or aspirated as fine mist into flame. Conversion of sample solution into an aerosol by atomizer. 2. Heat of the flame vaporizes sample constituents. 36 3. By heat of the flame and action of the reducing gas (fuel), molecules and ions of the sample species are decomposed and reduced to give atoms. eg. Na+ + e- --> Na 4. Heat of the flame causes excitation of some atoms into higher electronic states. 5. Excited atoms revert to ground state by emission of light energy, hν, of characteristic wavelength; measured by detector. Figure 2.23: Schematic diagram of a flame photometer [105]. 2.8.2 Quantitative Analysis Plot of emission intensity versus concentration of ionic species in the solution being measured is linear over wide range but with deviation at both low and high concentrations [105]. Figure 2.24 shows the plot of emission intensity versus concentration. 37 Figure 2.24: Plot of emission intensity versus concentration [105]. The plot in Figure 2.24 shows the following occurrence: 1. At very low concentration, emission falls below expected due to inoization, some atoms converted back to ions (eg. K --> K+ + e-). 2. Linear region. 3. Negative deviation at high concentration due to self-absorption. Photons emitted by excited atoms partly absorbed by ground state atoms in flame. CHAPTER 3 EXPERIMENTAL 3.1 Instrumentations ESR spectra in this work were recorded on a JEOL JES-FA 100 spectrometer, operating at X-band frequencies and 100 kHz, interfaced to a computer with JEOL system software. The ESR sample tube was made of quartz with diameter of 2.0 mm. The peak intensity and g value were automatically calculated by the data analysis progam. In-situ photolysis was carried out in the cavity of ESR spectrometer by using JEOL Ultraviolet radiation, ES-USH 500 Hg lamp, 500 W. The gas chromatography (GC) was performed by Hewlett-Packard chromatometer Model 6890. Tetrahyrofuran (THF), or dichrolometane (CH2Cl2), or ethyl acetate (EtOAc) was used as solvent with injection amount of 2 µL. Helium gas was used as the mobile phase while column Ultra 1 (100% polymethylsiloksane) with 0.11 µm thickness, 25.0 m length and internal diameter of 0.20 mm was used as the stationary phase. The column was operated from the temperature of of 50oC (maintained for 5 minutes) up to 250oC with the rate 8oC/min. The MS spectra were taken using coupled GC-MS Agilent Technologies spectrometer Model G 1540 N (GC) and G 2579 A (MS) with identical operation condition as in the GC analysis. Nucleus Magnetic Resonance spectrum for proton (1H NMR 400 MHz) was recorded with a Bruker spectrometer with deuterated chloroform (CDCl3) as solvent. Plate Merck pre-coated silica gel F254 with 2.0 mm thickness was used in the Thin Layer Chromatography (TLC). Silica gel Kieselgel 39 Merck with particle size 70-230 mesh was used as packing material for Gravity Column Chromatography (CC). X-ray diffractograms were obtained using D 500 Siemens Kristalloflex automated powder X-ray diffractometer with CuKα1 as the radiation source with λ = 1.548 Å at 40 kV and 30 mA. All zeolite samples were measured in the range of 2θ of 2 to 60 degree at room temperature with step intervals of 0.02 degree and scan speed of 4 deg. min-1. The reflection position, d value and peak intensity were calculated automatically by the data analysis program. Elemental analysis of Na in zeolites were done by using Jenway Flame Photometer (Model PFP 7). 3.2 Chemicals NaY zeolite (SiO2/Al2O3 = 5.1, unit cell size = 24.65 Ǻ, surface area = 900 2 m /g) was purchased from Zeolyst International. CH2Cl2 was purchased from MERCK and used after purification. Triethylamine (TEA), acetophenone (AcP), hexane, LiNO3 and KNO3 were obtained from MERCK. 2-Cyclohexenone (CH), vinyl acetate (VA), 1,2-diphenyletane (DPE), RbNO3 and CsNO3 were purchased from Fluka. Purified THF was obtained from Ajax Chemicals. HF (49 %) was obtained from Clean Room® Electronic Chemical. Ethyl acetate (EtOAc) and HCl (37 %) were obtained from Ashland and dehydrated MgSO4 was purchase from GCE Loboratory Chemicals. 3.3 UV Irradiation of H2 in NaY Zeolite A sample of NaY zeolite (50 mg) was activated at 300oC for 3 hours in a specially designed pyrex cell attached with ESR sample tube (Figure 3.1) under vacuum (10-4 Torr). Figure 3.2 shows the vacuum line used for all sample activations and sample degas in this research. In this experiment, specially designed pyrex cell replaced the sample activation bulk and the purified hydrogen was 40 introduced through the joint 2. Purified hydrogen was used straight from the tank and added to the activated sample at room temperature. Typical pressures of ~ 100 Torr were used. The pyrex cell was then taken off from the vacuum line and the activated zeolite was transferred to the ESR quartz tube by tilting and gently tapping the cell. This tube was then inserted into the ESR cavity and UV irradiated. The ESR spectra were recorded every 3 minutes until 15 spectra were obtained at 298 K. The ESR spectra showed 2 singlet peaks at 321.56 and 322.59 mT which correspond to g value of 2.0073 and 2.0008 . Ball joint ESR tube Valve B 14 joint Pyrex activation bulb Figure 3.1: Sample cell for activation and UV irradiation. 41 Teflon valve Mercury monometer Figure 3.2: Vacuum line used for sample activation and sample degassing. 3.4 ESR Study of the Photosensitization of Triethylamine by Acetophenone in NaY Zeolite NaY zeolite (200 mg) was activated under vacuum at 400oC for 3 hours. The zeolite was then added to a solution of acetophenone (AcP) (50 mg) in CH2Cl2 (2.5 mL) inside the glove box and stirred overnight. After filtration, the zeolite was then added to a solution of triethylamine (TEA) (50 mg) in dichloromethane (2.5 mL) and stirred overnight. The sample was then filtered and dried in air. The dried sample of NaY zeolite (40 mg) which contained AcP and TEA was degassed under vacuum (10-4 Torr) with a specially designed pyrex cell (Figure 3.1). The cell was then taken off from the vacuum line and transferred to an ESR quartz tube by tilting and gently tapping the cell. The sample tube was then inserted into the ESR cavity and UV irradiated. The ESR spectra of zeolite-AcP, zeolite-TEA, and zeolite-AcP-TEA were recorded after 1 hour of photolysis at room temperature. The ESR spectrum of zeolite-AcP showed a singlet peak at 322.83 mT while no ESR peak was observed for the zeolite-TEA sample. Zeolite-AcP-TEA sample 42 showed multiplet peaks at 319.21, 321.60, 322.13, 323.32, and 325.02 mT (A = 1.8 mT, g = 2.0072) and a singlet peak at 322.128 mT (g = 2.0023) in its ESR spectrum. 3.5 Preparation of Alkali Metal Cation-Exchanged Y Zeolites (MY) Nitrate solutions (0.5 M) of different cations (Li, K, Rb, and Cs) were prepared by dissolving the nitrates (LiNO3, KNO3, RbNO3, and CsNO3) with deionized water into volumetric flasks (200 mL). The nitrate solution (~65 –70 mL) was then added to NaY zeolite (5 g) in a Teflon bottle (200 mL) and put into the oil bath (90oC) with continued stirring for one hour. Then the zeolite was separated from the nitrate solution with centrifuge, and washed with deionized water. The exchange process was repeated three times. After filtration, the exchanged zeolites were washed thoroughly with distilled water and dried at 120oC for overnight. The dried samples were then weight and characterized using XRD. All the samples were saturated over concentrated NH4NO3 solution prior to XRD measurement in order to ensure complete hydration. The sample was ground to fine powder using pestle and mortar before mounting it on the sample holder. The XRD profile of the zeolites was compared with the simulated XRD pattern of standard. The crystallinity of the cation-exchanged zeolites were calculated using the intensity of the 3 reflection peaks, namely {533}, {642} and {840}. 3.5.1 Quantitative Analysis of the Cation-Exchanged Y Zeolites The samples of M+Y zeolite (M = Na, Li, K, Rb, and Cs) (~ 1 g) were dried at 120oC for 24 hours. The dried samples (0.1 g) were dissolved using HF 49% (5 mL) in a plastic beaker (5 mL). The diluted solution were then added with deionized water into a plastic volumetric flask (100 mL). Solution of HF was prepared as blank. The cation solutions were then sent for Na analysis using Flame Photometer. 43 The exchanged level (%) of cations was calculated depending on the replacement of Na cation in various MY zeolites. 3.6 Photochemical Hydrogen Abstraction by Acetophenone (AcP) in Toluene Solution and NaY Zeolites Slurry 3.6.1 Homogeneous Reaction Toluene solution of AcP (0.1 M, 5 mL) was added into a pyrex tube. The sample was flushed with purified argon gas for 1 hour prior to UV-irradiation in positive pressure condition. Figure 3.3 shows the experimental set up for the irradiation of sample in homogeneous solution and zeolite-solvent slurry. The sample was then UV-irradiated for 5 hours with continuous stirring in an inert condition. The conversion rate was determined using GC with m-xylene as an external standard. The sample was then concentrated under pressure using rotary evaporator and to give transparent liquid (40 mg).The GC chromatogram of the concentrated sample showed 5 peaks with retention time (Rt ) values of 18.62 , 19.60, 20.85 , 23.19, and 23.29 minutes. Peak 1 (Rt 18.62 minutes, 31.0 %); MS: m/z 182 [M+, C14H14], 165, 152, 91, 77, 65, 51, 41. Peak 2 (Rt 19.60 minutes, 2.0 %); MS: m/z 194, 179, 165, 116, 103, 91, 77, 65, 51. Peak 3 (Rt 20.85 minutes, 37.0 %); MS: m/z 194 [M+, C15H16O - H2O], 179 [194- CH3], 165, 152, 139,121, 105, 91, 77, 65, 51, 43. Peak 4 (Rt 23.19, 13.0 %); MS: m/z 210 [M+, C16H18O2 – H2O – CH3], 195 [210 – CH3], 181, 165, 121, 105, 91, 77, 65, 51, 43. 44 Peak 5 (Rt 23.29 minutes, 17.0 %); MS: m/z 210 [M+, C16H18O2 – H2O – CH3], 195 [210 – CH3], 181, 165, 121, 105, 91, 77, 65, 51, 43. Silicon oil Positive pressure Dry gas Dried silica gel Pryrex tube UV source A L U M I N I U M Solvent S H Zeolite E E T Stirrer hot plate Con. H2SO4 Argon Figure 3.3: Experiment set up for UV irradiations in homogeneous solutions and zeolite-solvent slurries. 3.6.2 Isolation of Photoproducts The reaction mixture was purified to further confirm the chemical structures of the photoproducts. 45 3.6.2.1 Thin Layer Chromatography (TLC) The sample was analyzed with TLC using CH2Cl2 (100%) as eluent. The TLC showed 3 major components with Rf 0.76, 0.50, and 0.31. 3.6.2.2 Gravity Column Chromatography (CC) The mixture (40 mg) was purified using CC. the gravity column (internal diameter 1.5 cm, height 10.0 cm) was packed with silica gel (7 g) in CH2Cl2 as eluent and 200 fractions were collected. Every fraction was analyzed by TLC. Fractions with the same TLC profile were combined and concentrated. However, only a pure compound was able to isolate. The combined fractions 120-160 after evaporation gave PH (13.8 mg, 34.5 %) with Rf 0.30 in CH2Cl2 (100 %). 1H NMR δ (CDCl3): 1.53 (3H, s, -CH3), 1.61 (3H, s, -CH3), 2.30 (1H, s, OH), 2.60 (1H, s, OH), 7.20-7.29 (10H, m, Aryl-H). 3.6.3 Photoreaction in NaY Zeolite Slurry Stock solution of acetophenone (AcP) in toluene (4 mg/mL) was prepared in a volumetric flask (25 mL). NaY zeolite (300 mg) was activated under vacuum at 300oC for 3 hours. Solution of AcP (3 mL = 12 mg AcP) was then added into the activated zeolite in a pyrex tube. After 3 hours stirring, the zeolite was washed twice with toluene (10 mL) and the combined solution was concentrated. The concentrated solution was analysed by GC to detect the absence of AcP. The sample was then degassed by three freeze-pump-thaw cycles. The resulting sample was irradiated for 5 hours under continuous stirring. The irradiated sample was centrifuged to separate the solvent and the zeolite. The zeolite obtained was then extracted with tetrahydrofuran (THF) for overnight. The concentrated solvent layer (supernatant) and the resulting extract was then analysed using GC. 46 The GC-MS results of the supernatant and the resulting THF extract were summarized in Table 3.1 and Table 3.2 respectively. The product ratios were calculated using peak area of product over peak area of total products. Table 3.1: GC-MS analysis of the supernatants in the photochemical hydrogen abstractions in NaY zeolite slurries. . Peak Retention time % of Area Rt (minutes) Molecular Molecular formula weight 1 18.08 12.78 - 182 2 18.62 76.95 C14H14 182 3 19.60 3.56 - 194 4 20.85 6.53 C15H16O 194 Table 3.2: GC-MS analysis of the tetrahydrofuran extracts in the photochemical hydrogen abstractions in NaY zeolite slurries. . Peak Retention time % of Area Rt (minutes) Molecular Molecular formula weight 1 18.62 13.00 C14H14 182 2 19.60 11.00 - 194 3 20.85 72.00 C15H16O 194 4 23.29 4.00 C16H18O2 242 3.7 Photodimerizations of 2-Cyclohexenone 3.7.1 Homogeneous Reactions Solution of 2-cyclohexenone (CH) (1) in hexane (10 mg/mL) was prepared in a volumetric flask (50 mL). The solution (5 mL) was transferred into a pyrex tube. The sample was flushed with purified argon gas for 30 minutes prior to UVirradiation in back pressure condition. The sample was UV-irradiated for 5 hours 47 with continuous stirring. Another 5 mL solution of CH (1) in hexane was UV irradiated for 5 hours in an open-air condition as the control experiment. The reaction mixture was then concentrated under reduced pressure using rotary evaporator and analysed with GC and GC-MS. The conversion rates for both the experiments were determined using GC with AcP as an external standard. The GC analysis of the photoproducts showed the presence of four peaks, Peak 1-4, each with Rt values of 23.40, 23.60, 23.89, and 24.12 minutes. All these four peaks gave the same molecular ion peaks with similar ion fragmentation patterns in the MS analysis; MS: m/z 192 [M+, C12H16O2], 175, 164, 149, 136, 121, 108, 96, 79, 68, 55, 41. Table 3.3 shows the obtained GC peak ratio of the photoproducts in this experiment. Table 3.3: GC peak ratios of the photoproducts in the photodimerizations of 2cyclohexenone in homogeneous reactions. Condition Peak 1 Peak 2 Peak 3 Peak 4 Inert 0.12 0.61 0.10 0.16 Open air 0.11 0.65 0.07 0.17 3.7.2 Solid State Photoreactions in Cation-Exchanged Y Zeolites The MY zeolite (M = Na, Li, K, Rb, and Cs) (300 mg) was activated under vacuum at 300oC for 3 hours. Hexane solution of CH (1.5 mL = 15 mg CH) prepared in Section 3.5 was added into volumetric flask (5 mL). An additional hexane was added until the solution reached a volume of 5 mL. The activated MY zeolite and the diluted solution of CH were then added into centrifuge tubes (50 mL) and stirred for 3 hours. The zeolite was washed twice with hexane (10 mL) and the combined washings were concentrated. The concentrated solution was analysed with GC to detect the absent of CH. The MY zeolite containing CH (CH-MY) were dried under vacuum (10-4 Torr) for 2 hours in a pyrex tube. The magnetically stirred dry powder sample was irradiated for 5 hours in the pyrex tube (magnetically stirred) in vacuum 48 condition. Figure 3.4 shows the experiment set up of solid state photoreaction. The irradiated zeolite was treated with HCl (1 N) and extracted with EtOAc. The resulting extract was then analysed with GC and GC-MS. The GC chromatograms showed 7 peaks with Rt 23.40, 23.60, 23.89, 24.12, 24.54, 25.66 and 25.87 minutes. Peak 1-6 gave the same molecular ion peaks with similar fragmentation patterns in the MS analysis; MS: m/z 192 [M+, C12H16O2], 175, 164, 149, 136, 121, 108, 96, 79, 68, 55, 41. With the similar fragmentation patterns with Peak 1-6, Peak 7 gave molecular ion peak at m/z 207 with molecular formula C12O12H16. Table 3.4 summarizes the GC peak ratios of the photoproducts obtained in photoreactions carried out in different MY zeolites. To vacuum pump Teflon valve Rubber stopper A L U M I N I U M UV source Zeolite Stirrer hot plate S H E E T Figure 3.4: Experiment set up for the solid state photoreactions. 49 Table 3.4: GC peak ratios of the photoproducts in the solid state photodimerizations of 2-cyclohexenone carried out in different cation-exchanged Y zeolites. MY Peak 1 Peak 2 Peak 3 Peak 4 Peak 5 Peak 6 Peak 7 LiY 0.18 0.10 0.06 0.50 0.05 0.10 0.02 NaY 0.17 0.15 0.09 0.31 0.10 0.15 0.04 KY 0.05 0.11 0.01 0.71 0.04 0.07 0.01 RbY 0.02 0.05 - 0.81 0.02 0.06 0.03 CsY 0.03 0.05 0.02 0.80 0.02 0.06 0.04 zeolite 3.7.3 Photoreactions in Cation-Exchanged Y Zeolite-Slurries The MY zeolite (M = Na, Li, K, Rb, and Cs) (300 mg) was activated under vacuum at 300oC for 3 hours. Hexane solution of CH (1.5 mL = 15 mg CH) prepared in Section 3.5 was added into a volumetric flask (5 mL). An additional hexane was added until the solution reached a volume of 5 mL. The activated MY zeolite and the diluted solution of CH were then transferred into centrifuge tubes (50 mL) and stirred for 3 hours. The zeolite was washed twice with hexane (10 mL) and the combined washings were concentrated. The concentrated solution was analysed with GC to detect the absent of CH. Hexane (5 mL) was then added to the washed sample in a pyrex tube. After purging with argon for 30 minutes, the magnetically stirred translucent MY-hexane slurry was irradiated for 2 hours. After the irradiation, the hexane layer was separated and concentrated for GC analysis. The irradiated zeolite was then dissolved with HCl (1N) followed by EtOAc extraction. The extract was analyzed using GC and GC-MS. The GC chromatograms showed 8 peaks with Rt 23.40, 23.60, 23.89, 24.12, 24.54, 25.66, 25.87 and 26.93 minutes. Peak 1-6 gave the same molecular ion peaks with similar fragmentation patterns in the MS analysis; MS: m/z 192 [M+, C12H16O2], 175, 164, 149, 136, 121, 108, 96, 79, 68, 55, 41. Peak 7 gave molecular ion peak at m/z 207 with molecular formula of C12O12H16. and Peak 8; gave the molecular ion 50 peak at m/z 277 with other fragment ions at m/z 222, 204, 160, 149, 135, 121, 104, 93, 76, 65, 57, 50. Table 3.5 summarizes the GC peak ratios obtained in these photoreactions. Table 3.5: GC peak ratios of the reaction mixture obtained in the photodimerizations of 2-cyclohexenone carried in cation-exchanged Y zeolite-slurries. MY Peak 1 Peak 2 Peak 3 Peak 4 Peak 5 Peak 6 Peak 7 Peak 8 zeolite LiY 0.14 0.05 0.06 0.32 0.14 0.10 0.08 0.12 NaY 0.28 0.06 0.07 0.16 0.19 0.16 0.06 0.02 KY 0.03 0.06 0.02 0.74 0.04 0.01 0.06 0.04 RbY 0.05 0.06 0.03 0.71 0.03 0.07 0.04 0.01 CsY 0.05 0.05 0.03 0.69 0.04 0.07 0.04 0.02 3.8 3.8.1 Photocycloaddition of 2-Cyclohexenone to Vinyl Acetate Homogeneous Photoreaction A hexane solution containing CH (1 g, 0.01 mol) and vinyl acetate (VA) (12.8 g, 0.15 mol) was prepared in a volumetric flask (50 mL). The solution (10 mL) was added into a pyrex tube and flushed with purified argon gas for 30 minutes prior to 5 hours of irradiations in the inert condition. The sample was stirred magnetically during the photolysis. The irradiated sample was then concentrated under pressure using rotary evaporator. The resulted solution was then analyzed using GC and GCMS. The GC chromatograms of the concentrated sample showed the presence of 5 peaks with Rt values of 19.58 (30.0 %), 19.71 (18.0 %), 19.83 (8.0 %), 19.94 (15.0 %), and 20.10 (29.0 %) minutes. These five peaks gave the same molecular ion peaks and similar fragmentation patterns; MS: m/z 158, 139 [M+, C10H14O3 –COCH3], 122 [M+- HOCOCH3], 111, 97, 84, 79, 55, 43. 51 3.8.1.1 Acid Test A small portion of the mixture (2 mL) was taken for the acid test. Evaporation of the solvent gave a mixture of photoproducts which were redissolved in HCl (1N) (5 mL) and hexane (1 mL), stirred for 15 minutes. The organic layer was separated, dried over anhydrous MgSO4 and analyzed by GC. The GC chromatogram showed no different in the peak profile or peak ratio as before the acid test. 3.8.2 Photoreactions in Cation-Exchanged Y Zeolite-Slurries The MY zeolite (500 mg) were activated under vacuum at 300oC for 3 hours. Hexane solution (0.5 mL) containing CH (10 mg/mL) and VA (128 mg/mL) were added into volumetric flask (10 mL). An additional hexane was added until the solutions reached a volume of 10 mL. The activated MY zeolite and the diluted solution were then transferred into a pyrex tube and stirred for 3 hours. The zeolite was washed twice with hexane (10 mL) to get rid of excess VA in solution. The washings were combined, concentrated and analyzed with GC. After purging with argon for 30 minutes, the magnetically stirred sample (CHVA-MY-hexane slurry) was irradiated for 5 hours under inert gas condition. After the irradiation, the hexane layer was separated and concentrated for GC analysis. The irradiated zeolite was then dissolved with concentrated HCl and isolated using EtOAc. The extract was then analyzed using GC. The GC chromatograms showed 5 peaks with Rt 19.58, 19.71, 19.83, 19.94, and 20.10 minutes. These 5 peaks gave the same molecular ion peaks and similar fragmentation patterns; MS: m/z 158, 139 [M+, C10H14O3 –COCH3], 122 [M+HOCOCH3], 111, 97, 84, 79, 55, 43. Table 3.6 summarizes the individual peak ratios obtained in these photoreactions. 52 Table 3.6: GC peak ratios of the photoproducts in the photocycloadditions of 2cyclohexenone to vinyl acetate in cation-exchanged Y zeolite-slurries. MY zeolite Peak 1 Peak 2 Peak 3 Peak 4 Peak 5 LiY NaY KY RbY CsY 0.21 0.25 0.21 0.34 0.21 0.05 0.06 0.05 0.05 0.06 0.28 0.24 0.29 0.21 0.25 0.08 0.09 0.08 0.05 0.06 0.38 0.36 0.37 0.35 0.42 CHAPTER 4 RESULTS AND DISCUSSION 4.1 ESR Study of the UV Irradiation of H2 in NaY Zeolite Paramagnetic probes may localize in sites characterized by different environmental mobility and polarity at the zeolite surface. With electron spin resonance (ESR) technique the precise structural and dynamical information about the probe and their environments can be studied by means of an accurate analysis of the spectral line shape [107]. Under conditions where the exchange rate among different sites is slow, the ESR signals at each site will contribute to superimposed adsorptions of overall spectra [108]. Due to their ability to interact with the surface sites, both paramagnetic metal ions [109, 110] and nitroxide radicals [108, 111-114] can be used as good spin probes in zeolite X and Y. The viscosity and the rotational motion of the adsorbed radicals depend on the dehydration degree of the zeolite, the nature of the support surfaces and the characteristic of the spin probe molecules [111]. In this research, locations of paramagnetic probe (H radical) in different adsorption sites of NaY zeolite was studied using Electron Spin Resonance spectroscopy in this section. Non-irradiated sample (Figure 4.1) of H2 in NaY zeolite gives only one singlet peak with the g value of 2.0001 is observed. This peak may due to the free electron from the defect sites on the zeolite framework. The other six peaks with the 54 A value 8.5 mT are distributed by the Manganese marker (I of Mn = 5/2) which was inserted as standard during recording of the spectrum. g = 2.0001 A mT Figure 4.1: ESR spectrum of H2 in NaY before UV irradiation. The resulting ESR spectrum for UV irradiation of H2 in NaY zeolite supercages after 45 minutes is shown in Figure 4.2. It gives 2 singlet peaks, with Peak 1 and Peak 2 corresponding to g values of 2.0008 and 2.0073. During UV irradiation, the H2 molecule is homolytically cleaved and the resulting H• radicals, being highly unstable on the zeolite surface, are easily ionised to produce H+ and a free electron. This process is given by the following equations: H2 2H• hv (Eq. 7) 2 H• 2H+ + 2e- (Eq.8) The free electron on proton is not seen as isotropic doublet lines (which indicates the hyperfine interaction between the free electron and a proton with I = ½) in the spectrum. This is further supported that the H radicals are ionised according to equation (8). 55 g = 2.0073, Peak 2 Peak 1, g= 2.0008 mT Figure 4.2: ESR spectrum of UV irradiation (after 45 minutes) of H2 in NaY zeolite supercages. Broussard et al. [115] have proved that both hydrated and dehydrated zeolites are known to present a locally homogeneous distribution of cations and surface groups (both O- and OH-), and it was proved that the FAU cavities bear different adsorption sites [116]. From literatures [2, 117], the charge compensating cations are known to be occupied in three main positions in FAU zeolites (zeolite X and zeolite Y). Type I cations with 16 per unit cell are located on the hexagonal prism faces between the sodalite cages. Type II, with 32 per unit cell is located in the open hexagonal faces. While the type III cations, with 38 per unit cell in zeolite X and only 8 per unit cell in the case of zeolite Y, is located on the walls of the large cavity. Cations in the zeolite exhibit a high mobility and are not rigidly located at their cations position, thus cations also can be found in sites I’ and site II’ in the sodalite units which are slightly displaced from the ideal positions. These extraframework cation sites are indicated in Figure 4.3. Only cations of site II and III in supercages are readily accessible to organic molecules. It is because the site I cations enjoy an octahedral coordination of six oxygens are well shielded from the guest molecules. The locations of the cations also vary depending on the presence of water and other molecules [2, 118, 119]. For the case of NaY, diffraction measurements have 56 determined that mainly two sites are occupied by Na+ cations in the dehydrated zeolite Y: site I and site II [120]. Site III Site I Site II Figure 4.3: Stucture of the FAU zeolite with cation position site II and site III in the supercages, site I’ and site II’ in sodalite units, and site I in the centers of hexagonal prisms [118]. From the 23Na synchronized double-rotation NMR spectra the cations on site I are more shielded than site II [117]. Therefore, we can assign Peak 1 represented the free electrons adsorb at sites II and Peak 2 represented the free electrons stayed at sites I. Since, we do not observed any extra peaks of isotropic lines, the interaction of unpaired electron and Na+ cation (I = 3/2) can be ignored. This is further proved by Gutjahr et al. [114] who claimed the conventional continuous wave (CW) ESR is not sufficient to resolve the hyperfine interaction of the unpaired electron with alkali metal ions in Y zeolite. Figure 4.4 shows Peak 1 and Peak 2 (refer to Figure 4.2) intensity plot against UV irradiation time. From the graph, it shows that the electrons diffused from site II to site I when the UV irradiation time was increased after 19 minutes. However, the intensity of Peak 1 (1200-1350) is higher than Peak 2 (0-300). This indicates that the electrons preferred to stay at sites II than sites I. The electric field is stronger at sites II than at sites I and the cations at sites II are poorly shielded. 57 Therefore, the radicals should preferentially interact at site II where the free electrons Peak 1 Intensity experience a rather strong spin polarization. 1340 1320 1300 1280 1260 1240 1220 1200 0 10 20 30 40 50 UV Irradiation Time (min) Peak 2 Intensity (a) 300 250 200 150 100 50 0 0 10 20 30 40 50 UV Irradiation Time (min) (b) Figure 4.4: (a) Peak 1 intensity and (b) Peak 2 intensity against UV irradiation time. 58 4.2 An ESR Investigation of Amine Dimers Radical Cation in the Photosensitization of Triethylamine by Acetophenone in NaY Zeolite Supercages One of the intermolecular photoreactions have been studied within zeolite was the use of triplet sensitizer. In order to study this approach, the reaction intermediate in the photosensitization of triethylamine (TEA) by AcP in NaY zeolite was further investigated by using ESR. Detailed mechanism were devised to explain the range of products and the overall kinetic course of the reaction. ESR spectroscopy has extended mechanistic studies since radical intermediates can now directly be detected and identified, and the intermediate rates of decay and the interconversion of the radical can be measured. Photolysis is probably the most common and versatile method to generate radicals. Using a microwave cavity field with grating designed to let light shine onto the cell, but not to perturb the microwaves, in-situ photolysis becomes quite simple. The technique is very effective because high stationary concentration of the radicals could be accumulated even when they are lost by diffusion controlled radical-radical reactions [97, 101]. The photosensitizations of aliphatic amines by acetophenone inside NaY zeolite have been studied by Scaino et al. [26] using laser flash photolysis (308 nm). The results showed a distinctive pattern including detection of amine dimer radical cations which were not observed for the same reaction in polar solvents. Analysis of the products have revealed the formation of hydrazines. In this study ESR spectroscopy is used instead of laser flash photolysis to study the free radical mechanisms of the photosensitization of amines occurred inside zeolite NaY. The disadvantage of laser flash photolysis is that the optical spectra are often very broad, and hence may lack the detail needed to give firm identification. This means that arguments based on kinetic studies, product analysis and simply chemical expectation have often been used to identify a given intermediate [101]. Hopefully with ESR, the mechanism could be further established. 59 The sample NaY zeolite containing AcP and TEA was prepared by sequential adsorptions. In- situ UV irradiations were carried out on the dried sample in ESR spectrometer. The ESR spectrum of AcP-zeolite (Figure 4.5) showed only a singlet peak with g = 1.9996 indicated the AcP triplet state, AcP*. AcP was reported to have its two triplets so close in energy that they may be inverted simply by adjusting the polarity of the solvent. In non-polar solvents the lowest triplet state of AcP is known to be nπ* not ππ* which is slightly higher in energy than nπ* [121]. In hydrocarbon glass, the emission from AcP is characteristic of the nπ* triplet, whereas in polar hydrogen-bonding media such as silica gel, a long lived (~ 300 ms) emission characteristic of the ππ* triplet is observed [122,123]. Shailaja et al. [124] have reported the nature of the emitting triplet in AcP switches from the nπ* (nonpolar media) to ππ* within zeolites. In LiY, NaY, and KY zeolites, the lowest triplet state is identified as ππ*, whereas in RbY and CsY, two emissions characteristic of nπ* and ππ* were observed. In this case, the cation-carbonyl interaction, not polarity was reported responsible for the state switch. Beside, the nature and reactivity of the AcP triplet also was found strongly depends on the phenyl ring substitution [125]. When further experiment was performed on TEA-zeolite, no ESR peak was observed (Figure 4.6). This may imply that TEA could not be easily excited. O g = 1.9996 CH3 AcP* mT Figure 4.5: ESR spectrum of UV photolysis (after1 hour) of AcP in NaY zeolite. 60 mT Figure 4.6: ESR spectrum of UV photolysis (after1 hour) of TEA in NaY zeolite When the zeolite sample containing AcP and TEA was UV irradiated, the resulting ESR spectrum (Figure 4.7) shows several prominent peaks. Albeit complex, the spectrum could be characterized with detail analysis. The five peaks at 318.04, 319.81, 321.60, 323.32, and 325.02 mT (labeled as “x” in the spectrum) with hyperfine splitting constant, A = 1.8 mT are assigned to amine dimer cation radical (I (spin) for nitrogen atom = 1). The total of lines observed for the amine dimer cation radical are in agreement with the formula N = 2nI + 1, where ‘N’ is the number of line and ‘n’ is the number of atoms with spin I. Even though the intensity of the peaks does not exactly give the ratio 1: 2: 3: 2: 1 as suggested by Pascal’s Triangle Theory, it could still be accepted considering the presence of anisotropic effect in a solid state. mT Figure 4.7: ESR spectrum of UV photolysis (after1 hour) of AcP + TEA in the NaY zeolite supercages. 61 The appearance of the peak with g value of 2.0023 (value for free electron) suggests that upon light absorption, a single electron transfer (SET) from the amine to AcP triplet excited state was taken place inside NaY zeolite supercages. In this case, the TEA acts as electron donor while the AcP as the acceptor. The process of SET of TEA was photoinduced by AcP (photosensitizer) for it to achieve the excited triplet state. No peaks corresponding to TEA radical cation was detected in the ESR spectrum. This could be due to its short lifetime. The obtained results are comparable to the same reaction studied by Scaino et al. [26] using laser flash photolysis. Scheme 4.1 shows the proposed mechanism [26]. They reported that the most remarkable observation of the AcP-amine system inside zeolite is the detection of radical cation of amine dimer. These species have not been observed in solution, except in the case of rigid diamine holding the two nitrogen atoms in close proximity [26, 126]. This reflected the fact that the photolysis was taking place in a confined space. Thus, for those amine radical cation sharing the zeolite cage with another molecule of neutral amine, the interaction [N….N]+• has to be favored as the result of mobility and conformation restriction imposed by the rigid framework. This so-called internal-pressure effect was previously observed for charge-transfer complexes [127, 128]. The distinctive behavior of NaY zeolite arises from the combined contribution of a polar environment stabilizing positively charged intermediates and confined reaction cavity favoring aggregation of amine radical cation. O CH3 hv H AcP * OH AcP CH3 AcP Et N Et Et Et N Et Et Et Et N Et Et N Et Et Scheme 4.1: The proposed mechanism of amine photosensitization by AcP inside NaY zeolite supecages [26]. 62 4.3 Alkali Metals Cation-Exchanged Y Zeolites In order to study the size constriction effect and cation-guest interactions played by different cations in the photoreaction, NaY zeolite was further exchanged with alkali metal cations (Li, K, Rb, and Cs) from the respective nitrate solutions. Figure 4.8 shows the X-ray diffractograms of the alkali metal cationexchanged Y zeolite, MY (M= Li, K, Rb, and Cs) compared to NaY zeolite. The XRD profile of the parent NaY sample and all of its modified forms essentially showed that the characteristic peaks closely match those of the reported data [129]. The relative intensities of the XRD peaks of cation-exchanged zeolites were found to be affected to different extents depending on the nature and the concentration of nonframework cationic size without any significant shift in the positions of reflections [118]. The influences of the extraframework cationic size on the change in the relative intensities of the characteristic peaks were also examined assuming identical site occupancies of the different cations. The plane’s values of the Y zeolite were obtained from literature [129]. The relative intensity of the peak on the plane {642}, 2θ = 27.04o was found to be least affected by the size of the cation. However, the decrease in the relative intensity due to plane {111}, 2θ = 6.27o and an increase in the relative intensity due to plane {733}, 2θ = 29.63o with increased cationic size were observed. The variation in peak intensity may be attributed to the higher/lower scattering power of X-rays because of the variation in the charge-to-size ratio of cationic species and the framework distortion to some extent [118, 130]. The diffractograms of the samples also showed an increased background from LiY to CsY (compared to parent NaY), indicating an increased amorphous fraction in the material. It may due to the increasing of basicity with the decreasing electronegativity of the alkali cation. 63 3000 2900 2800 2700 2600 X Relative Intensities (Cps) 2500 2400 2300 X 2200 2100 2000 X 1900 1800 NaY 1700 1600 1500 1400 1300 LiY 1200 1100 1000 900 KY 800 700 600 500 RbY 400 300 200 CsY 100 2 10 20 30 40 50 60 2-Theta - Scale Figure 4.8: X-ray diffractograms of the alkali metal cation-exchanged Y zeolites compared to parent NaY zeolite. 64 Graph of crystallinity versus various MY zeolites (Figure 4.9) was plotted using average value (Cps) of total the intensities on plane {533}, {642} and {840} correspondence to 2θ 23.65o, 27.04o, and 32.46o of the parent NaY. These peaks were marked by a “X” in Figure 4.8. Crystallinity 600 500 400 300 200 100 0 NaY LiY KY RbY CsY Cation-exchanged Y zeolites Figure 4.9: Crystallinity versus cation-exchanged Y zeolites. The degree of ion exchange was determined by flame emission photometry (FEP). The exchange levels were calculated based on the replacement of Na cation in the zeolites (assuming that NaY was 100% ion exchanged). Figure 4.10 is the calibration line of Na FEP analysis. Table 4.1 shows the concentration of Na+ in the Emission Intensity exchanged zeolites and its corresponding exchange levels. 1 y = 0.0069x + 0.0911 R2 = 0.9984 0.8 0.6 0.4 0.2 0 0 50 100 150 Concentration (mg/L) Figure 4.10: Emission intensity versus concentration of Na analysis in flame emission photometry . 65 Table 4.1: Exchanged levels of alkali metal cations exchanged Y zeolites. MY zeolites Emission Intensity NaY LiY KY RbY CsY 0.584 0.208 0.031 0.211 0.263 Na+ concentration (mg/L) 71.434 12.590 Negligible 17.377 20.565 Exchange level 100.0% 82.4% ~ 99% 75.7% 71.3% The exchanged levels and distributions of cation in Y zeolite are very much depends on the size and the nature of the cations. The exchange levels of RbY (75.7%) and CsY (71.3%) are relatively low compared to LiY (82.4%) and KY (almost all the Na+ have been replaced) which has been explained by their inability to replace the site I cation in the hexagonal prism due to the increase in atom radius of Rb+ and Cs+[119,131]. However, the exchange levels of 80 % have been achieved in conventional exchange of NaY with Rb+ and Cs+ [132,133]. It was reported that the Li+ ions are not preferred in exchanged the cations in site I and site I’ due to its hydrated ions which has bigger size compared to K+ ions and weak coulombic interactions between the hydrated counterions and anionic sites [131]. It gives the reason of the low exchange level of Li+ compared to K+. 4.4 Photochemical Hydrogen Abstraction by AcP in Toluene Solution and NaY Zeolites Slurry In this part, we continued our study to another important approach in the intermolecular photoreaction within zeolite, the “spectator” approach. This approach was studied in the photochemical hydrogen abstraction by AcP in zeolite-toluene slurry. 4.4.1 Homogeneous Photoreaction Photochemical hydrogen abstraction by AcP (8) in toluene (9) homogeneous solution and NaY zeolite slurry have been studied in this research. The 66 homogeneous reaction gave 95 % conversion after 5 hours of irradiation under inert condition. Figure 4.11 shows the GC chromatograms before and after the hydrogen abstraction reaction. The GC analysis of the photoproducts showed the presence of 4 significant peaks, each with Rt value of 18.62, 20.85, 23.19, and 23.29 minutes. Toluene AcP (a) Before photoreaction Peak 1 m-Xylene Peak 2 Toluene Peak 3 & 4 Unknown product (14) (b) After photoreaction AcP 5 10 15 20 25 30 min Figure 4.11: GC chromatograms (a) before and (b) after the homogeneous photoreactions of acetophenone in toluene solution. Peak 1 at Rt 18.62 minutes, gave a molecular ion peak at m/z 182 (Appendix 1) in GCMS analysis which was in agreement with a molecular formula of C14H14. The fragmentation pattern of this compound was matched with 1,2-diphenylethane (DPE) (10) cited by the Wiley database of the GCMS system. 67 Peak 2 (Rt 20.85 minutes) which was the major peak in the mixture (37 %) showed an ion peak at m/z 194 (Appendix 2). It’s fragmentation pattern was 90 % matched with 2,3-diphenylpropan-2-ol (DPP) (11) which has a molecular formula C15H16O (M+ 212). The ion peak at m/z 194 confirmed the loss of a molecule of H2O from the molecular ion Peak 3 (Rt 23.19 minutes) and 4 (Rt 23.29 minutes) gave the same fragmentation patterns and the ion peak at m/z 208 in their GCMS spectrum (Appendix 3). The fragmentation of both peaks were matched with 2,3diphenylbutan-2,3-diols (DPB) (12), with molecular formula C16H18O2 (M+ 242). The ion peak at m/z 208 was suggested due to the loss of a H2O and CH3 from the parent ion. The 1H NMR spectrum (Appendix 4) of the isolated compound (PH) exhibited two singlet signals at δ 2.30 and 2.60 attributed to two hydroxyl protons of the compound. Another two singlets at δ 1.53 and 1.61 were assigned to two methyl groups. A multiplet signal which resonated at δ 7.20 – 7.29 was due to ten aryl protons. A small peak with Rt19.60 was also observed in the GC chromatogram. It gave an ion peak at m/z 194 in the MS spectrum. The compound was unable to identify and was named as unknown product (14). Interestingly, there was no 1,2phenylethanol (13) being observed as proposed in the mechanism (Scheme 4.2). The products ratios in this photoreaction were calculated using peak areas in the GC chromatogram, with the assumption of all the photoproducts produced the same response to the FID detector. Table 4.2 shows the calculated ratios. OH OH OH (10) (11) (12) 68 Table 4.2: Product ratios calculated by GC in the photochemical hydrogen abstraction by acetophenone in toluene solution. Condition Conversion DPE (10)a DPP (11) DPB (12) Homogeneous 95% 0.31 0.37 0.30 a Unknown product (14) 0.02 Numbers reported are the average of at least two measurements. Error limit of the analysis is ± 3% From Table 4.2 photoproducts (10), (11), and (12) gave the ratio of 1:1:1 different from the reported ratio 1:2:1 by Lei and Turro [134]. The data cannot be compared because the products ratio in this research was calculated based on the GC peak area while the reported ratio was based on isolated yield. 4.4.2 Photoreaction in NaY Zeolite Slurry The zeolite sample containing AcP was irradiated under degassed condition. Extraction of the photoproducts from the zeolite was done by using THF. Figure 4.12 shows the resulting GC chromatograms of the photoreactions in NaY zeolite slurry. The analysis of the supernatant (toluene solution after photoreaction) showed the present of DPE (10) and an unidentified compound (15). Small amounts of AcP and DPP (11) were also detected. The present of AcP shows that the starting material (AcP) might “escaped” from the supercage of zeolite during process of stirring in the photoreaction. DPE (10) and DPP (11) may formed by the hydrogen abstraction occurred in toluene solution. The high yield of DPE (10) could also caused by the benzyl radicals formed from zeolite supercage because these radicals are small in size and less polar compared to AcP-OH radical. The extraction of the photolysed zeolite by THF gave high yield of asymmetric coupling product DPP (11) due to geminate cage combination. The 69 yields of free radical products DPE (10) and DPB (12) were drastically reduced compared to the homogeneous reaction. The unknown compound (14) was also detected. Table 4.3 shows the product ratios of the extract in the photolysed zeolite calculated using product peaks area in the GC chromatogram. Although the conversion rate was not calculated in this experiment, the ratio of total products/starting material (AcP); 1.65, predicted that the conversion rate is much more lower in the zeolite slurry as compared to the homogeneous reaction (the ratio was 45.25) which occurred in the same reaction period. DPE (a) Supernatant after photolysis Unknown product (15) AcP DPP AcP (b) THF extract of the photolysed zeolite DPP Unknown product (14) DPE DPB 12 14 16 18 20 22 24 Figure 4.12: GC chromatograms of the supernatant and the resulting tetrahydrofuran extract. 70 There was no AcP detected after 2 hours of stirring in supernatant. It is due to the polar carbonyl group of AcP which is strongly bound to the zeolite internal surface. The loading level of 4 mg/100 mg (AcP/NaY) was reported to enable each zeolite cavity contained an AcP molecule and a molecule of toluene [134]. The excess toluene molecules will fill the void space of the NaY zeolite supercage. These hydrocarbon molecules are expected to serve as blockers that inhibit the diffusion of geminate radical pairs and therefore inhibit the free radical formation. Figure 4.13 shows the different molecules distribution in homogeneous solution and zeolite slurry. We can conclude that the different in chemoselectivity obtained in zeolite slurry is the result from a combination of strong preferential adsorption of the AcP to the internal surface of NaY zeliote and the inhibition of the diffusional motion of the geminate radical pairs produce by the toluene radicals. Scheme 4.2 summaries the reaction mechanism and also how the high yield of asymmetric coupling product can be obtained by using the “spectator” approach in the zeolite slurry in this study [134]. Table 4.3: Product ratios in the tetrahydrofuran extract of the photolysed NaY zeolite. Condition a Zeolite slurry DPE (10)a 0.13 DPP (11) DPB (12) 0.72 0.04 Unknown product (14) Total products/AcP 0.11 1.65 Numbers reported are the average of at least two measurements. Error limit of the analysis is ± 2% In homogeneous solution In zeolite slurry Figure 4.13: Molecule distributions in homogeneous solution and zeolite slurry (spectator approach) [147]. 71 O Ph C CH3 O Ph C CH3 hv 1 O Ph C CH3 ISC 3 AcP (8) O Ph C CH3 OH Ph C CH3 hv 3 + Ph CH3 zeolite -slurry + PhCH2 Geminate cage combination geminate radical pair Toluene (9) OH Ph C CH2Ph CH3 DPP (11) major product Free Radicals combination homogenous PhCH2CH2Ph + DPE (10) symmetric statistical distributions: OH + Ph C CH3 Ph CH3 (9) 1 OH OH Ph C C Ph CH3 CH3 OH Ph C CH2Ph + CH3 DPP (11) asymmetric : 2 DPB (12) symmetric : 1 OH Ph C CH3 H 2-phenylethanol (13) expected minor product Scheme 4.2: The mechanism of photochemical hydrogen abstraction by acetophenone in toluene solution and zeolite NaY slurry [134]. 72 4.5 Regioselective Photodimerizations of 2-Cyclohexenone (CH) in Alkali Metal Cation-Exchanged Y Zeolites We continued our research on the confiment effect of zeolite by further study the effect of alkali metal cation-exchanged Y zeolites in the photodimerization of cyclohexenone. 4.5.1 Photodimerizations of 2-Cyclohexenone in Homogeneous Solution Homogenoues photodimerization of 2-cyclohexenone (CH) was first carried out in n-hexane. Reaction in open air was used as control experiments. The GC analysis of the photoproducts (Figure 4.14) showed the presence of four peaks, each with Rt values of 23.40, 23.60, 23.89, and 24.12 minutes. The MS analysis showed that peaks 1-4 gave the same molecule ion peaks at m/z 192 (Appendix 5-8) which was in agreement with a molecular formula of C12H16O2. The fragmentation pattern of this compound was matched with the fragmentation pattern of CH dimer. Peak 2 (Rt 23.60 minutes) and peak 4 (Rt 24.12 minutes) were further confirmed to be dimers of head-to-tail (HT) (16) and head-tohead (HH) (17) with the comparison of literatures [6, 10,135,136]. Peak 1 (Rt 23.40 minutes) and peak 3 (Rt 23.89 minutes) were assigned to either dimer (18) or (19) and classified under other products. Scheme 4.3 summarizes the products distribution in the photodimerization of CH [10]. Products ratios were calculated using the peak area in GC chromatograms. Table 4.4 shows the (16)/(17) dimer ratios in the homogeneous photodimerizations of CH in n-hexane in different conditions. Figure 4.14 also shows the significant different of product selectivity of the photoreactions conducted in homogeneous condition compared to solid-state reaction. 73 HT (16) 2 3 1 (a) Homogeneous photoreaction in inert Condition (b) Homogeneous photoreaction in free air condition 4 HH (17) HT (16) HH (17) HH (17) (c) HT (16) 20 22 24 26 Solid state photoreaction in CsY 28 30 Figure 4.14: GC chromatograms of the homogeneous photoreactions of 2- cyclohexenone compared to solid state photoreactions (a)-(c). 74 O O O O O O hv + + (16) O anti-cis HT (1) (17) anti-cis HH O (18) trans-fused HH O O or + (19a) O alternative trans-fused HT (19b) alternative trans-fused HH Scheme 4.3: Photodimerization of 2-cyclohexenone (1) [10]. Table 4.4: Product ratios of the photodimerizations of 2-cyclohexenone in n-hexane. Condition Conversion HT (16)a HH (17) (16)/(17) Inert Open air 67% 10% 0.61 0.65 0.16 0.17 3.81 3.82 a Other products 0.23 0.28 Numbers reported are the average of at least two measurements. Error limit of the analysis is ± 3% Table 4.4 shows that the ratio of (16)/(17) is independent from the % of conversion. The existence of oxygen molecules in the open air experiment will not inhibit the photoreaction but will slow the rate of the reaction. Oxygen usually will form superoxide radical which can terminate the radical reaction. It explains the low conversion compared to the same experiment carried in inert condition. It is reported that the ratio of anti-dimers HT (16) and HH (17) is in the function of both the solvent and the concentration of the starting material, CH. In non-polar solvent, the photodimerization of CH gives predominantly the HT photodimer (in n-hexane, ratio HT/HH 5.21:1), but the regioselectivity switched in favor of the HH photodimer in polar solvent (in acetonitrile, ratio HT/HH 0.5:1). Irradiation of 0.5 M and 3.0 M CH in benzene give the HT/HH ratio of 2.50:1 and 1.52:1 respectively and photodimerization of the neat ketone gave the dimer ratio 1.06:1 [6, 10,135, 136]. 75 Lam et al. [135] and Wagner and Bucheck [6] have suggested that the photodimerization of CH occurs via it lowest triplet states, ππ*. It is believed there are two distinct structural possibilities for the metastable intermediate: an excited state charge-transfer complex and a 1,4-biradical. The charge-transfer complex itself cannot provide a completely satisfying explanation of the stereochemistry of the photodimerizations [6]. Corey’s model (Figure 4.15) would predict the HH dimer is predominant in this reaction. O δ− O δ− O δ+ δ+ δ+ better than δ+ O ππ* (b) HT alignment 3 ππ* (a) HH alignment 3 Figure 4.15: Corey’s model [5]. Another factor influences the course of cycloaddition is the dipole moment of the collision complex that precedes chemical reaction. A HT alignment of the excited enone and ground state enone should be greatly favored over a HH approach in non polar solvents, but less so in polar solvents. The Scheme 4.4 incorporates the various intermediates which can lead to dimers. A and B are probably π complex (exciplexes) but may be simply collision complexes intermediates. In either case, the specific charge-transfer interactions suggested by Corey would favor B while dipole effects would favor A. In non-polar solvents, biradicals b and c would probably rotate to conformations with lower dipole moments but which would have their radical sites too far apart for effective coupling. Polar solvent however would help to maintain the dipole moments in b and c [6]. However, evidences have been suggested that exciplexes may not involved in cycloadditon reaction of alkenes with cyclic enones and the regiochemistry can be explained in terms of the properties of triplet 1,4-biradiacl intermediates [4, 137139]. 76 O* O* O HT O + A O* O O O a O O O O or B b HH c Scheme 4.4: Various intermediates which can lead to 2-cyclohexenone dimers. 4.5.2 Solid State Photodimerizations of 2-Cyclohexenone in Alkali Metal Cation-Exchanged Y Zeolites The [2+2] photocycloaddition of enone to alkene is one of the most widely used photochemical reactions. Several factors in modifying the regioselectivity and stereoselectivity such as chain length [140, 141], substituents of the system [142,143] and incorporation of the conjugated double bond into a ring [143], have been reported. In addition, reaction medium also seems to be a main factor in controlling the regiochemistry and stereochemistry. In this research, alkali metal cation exchanged Y zeolites were used to obtain the selectivity of the photoproducts in [2+2] photodimerization of CH. The relative efficiency of dimerization was calculated by irradiating all five zeolite complexes under identical condition, thus the results of this experiment are comparable The zeolite samples containing CH were prepared by stirring the zeolite in hexane. The samples were then degassed and UV irradiated as dry powder. The photoproducts were firstly obtained by stirring the photolysed samples overnight in THF. Table 4.5 shows the product ratios obtained from of the solid state photodimerization of CH in cation-exchanged Y zeolites using THF extraction. 77 Table 4.5: Product ratios of the solid state photodimerizations of 2-cyclohexenone in alkali metal cation-exchanged Y zeolites with tetrahydrofuran extractions. a M+Y LiY NaY KY RbY CsY HT (16)a 0.06 0.13 0.03 0.06 0.02 HH (17) 0.70 0.21 0.88 0.78 0.86 (16)/(17) 0.09 0.62 0.03 0.08 0.02 Other products 0.24 0.66 0.09 0.16 012 Numbers reported are the average of at least two measurements. Error limit of the analysis is ± 2% However, when the remained zeolites were further dissolved in 1N HCl and extracted using ethyl acetate (EtOAc), the resulting GC chromatograms showed there were still some products (residue) trapped in the zeolites. Thus we concluded that the THF extraction was not a suitable method to extract the product. GC chromatograms of LiYCH-THF-HCl (Figure 4.16 (b)) and CsYCH-THF-HCl (Figure 4.16 (d)) clearly show the remained products which trapped in the zeolites after THF extractions. The photoreactions were repeated by using other extraction method. 78 HH (a) LiYCH-THF HT 1 4 HH 2 HT 3 5 (b) LiYCH-THF-HCl 6 7 HH (c) CsYCH-THF HH HT 20 22 (d) CsYCH-THF-HCl 24 26 28 Figure 4.16: GC chromatograms (b) and (d) show the remained products which trapped in the zeolites after tetrahydrofuran extractions. 30 79 The photolysed zeolites were dissolved in acid 1 N HCl followed by ethyl acetate (EtOAc) extraction. The dimers have been proved to be stable under acid condition [10]. The resulting chromatograms (Figure 4.17) gave 7 peaks (can be observed clearly in LiY and NaY) instead of 4 peaks in homogeneous reaction (Figure 4.14). The extra peaks 5, 6, and 7 (Figure 4.17) (show with arrows in chromatogram) were observed at Rt 24.54, 25.66 and 25.87 minutes. Peak 5 and 6 gave ion peaks at m/z 192 in their MS spectra. Their fragmentation patterns were matched with the fragmentation patterns of CH dimer. While peak 7 gave the ion peak at m/z 207. Their MS pattern is similar to CH dimers. It is believed that these unknown products are CH dimer with a methyl substitution. The product ratios were calculated based on the peak areas of these 7 peaks in the GC chromatogram and was shown in Table 4.6. The dimer ratios did not change even the dimers were allowed to remain within the zeolite for up to 24 h after irradiation. Table 4.6: Products ratio obtained by solid state photodimerization of 2- cyclohexenone in alkali metal cation-exchanged Y zeolites with HCl treatment and ethyl acetate extraction. a M+Y HT (16)a HH (17) (16)/(17) Other products LiY NaY KY RbY CsY 0.10 0.15 0.11 0.05 0.05 0.50 0.31 0.71 0.81 0.80 0.20 0.48 0.15 0.06 0.06 0.40 0.54 0.14 0.15 0.15 Total products/ (CH+ Total products) 0.37 0.81 0.96 0.63 0.98 Numbers reported are the average of at least two measurements. Error limit of the analysis is ± 2% 80 4 HH 1 HT 2 (a) CH-LiY-HCl HH HT 1 6 5 3 7 (b) CH-NaY-HCl HH HT (c)CH-KY-HCl HH 4 1 5 HT 6 (d) CH-RbYHCl HH (e) CH-CsY-HCl HT 20 22 24 26 28 30 Figure 4.17: GC chromatograms of the solid state photodimerizations of 2- cyclohexenone in alkali metal cation-exchanged Y zeolites (a)-(e). 81 It is observed that the solid state photoreactions give a distinctive difference in products selectivity compared to homogeneous reactions (shown in Figure 4.14). The major photoproduct in homogenous reaction is HT (16) (Table 4.4) while HH (17) become dominant in all the solid state photoreactions (Table 4.6). In most zeolites, the minor (other) products are formed in comparable amounts (Table 4.6). In NaY and LiY they had became prominent products. In NaY, these products formed (54 % of total products) even greater than HT (16) and HH (17) dimers. It further showed that the solid state photodimerization of CH can result in totally different product selectivity compared to solution reaction (Table 4.4). It is suggested that the complexing effect of the charge compensating cation and the size constriction factor in the zeolite supercage is the main factor which contributes to the different products selectivites compared to homogenous reaction [10, 144]. There is a great decrease in the HT (16)/ HH (17) ratio from NaY (0.48) to CsY (0.06) in solid state photoreactions (Table 4.6). The greatest reversal in the regiochemisrty from HT to HH dimer in RbY and CsY could be due to the smallest supercage volume and weakest electrostatic interaction of ion in the zeolites [10, 145]. In RbY and CsY, the constriction factor most probably is the dominant factor. It was reported even at very low loading levels there are supercages with double occupancies [146]. It has been established through solid state NMR and diffusion measurement studies, the translation and rotation motions of aromatic as well as aliphatic molecules are reduced within zeolites [147]. Thus, the formation of high HH in RbY and CsY are caused mainly by the confined space. While in LiY, the relatively large effect of Li+ (strong van der Waals) are expected to provide strong interaction/binding between the carbonyl group of CH with the zeolite surface [28,144]. LiY has been reported to give the greatest enhancement of Norrish Type I products in the photolysis of macrocyclic ketone within zeolites [144]. Lithium ions have been suggested to exert a large effect due to their small size, give rise to high charge density [148,149]. The factor of strong electrostatic field in LiY may suggest why LiY give the higher HH product compared to NaY with smaller supercage volume. 82 The results obtained are compared to experiments done by Lem et al. [10]. Most of the ratio HT(16)/HH(17) are quite similar (for the case NaY, KY, and CsY). While the ratio gives a significant different in LiY (this research 0.20, Lem et al. 0.46). It probably due to the higher exchanged level of Li obtained in this experiment. RbY was not tested by Lem et al. Figure 4.18 shows the ratio HT(16)/HH(17) obtained in this research compared to Lem et al. [10]. Conversion range of 5-40% was reported by Lem et al. [10] correspondent to 1.25 hours of irradiation time. Although conversion was not calculated in this experiment, the wide range differences between the ratio of total products/(starting material (CH) + total products) (Table 4.6) indicated the big different of product conversion rates in various cation-exchanged zeolites. It was shown that products ratios also do not change accordingly as a function of irradiation times (5 hours in this experiment compared to 1.25 hours in Lem et al.’s experiment) (Figure 4.18). Ratio (16)/(17) 0.6 0.5 0.4 This Research Lem et al. 0.3 0.2 0.1 0 LiY NaY KY RbY CsY Zeolites Figure 4.18: Ratio HT(16)/HH(17) obtained in this research compared to ratio obtained by Lem et al. [10]. 83 4.5.3 Photodimerizations of 2-Cyclohexenone in Alkali Metal CationExchanged Y Zeolite Slurries Different handling in the procedures for sample preparing as well as the different method of sample preparation sometimes will induce contradicting observations even for similar system [1]. Majority of the studies of photochemistry in zeolite are concerned with photolyses of organic molecules in dry powder zeolite (solid state) in the absent of solvent. It has been shown that the product distribution obtained upon UV irradiation of organic molecules included in zeolite-solvent slurries is distinctly different from conventional dry powder photolysis. The difference in the product distribution obtained between zeolite-solvent slurry and a homogeneous solution is often higher than that between the dry powder zeolites and homogeneous solution. Solvent present within the supercages of zeolite X and Y was reported to provide constraint on the mobility of the included guest molecules, thus one might able to modify the photoreactivity of the guest molecules. Photolysis of acenapthylene [12] in the RbY-hexane slurry gave a high yield of cis dimer compared to solid state reaction and it was believed that the migration of acenaphtlylene between cages was blocked or inhibited by the solvent hexane. In order to study the effect of solvent on the HT(16) /HH (17) ratio, the reactions were carried out in zeolite-hexane slurry. Since the adsorption of CH is achieved by the same process (stirring in hexane) both for dry (hexane evaporated off) and slurry (hexane left within zeolite) irradiations, we may assume that the distribution pattern remains the same for both the dry and slurry samples. Although the dimerization of CH is more complicated compared to dimerization of acenaphtlylene (which gives only two products), we tried to make a relation between the formations of HH (17) to cis dimer of acenaphtlylene within zeolites. If the high yield of HH (17) in the solid state reaction are mainly due to the size constriction effects within the supercage, this would be expected to get a higher yield of HH (17) in zeolite-hexane slurry system. Hexane was always reported to be the best solvent for zeolite-solvent slurry preparations since total adsorption of the reactant molecules to zeolites occurred in hexane slurries [12,150]. 84 Figure 4.19 shows the resulted GC chromatograms of the photoproducts obtained from the photodimerizations of CH in alkali metal cation-exchanged Y zeolite-hexane slurries. Most of the chromatograms show similar pattern with those obtained in dry powder (solid state) reaction (Figure 4.17), except in the NaY and LiY zeolite- slurry systems. These systems showed a significant increase of background peaks at the Rt 20-23 minutes (shown with circles in the chromatogram), which were not observed in the solid state systems. These products could not be identified. Solid state reaction seen to be provided a “cleaner” reaction compared to slurry system. An extra peak (marked with arrow in Figure 4.19, Peak 8) with the Rt 26.93 minutes could be clearly observed in LiY, NaY and KY-slurry systems compared to dry powder systems. This peak gave a molecular ion peak at m/z 277 in the MS analysis. Its fragmentation patter did not matched with any of the compounds cited in the Wiley database of the GC-MS system. However, this unknown compound (Peak 8) formed an important portion (12 %) of total products in LiY-slurry system. The product distributions are calculated based on the peak areas of these 8 peaks in the GC chromatogram. Table 4.7 shows the calculated results. 85 HH 8 (a) CH-LiY-S HT HH 1 4 5 HT 6 2 3 7 8 (b) CH-NaY-S HH (c) CH-KY-S HT 4 1 HT HH 5 6 (d) CH-RbY-S HH HT 20 22 (e) CH-CsY-S 24 26 28 30 Figure 4.19: GC chromatograms of the photodimerizations of 2-cyclohexenone in alkali metal cation-exchanged Y zeolite-hexane slurries (a)-(e). 86 Table 4.7: Product ratios of the photodimerizations of 2-cyclohexenone in alkali metal cation-exchanged Y zeolite-hexane-slurries. MY zeolites HT (16)a HH (17) (16)/(17) Other products LiY NaY KY RbY CsY 0.05 0.06 0.06 0.06 0.05 0.32 0.16 0.74 0.71 0.69 0.16 0.38 0.08 0.08 0.07 0.63 0.78 0.20 0.23 0.26 a Total products/ (CH+ Total products) 1.00 0.94 0.92 0.91 0.97 Numbers reported are the average of at least two measurements. Error limit of the analysis is ± 2% In LiY, NaY and KY-slurry systems, the formations of HT (16) are lesser, thus provided lower HT/HH value compared to solid state systems (Table 4.6). The formations of HT (16) in hexane solution are more than 60% (Table 4.4). However, the formations of HT (16) and HH (17) decreased in NaY-slurry system, the “other products” become the main portion (78%) and Peak 1 (trans-fused photodimers (18) or (19) is the prominant compound (28 %). Except for the KY-system, the formations of all HH dimers (17) are lesser compared to solid-state reactions. The zeolite-slurry system gave rise to more “other products” compared to HT (16) and HH (17). Introducing of the solvent (hexane) molecules to zeolite supercages in this expeiment did not give higher yield of HH (17) as expected. Hence, it is further proved that the size constriction effect is not the single factor that bringing higher formation of HH (17) within the zeolite supercage. Generally, conversion of the starting material to products is higher in zeoliteslurry compared to dry powder. It can be observed on the value of total products/(CH + total products). All the zeolite-slurry systems gave a value greater than 0.9 even in a shorter irradiation time (3 hours shorter compared to the solid state reaction). Thus, we concluded that photodimerization of CH in alkali metal cationexchanged Y zeolites in solid state and slurry system has successfully gaining favorable result in controlling the regio and steroeselectivity of the photodimer. The results obtained are comparable to the previous report [10]. The selectivity of the 87 photoproducts is most probably due to the combine factors of size constriction effects and the complexing ability of the charge-compensating cation of the zeolites. In overall, the modified (cation-exchanged) zeolites gave better product selectivity compared to the parent NaY zeolite. The presence of solvent molecules in the system did not give big different in the ratio of HT (16)/HH (17). 4.6 Photocycloadditions of 2-Cyclohexenone to Vinyl Acetate (VA) in Alkali Metal Cation-Exchanged Y Zeolite-Slurries After the photodimerization of 2-cyclohexenone (CH) within zeolites has successfully gaining high regioselective of HH dimer, we tried to apply the same approach in a more complicated reaction. Photodimerization only involved a single reactant. We further our study to another similar cycloaddition reaction, by introducing an alkene (vinyl acetate), VA to the enone within the zeolite. 4.6.1 Homogeneous Solution Photocycloadditon of CH (1) to VA (21) (1: 15 mol) was carried out in hexane solution, in inert condition for 5 hours. The reaction mixture was analyzed using GC and GC-MS. The GC analysis of the mixture (Figure 4.20) showed the presence of five significant peaks, each with Rt values of 19.58, 19.71, 19.83, 19.94, and 20.10 minutes. All these five peaks gave similar fragmentation patterns and the ion peaks at m/z 158, 139, 122, 111, 97, and 43 in their GCMS analysis (Appendix 9 -13). Their fragmentation patterns were matched with the cyclohexane-cyclobutene adduct (22) or (23) (Figure 4.21) which has a molecular formula of C10H14O3 (M+, 182). The ion peak at m/z 122 confirmed the loss of HOCOCH3 from the molecular ion. Small amounts of CH dimer were also detected in this reaction. 88 1 2 45 3 16 18 20 22 24 Figure 4.20: GC analysis on the reaction mixture in the photocycloaddition of vinyl acetate to 2-cyclohexenoen in hexane. O O OAc O OAc hv + + OAc (1) (21) HT (22) HH (23) Figure 4.21: Photocycloaddition of vinyl acetate to 2-cyclohexenone. In early 1960, Corey et al. [5] have reported the photocycloadditon of CH to VA, they have only able to characterize three HT streoisomeric acetoxy ketones as major products. However, we believe the cycloadducts distribution of the addition of CH to VA (CH2=CHOAc) are similar to the addition of CH to ethoxyethene. Seven racemic cycloadducts (25)-(31) in the ratio of 7: 2: 10: 23: 28: 7: 23 were reported by Maradyn et al. [139] in the photocycloaddition of CH (1) to ethoxyethene (CH2=CHOEt) (24) (Scheme 4.5). Although we are not able to determine the stereochemistry of these cycloadducts, we hope to compare the different of product distribution between solution reaction and reaction in confinement space of zeolite. The cyclobutene 89 adducts were then named as (P1)-(P5) and the product ratios were calculated based on the peak areas obtained in the GC analysis. O H OEt O OEt + (1) H hv HH (25)-(27) (24) O H H OEt HT (28), (29) O H H OEt HT (30), (31) Scheme 4.5: Photocycloadditon of 2-cyclohexenone to ethoxyethene. 4.6.2 Photocycloadditions in Alkali Metal Cation Exchanged Y Zeolite Slurries At first, we faced with difficulties in finding a suitable sample preparation method. We tried to include both the CH and VA in zeolite and irradiated it in solid state. However, the products analysis only revealed the CH dimer. It is believed that the VA had all evaporated off in the process of drying and degassing the sample because VA is a highly volatile compound. Thus, the “solid state” method is not suitable for this reaction. Then, we applied the “spectator” zeolite-slurry method (as what we have done in the hydrogen abstraction experiment) by absorbing the CH in VA solution. After the irradiation, the sample turned to become a gummy transparent liquid, stuck with the zeolite. The gummy liquid was believed to be the polymerized form of VA, polyvinyl acetate. Again, this method failed. The triplet sensitization technique was not in our consideration, because it will only complicate the problem. Finally, the reactions were carried out by absorbing both the CH and VA in MY zeolites and photolysed it as hexane slurry. The products were extracted by dissolving the irradiated zeolites with concentrated HCl and isolated with EtOAc. No new products or a difference in the product ratio were observed relative to the 90 mixture before acid treatment (refer to the acid test in Section 3.8.1.1). Figure 4.22 shows the resulted GC chromatograms of the extracted photoadducts and the resulted product ratios in different mediums are reported in Table 4.8. There is no new compound obtained in the photocycloaddition within zeoliteslurries system. However, distinctly difference of product selectivity was observed compared to the homogeneous reaction. There are drastically decrease of product portion of (P2) and increase of the ratio of (P3). Surprisingly, all the MY zeolites provided a similar pattern of product distributions. There is no significant change of product ratios from LiY to CsY. LiY and KY systems almost provided the same product distribution as solution reaction. RbY system gave the highest yield of (P1) while CsY system provided (P5) as the largest portion in the products mixture. We also tried to increase the loading level of CH (4 mg CH/100 mg zeolite) to increase the confinement effect. However the obtained results provided a high yield of CH dimer, indicated that photodimerization reaction was more prominant compared to photocycloaddition of VA. Thus, we tried to reason the failure of cation to vary the selectivity in this reaction is caused by the “passive” cavity. A zeolite reaction cavity has been characterized to be “active” when the interaction between a guest molecule and the cavity is attractive or repulsive. While there is no significant interaction, it is considered to be “passive”. The interactions may vary from weak van der Waals forces, to hydrogen bonds to strong electrostatic forces between charged centers [69]. VA (CH2=CHOCOCH3) molecule has two polar groups, C-O and C=O which can compete with the C=O group in the CH molecule to interact electrostatically with the surface cations. Because the VA molecules are present in a large amount in the supercage, it may also shielded and weaker the electric field created by the cations. It explains why the alkali metal cation-exchanged Y zeolites are not able to control the products selectivity in this system. 1 91 3 5 (a) CHVA-LiY (b) CHVA-NaY (c) CHVA-KY 1 3 5 (d) CHVA-RbY (e) CHVA-CsY 16 18 20 22 24 Figure 4.22: GC chromatograms photproducts in photocycloadditons of 2- cyclohexenone to vinyl acetate in alkali metal cation-exchanged Y zeolite-slurries. 92 Table 4.8: Product ratios obtained in photocycloadditions of 2-cyclohexenone to vinyl acetate in different mediums. Medium (P1)(a) (P2) (P3) (P4) (P5) Hexane LiY NaY KY RbY CsY 0.30 0.21 0.25 0.21 0.34 0.21 0.18 0.05 0.06 0.05 0.05 0.06 0.08 0.28 0.24 0.29 0.21 0.25 0.15 0.08 0.09 0.08 0.05 0.06 0.29 0.38 0.36 0.37 0.35 0.42 a Total products/ (CH+ Total products) 1.00 0.97 1.00 0.83 0.70 1.00 Numbers reported are the average of at least two measurements. Error limit of the analysis is ± 2% CHAPTER 5 CONCLUSIONS In summary, this research have achieved all the stated objectives. The ESR study of paramagnetic probe (H radical) in NaY zeolite showed the radicals stayed in two different adsorption sites. The radicals are more preferable to stay in site II most probably because it is less shielded. ESR results also showed that the radical cation of amine dimer was formed resulted from the constriction effect of the zeolite Y supercage in the photosensitization of triethylamine by acetophenone. Photoreactions in the confined space of zeolite Y supercage produced remarkable differences in product distribution comp red to conventional solution reactions. In the photochemical hydrogen abstraction by acetophenone in toluene, the NaY zeolite-toluene slurry provided high yield of asymmetric product (0.72) compared to homogeneous reaction (0.37). Cation-exchanged Y zeolites has successfully gaining favourable result in controlling the regio and setereoselectivity of photodimer in the solid state photodimerization of 2-cyclohexenone. The reactions showed a great reversal of head to tail (HT) cyclohexenone dimer, to head to head (HH) cyclohexenone dimer increasing from LiY to CsY zeolite. The complexing effect of the charge compensating cation and the size constriction factor is the main factor which contributes to the different product selectivities. The appearance of solvent molecules in zeolite slurry reactions, however did not give any significant change in product distributions compared to solid state reactions. 94 The study of regioselectivity in the photocycloaddition of 2-cyclohexenone to vinyl acetate in zeolite slurry also showed a drastically change of product yield compared to homogeneous reaction. However, the cation-exchanged Y zeolites did not play an important role in controlling the product selectivity. The failure of the alkali metal cations to vary the selectivity was due to the passive cavity effect. The use of zeolites in controlling product selectivity produced favourable results. However, attentions must be given in choosing the suitable zeolites (regarding their pore size, framework structure, etc) and reactants (molecular size, polarity, volatility, etc). The possibility of using other type of zeolites (ZSM-5 or zeolite-beta) in controlling the regioselectiviy of the photocycloaddition of 2cyclohexenone to vinyl acetate should be further investigated. Further experiments can be carried out using different organic molecules from other functional group such as alkene. REFERENCES 1. Hashimoto, S. Zeolite Photochemistry: Impact of Zeolites on Photochemistry and Feedback from Photochemistry to Zeolite Science. J. Photochem. Photobiol. C: Photochem. Rev. 2003. 4: 19-49. 2. Lakshminarasimhan. P. Photochemical Reactions in Zeolites – Effect of Acidity, Confinement and Non-Bonding Interaction. Ph.D. Dissertation. Tulane University; 2001. 3. Weedon, A. C. Synthetic Organic Photochemistry. Horspool, W. M. ed. New York: Plenum. 1984. 4. Schuster, D. I., Lem, G. and Kaprinidis, N. A. New Insight Into Old Mechanism: [2+2] Photocycloaddition of Enones to Alkenes. Chem. Rev.1993, 93: 3-22. 5. Corey, E. J., Bass, J. D., LeMahieu, R. and Mitra, R. B. A Study of the Photochemical Reactions of the 2-Cyclohexenones with Substituted Olefins. J. Am. Chem. Soc. 1964. 86: 5570-5583. 6. Wagner, P. J. and Bucheck, D. J. A Comparison of the Photodimerization of 2-Cyclopentenone and 2-Cyclohexenone in Acetonitrile”, J. Am. Chem. Soc. 1969, 91: 5090-5097. 7. Joy, A., Warrier, M. V. and Ramamurthy, V. Enforcing Molecules to Behave. The Spectrum. 1999. 12: 1-7. 96 8. Turro, N. J. From Molecular Chemistry to Supramolecular Chemistry to Superdupermolecular Chemistry. Controlling Covalent Bond Formation through Non-Covalent and Magnetic Interactions. Chem. Comm. 2002. 22792292. 9. Turro, N. J., Lei, X. G., Li, W., Liu, Z. Q., McDermott, A., Ottaviani, M. F. and Abrams, L. Photochemical and Magnetic Resonance Investigations of the Supramolecular Structure and Dynamics of Molecules and Reactive Radicals on the External and Internal Surface of MFI Zeolites. J. Am. Chem. Soc. 2000. 122: 11649-11659. 10. Lem, G., Kaprinidis, N. A., Schuster, D. I., Ghatlia, N. D. and Turro, N. J. Regioselective Photodimerization of Enones in Zeolites. J. Am. Chem. Soc. 1993. 115: 7009-7010. 11. Turro, N. J., Han, N., Lei, X. G., Fehlner, J. R. and Abrams, L. Mechanism of Dichlorination of n-Dodecane and Chlorination of 1-Chlorododecane Adsorbed on ZSM-5 Zeolite Molecular Sieves. A Supramolecular Structural Interpretation. J. Am. Chem. Soc. 1995. 117: 4881-4893. 12. Ramamurthy, V., Corbin, D. R., Turro, N. J., Zhang, Z. and Garcia-Garibay, M. A. A Comparison between Zeolite-Solvent Slurry and Dry Solid Photolyses. J. Org. Chem. 1991. 56: 255-261. 13. Ghatlia, N. D. and Turro, N. J. Diastereoselective Induction in Radical Coupling Reactions: Photolysis of 2,4-Diphenylpentan-3-ones Adsorbed on Faujasite Zeolites. J. Photochem. Photobiol. A: Chem. 1991. 57: 7-19. 14. Cheung, E., Chong, K. C. W., Jayaraman, S., Ramamurthy, V., Scheffer, J. R. and Trotter, J. Enantio- and Diastereodifferentiating cis, trans- Photoisomerization of 2β, 3β-Diphenylcyclopropane-1α-carboxylic Acid Derivatives in Organized Media. Org. Lett. 2000. 2: 2801-2804. 97 15. Blatter, F. and Frei, H. Selective Photooxidation of Small alkenes by O2 with Red Light in Zeolite Y. J. Am. Chem. Soc. 1994. 116:1812-1820. 16. Vasenkov, S. and Frei, H. UV-Visible Absorption Spectroscopy and Photochemistry of an alkene-O2 Contact Charge-Transfer system in Large NaY Crystals. J. Phys. Chem. B.1997. 101: 4539-4543. 17. Takeya, H., Kuriyama, Y. and Kojima, M. Photooxygenation of Stilbenes in Zeolite by Excitation of Their Contact Charge Transfer Complexes with Oxygen. Tetrahedron Lett. 1998. 39: 5967-5970. 18. Leibovitch, M., Olovsson, G., Sundarababu, G., Ramamurthy, V., Scheffer, J. R and Trotter, J. Asymmetric Induction in Photochemical Reactions Conducted in Zeolites and in Crystalline State. J. Am. Chem. Soc. 1996. 118:1219-1220. 19. Joy, A., Scheffer, J. R. and Ramamurthy, V. Chirally Modified Zeolites as Reaction Media: Photochemistry of an Achiral Tropolone Ether. Org. Lett. 2000. 2: 119-121. 20. Cheung, E., Netherton, M. R., Scheffer, J. R., Trotter J. and Zenova, A. Asymmetric Induction through the Use of Remote Covalent and Ionic Chiral Auxiliaries in the Solid State Photochemistry of 2-Benzoaladamatane-2Carboxylic Acid Derivatives. Tetrahedron Lett. 2000. 41: 9673-9677. 21. Cheung, E., Rademacher, K., Scheffer, J. R. and Trotter, J. An Investigation of the Solid-State Photochemistry of α-Meistylacetophenone Derivatives: Asymmetric Induction Studies and Crystal Structure- Reactivity Relationships. Tetrahedron Let. 2001. 56: 6739-6751. 22. Jayaraman, S., Uppili, S., Natarajan, A., Joy, A., Chong, Kenneth C. W., Netherton, A. R., Zenova, A., Scheffer, J. R. and Ramamurthy, V. The Influence of Chiral Auxiliaries is Enhanced within Zeolites. Tetrahedron Lett. 2000. 41: 8231-8235. 98 23. Natarajan, A., Wang, K., Scheffer, J. R. and Patrick, B. Control of Enantioselectivity in the Photochemical Conversion of α-Oxoamides into βLactam Deriavatives. Org. Lett. 2002. 4: 1443-1446. 24. Jayaraman, S., Scheffer, J. R., Chandarasekhar, J. and Ramamurthy, V. Confined Space and Cations Enhance the Power of a Chiral Auxiliary: Photochemical of 1,2-dipheylcyclopropane Derivatives. Chem. Commun.. 2002. 830-831. 25. Chong, K. C. W., Sivaguru, J., Scichi, T., Yoshimi, Y., Ramamurthy, V. and Scheffer, J. R. Use of Chirally Modified Zeolites and Crystals in Photochemical Asymmetric Synthesis. J. Am. Chem. Soc. 2002. 124: 28582859. 26. Scaiano, J. C., Garcia, S. and Garcia, H. Intrazeolite Photochemistry. 18. Detection of Radical Cations of Amine Dimers upon Amine Photosensitization with Acetophenone in the Zeolite NaY. Tetrahedron Lett. 1997. 38: 5929-5932. 27. Brancaleon, L., Brousmiche, D., Jayathirtha R., Johnston, L. J. and Ramamurthy, V. Photoinduced Electron Transfer Reactions within Zeolites: Detection of Radical Cations and Dimerization of Arlyalkenes. J. Am. Chem. Soc. 1998. 120: 4926-4933. 28. Pitchumani, K., Warrier, M., Scheffer, J. R. and Ramamurthy, V. Novel Approaches Towards the Generation of Excited Triplets of Organic Guest Molecules with Zeolites. Chem. Commun. 1998. 1197-1198. 29. Wada, T., Shikimi, M., Inoue, Y., Lem, G. and Turro, N. J. First Photosensitized Enantiodifferentiating Isomerization by Optically Active Sensitizer Immobilized in Zeolite Supercages. Chem. Commun. 2001. 18641865. 99 30. Frank, J. and Rabinowitch, E. Trans. Faraday Soc. 1934. 30: 120. 31. Barrer, R. M. Hydrothermal Chemistry of Zeolites. London: Academic Press; 1982. 32. Breck, D. W. Zeolite Molecular Sieves: Structure, Chemistry and Use. New York: Wiley Interscience; 1974. 33. Dyer, A. An Introduction to Zeolite Molecular Sieves. New York: Wiley; 1988. 34. Flanigen, E. M. Zeolites and Molecular Sieves. A Historical Perpective. In: van. Bekkum, H., Flanigen, E. M. and Jasen, J. C. eds. Introduction to Zeolite Science and Practice. Amsterdam: Elsevier .1991. 35. Cronstedt, A. F. Akad. Handl., Stoockholm. 1756. 17:120. 36. Mumpton, F. A. Zeolite Exploration: The Early Day. In: D. Olson, and A. Bisio eds. Proc. 6th. Intl. Zeolite Conference. UK: Butterworths-Guiford; 1983. 37. Barrer, R. M. Separation of Mixture Using Zeolites as Molecular Sieve. I. Three Classes of Molecular Sieve. J. Soc. Chem. Ind. 1945. 64: 130-11. 38. Barrer, R. M. Synthesis of a Zeolite Mineral with Chabazite Like Sorptive Properties. J. Chem. Soc. 1948. 217-232. 39. Milton, R. M. Molecular Sieve Science and Technology: A historical Perspective. In: Occelli, M. L. and H. E. Robson eds. Zeolite Synthesis. ACS Symp. Ser. No. 393. Washington: American Chemical Society, 1989. 40. Frost & Sullivan. D458: Zeolites: Industry Trends and Worldwide Market in 2010. New York: A Report from Technical Insights; 2001. 100 41. Tung, C. H., Wu, L. Z., Yuan, Z. Y., Guan, J. Q., Ying, Y. M., Wang, H. H. and Xu., X. H. Remarkable Product Selectivity in Photochemical Reactions within Supramolecular Systems. Materials Sc. & Eng. C. 1999. 10: 75-81. 42. Matsubara, C. and Kojima, M. Effect of Coadsorbed Water on Photodimerization and Photooxygenation of 4-Methoxystrene Included in NaY. Tetrahedron Lett. 1999. 40: 3439-3442. 43. Noh, T., Kwon, H., Kyungsun Choi and Kyungin Choi. Intramolecular Photocycloaddition of 3-(3-Butenyl)cyclohex-2-enone and 3-(2- Propenoxy)cyclohex-2-enone in Zeolites. Bull. Korean Chem. Soc. 1999. 20: 76-80. 44. Yoon, K. B. Electron- and Charge- Transfer Reactions within Zeolites. Chem. Rev., 1993. 93: 321-339. 45. Dempsey, E. Molecular Sieves. London: Society of Chemical Industry; 1968. 46. Preuss, E., Linden, G. and Peuckert, M. Model Calculation of Electrostatic Field and Potentials in Faujasite Type Zeolites. J. Phys. Chem. 1985. 89: 2955-2961. 47. Sarkar, N., Das, K., Nath, D. N. and Bhattacharyya, K. Twisted Charge Transfer Process of Nile Red in Homogeneous Solution and in Faujasite Zeolite. Langmuir. 1994. 10: 326-329. 48. Uppili, S., Thomas, K. J., Crompton, E. M. and Ramamurthy, V. Probing Zeolites with Organic Molecules: Supercages of X and Y are Superpolar. Langmuir. 2000. 16: 265-274. 49. Hashimoto, S. Fukazawa, N, Fukumura, H. and Masuhara, H. Diffuse Reflectance Laser Photolysis Study on Triplet Complex Between Aromatics 101 and Tl+ Included in Tl+-Exchanged X-Type Zeolite. Chem. Phys. Lett.1994. 223: 493-500. 50. Mellot, C., Siminot-Grange, M. H., Pilverdier, E., Bellat, J. P.and Espinat, D. Adsorption of Gaseous p- or m-Xylene in BaX Zeolite: Correlation between Thermodynamic and Crystallographic Studies Langmuir.1995.11: 1726-1730. 51. Vitale, G., Mellot, C. F., Bull, L. M. and Cheetham, A. K. Netron Diffraction and Computational Study of Zeolite NaX: Influent of SIII’ Cations on Its Complex with Benzene. J. Phys. Chem. B.1997.101: 4559-4564. 52. Yeom, Y.H., Kim, A. N., Kim, Y., Song, S. H. and Seff, K. Crystal Structure of a Benzene Sorption Complex of Dehydrated Fully Ca2+-Exchanged Zeolite X. J. Phys. Chem. B.1998.102: 6071-6077. 53. Pitchon, C., Méthivier, A., Siminot-Grange, M. H. and Baerlocher, C., Location of Water and Xylene Molecules Adsorbed on Prehyrated Zeolite BaX. A Low-Temperature Neutron Powder Diffraction Study. J. Phys. Chem. B. 1999. 103:10197-10203. 54. Gokel, G. W., De Wall, S. L. and Meadows, E. S. Experimental Evidence for Alkali Metal Cation-π Interaction. Eur. J. Org. Chem., 2000. 17: 2967-2978. 55. Kim, K. S., Tarakeshwar, P. and Lee, J. Y. Molecular Clusters of π-systems: Theoretical Studies of Structures, Spectra, and Origin of Interaction Energies. Chem. Rev. 2000. 100: 4145-4185. 56. Fujii, T. Alkali Metal Ion/Molecule Association Reactions and Their Applications to Mass Spectrometry. Mass Spectrom. Rev. 2000. 19:11-138. 57. Kirschhock, C. and Fuess, H. m-Dinitrobenzene in Zeolite NaY: Four Different Arrangements. Zeolites. 1996. 17: 381-388. 102 58. Lim, K. H. and Grey, C. P. Characterization of Extra-Framework Cation Positions in Zeolites NaX and NaY with Very Fast 23 Na MAS and Multiple Quantum MAS NMR Spectroscopy. J. Am. Chem. Soc. 2000.122: 9768-9780. 59. Lim, K. H., Jousse, F., Auerbach, S. M. and Grey, C. P. Double Resonance NMR and Molecular Simulation of Hydrocarbon Binding on Faujisite Zeolite NaX and NaY: The Importance of Hydrogen Binding in Controlling Adsorption Geometries. J. Phys. Chem. B. 2001. 105: 9918-9929. 60. Srivatsan N. and Norris, J. R. Electron Paramagnetic Resonance Study of Oxidized B820 Complexes. J. Phys. Chem. B. 2001. 105 : 12139-12398. 61. Thomas, K. J., Sunoj, S. B., Chandrasekhar, J. and Ramamurthy, V. Cation Interaction Promoted Aggregation of Aromatic Molecules and Energy Transfer within Zeolites. Langmuir. 2000. 16: 4912-4921. 62. Ramamurthy, V., Sanderson, D. R. and Eaton, D. F. Distribution of Organic Molecules within Zeolites as Revealed by Aromatic Photophysical Probes: Role of Water and Other Coadsorbents. J. Phys. Chem. 1993. 97: 1338013386. 63. Hong, S. B., Cho, H. M. and Davis M. E. Distribution and Motion Organic Guest Molecules in Zeolites. J. Phys. Chem. 1993. 97:1622-1628. 64. Kärger, J. and Ruthven D. M. Diffusion in Zeolites and Other Molecular Sieves. New York: Wiley; 1992. 65. Chen, N. Y., Degnan, T. F. and Smith, C. M. Molecular Transport and Reaction in Zeolies. New York: VCH; 1994. 66. Auerbach, S. M. and Metiu, H. I. Modeling Orientational Randomization in Zeolites: A New role of Intracage Mobility, Diffusion and Cation Disorder. J. Chem. Phys. 1997. 106: 2893-2905. 103 67. Auerbach, S. M., Bull, L. M., Henson, N. J., Metiu, H. I. and Cheethaam, A. K. Behaviour of Benzene in NaX and NaY Zeolites: A Comparative Study by 2 68. H NMR and Molecular Mechanics. J. Phys. Chem. 1996. 100: 5923-5930. Favre, D. E., Schaefer, D. J, Auerbach, S. M. and Chmelka, B. F. Direct Measurement of Intrercage Hopping in Strongly Absorbing Guest-Zeolite System. Phys. Rev. Lett. 1998. 81: 5852-5855. 69. Ramamurthy, V. Controlling Photochemical Reaction via Confinement: Zeolite. J. Photochem. Photobiol. C: Photochem. Rev. 2000. 1: 145-166. 70. Joy, A. Studies on Asymmetric Photoreactions in Zeolits. Ph.D. Abstract. Tulane University, 2000. 71. Márquez, F., Zicivich-Wilson, C. M., Corma, A., Palomares, E. and García, H. Napthalene Included within All-Silica Zeolites: Influence of the Host on the Napthalene Photophysics. J. Phys. Chem. B. 2001. 105: 9973-9979. 72. Márquez, F., García, H., Palomares, E., Fernández, L. and Corma, A. Spectroscopic Evidence in Support of the Molecular Orbital Confinement Concept: Case of Antracene Incorporated in Zeolites. J. Am. Chem. Soc. 2000. 122: 6520-6521. 73. Márquez, F., Marti, V., Palomares, E., García, H. and Adam, W. Observation of Azo Chromophore Fluorescence and Phosphorescence Emissions from DBH by Applying Exclusively the Orbital Confinement Effect in Siliceous Zeolites Devoid of Charge-Balancing Cations. J. Am. Chem. Soc. 2000, 124: 7264-7265. 74. Corma, A., García, H., Sastre, G. and Viruela, P. M., Activation of Molecules in Confined spaces: An Approach to Zeolite-Guest Supramolecular Systems. J. Phys. Chem. B 1997. 101: 4575-4582. 104 75. Thomas J. B. Primary Photoprocess in the Biology. North Holland: Amsterdam; 1965. 76. Braslacsky, S. E. and Houk, K. N. International Union of Pure and Applied Chemistry. Organic Chemistry Division. Commission on Photochemistry: Glossary of Terms Used in Photochemistry. Pure Appl. Chem. 1988. 60: 1055-1106. 77. Kopecky, J. Organic Photochemistry: A Visual Approach. USA: VCH; 1992. 78. Kan R. O. Organic Photochemistry. New York: McGraw-Hill; 1966. 79. Barltrop J. A. and Coyle J. D. Principle of Photochemistry. London: Wiley; 1975. 80. Jaffe, H. H. and Orchin, M. Theory and Applications of Ultraviolet Spectroscopy. London: Wiley; 1962. 81. Turro, N. J. Modern Molecular Photochemistry. Philippines: The Benjamin/Cumming Publishing Company; 1978. 82. Pescock, R. L., Shields, L. D., Cairns, T. and McWilliam, I. G. Modern Methods of Chemical Analysis. 2nd ed. Canada: Wiley; 1968. 83. Rau. H. Spectroscopic Properties of Organic Azo Compounds. Angew. Chem. Int. Ed. Engl. 1973. 12: 224-235. 84. March, J. Advanced Organic Chemistry: Reactions, Mechanisms and Structure. London: McGraw-Hill International; 1977. 85. Kauzman, W. Quantum Chemistry. New York: Academic Press; 1957. . 86. Bruice, P. Y. Organic Chemistry. 2nd ed. United States of America: PrenticeHall; 1998. 105 87. Hoffmann, R. An Extended Huckel Theory. I. Hydrocarbons. J. Chem. Phys. 1963. 39: 1397-1412. 88. Hoffmann, R. Extended Huckel Theory. IV. Carbonium Ions. J. Chem. Phys. 1964. 40: 2480-2488. 89. Zimmerman, H. E. and Swenton. J. S. Mechanistic Organic Photochemistry. VIII. Identification of the n-π* Triplet Excited State in the Rearrangement of 4,3-Diphenylcyclohexadienone. J. Am. Chem. Soc. 1964. 84: 1436-1437. 90. Margaretha, P. Synthesis and Photochemical Behaviour of Enones Lactones. Angew. Chem. Int. Ed. Engl. 1972. 11: 327-330. 91. Baldwin, S. W. and Wilkinson, J. M. Cyclohexenones by the Photoannelation of alkenes with 2,2-Dimethyl-3(2H)-Furanone. Tetrahedron Lett. 1979. 20: 2657-2660. 92. Cantrell, T. S. Heller, W. S. and Williams, J. C. Photocycloadditon Reactions of Some 3-substituted Cyclohexenones. J. Org. Chem. 1969. 34: 509-519. 93. Quevillon, T. M. and Weedon, A. C. The Photochemistry of 3-Nitro-2Cyclohexenone. Tetrahedron Lett. 1996. 37: 3939-3942. 94. Swenton, J. S. and Fritzen, E. L. Jr. Regioselective Photochemical Cycloadditon of 2-Trimethylsilylcyclopentenone. Tetrahedron Lett. 1979. 20 : 1951-1954. 95. Cowan, D. O. and Drisko, R. L. E. The Photodimerization of Acenapthylene. Heavy-atom Solvent Effects. J. Am. Chem. Soc. 1970. 92: 6281-6286. 96. Omar, H. I., Odo, Y., Shigemitsu, Y., Shimo, T. and Somekawa, K. Transition State Analysis on Regioselectivity in [2 + 2] Photocycloaddition Reactions of Substituted 2-Cyclohexenone with Cycloalkenecarboxylates. Tetrahedron. 2003. 59: 8099-8105. 106 97. Wertz, J. E. and Bolton, J. R. Electron Spin Resonance: Elementary Theory and Practical Applications. United States of America: McGraw Hill; 1973. 98. Atkins, P. and Paula, J. de. Atkins’ Physical Chemistry. 7th ed. India: Oxford University Press; 2002. 99. Banwell, C. N. Fundamentals of Molecular Spectroscopy. England: McGraw Hill; 1983. 100. Yacob, A. R. Matrix Isolation Study of Group One Alkali Metals. M. Phil. Thesis. University of Wales Cardiff; 1996. 101. Symons, M. Chemical and Biochemical Aspects of Electron Spin Resonance Spectroscopy. Bath: Van Nostrand Reinhold Company; 1978. 102. Flohr, J. K. X-ray Powder Diffraction. USGS: Science for a Changing World. U. S.: Department of the Interior; 1997. 103. Herschel, J. F. W. Trans. R. Soc. Edin. 1823. 9: 445. 104. Talbot, W. H. F., Brewster’s J. Sci. 1826. 5:77. 105. Schrenk, W. G. Analytical Atomic Spectroscopy. London: Plenum Press; 1975. 106. Dean, J. A. Flame Photometry. United States of America: McGraw-Hill; 1960. 107. Kasai, P. and Bishop, R. J. In: Rabo, J. A. ed. Zeolite Chemistry and Catalysis, ACS Monograph. 171. Washington: American Chemical Society; 1976. 107 108. Ottaviani, M. F., Garibay, M. G. and Turro, N. J. TEMPO Radicals as EPR Probes to Monitor the Adsorption of Different Species into X Zeolite. Coll. and Sur. A: Physiochem. & Eng. Asp. 1993. 72: 321-332. 109. Martini, G., Ottaviani, M. F. and Seravalli, G. L. Electron Spin Resonance of Vanadyl Complexes Adsorbed on Synthetic Zeolites. J. Phys. Chem.1975. 79: 1716-1720. 110. Kevan, L. and Narayana, M. In: Stucky, G. D. and Dwyer, F. G. eds. Intrazeolite Chemistry, ACS Symp. Ser.218. Washington: American Chemical Society. 1982. 111. Damian, G., Cozar, O., Miclăus, V., Paizs, C., Znamirovschi, V., Chis, V. and David, L. ESR Study of the Dynamics of Adsorbed Nitroxide Radicals on Porous Surfaces in the Dehydration Process. Coll. and Sur. A: Physiochem. & Eng. Asp. 1998. 137: 1-6. 112. Jockusch, S., Liu, Z., Ottaviani, M. F. and Turro, N. J. Time Resolved CWEPR Spectroscopy of Powder Samples: Electron Spin Polarization of Nitroxyl Radical Absorbed on NaY zeolite, Generated by Quenching of Excited Triplet Ketones. J. Phys. Chem. B. 2001. 105: 7477-7481. 113. Gutjahr, M., Pöppl, A., Böhlmann, W. and Böttcher, R. Electron Pair Acceptor Properties of Alkali Cations in Zeolite Y: An Electron Spin Resonance Study of Adsorbed Di-tert-butylnitroxide. Coll. and Sur. A: Physiochem. & Eng. Asp. 2001. 189: 93-101. 114. Gutjahr, M., Böttcher, R. and Pöppl, A. Characterization of the Di-tert-butyl Nitroxide: Li+ Adsorption Complex in LiY Zeolites by One- and Two Dimensional Electron Spin-Echo Envelope Modulation Spectroscopy. J. Phys. Chem. B. 2002. 106: 1345-1349. 115. Broussard, L. and Shoemaker, D. P. The Structures of Synthetic Molecular Sieves. J. Am. Chem. Soc. 1960. 82: 1041-1051. 108 116. Turkevich, J. J. Catal. Rev.1968. 1: 1. 117. Jelinek, R., Malek, A. and Ozin. G. A. 23 Na Synchroized Double-Rotation NMR Study of Cs+, Ca2+, and La3+ Cation-Exchanged Sodium Zeolite Y. J. Phys. Chem. 1995. 99: 9236-9240. 118. Hunger, M., Schenk, U. and Buchholz, A. Mobility of Cations and Guest Compounds in Cesium-Exchanged and Impregnated Zeolites Y and X Investigated by High-Temperature MAS NMR Spectroscopy. J. Phys. Chem. B. 2000. 104:12230-12236. 119. Savitz, S., Siperstein, F. R., Huber, R., Tieri, S. M., Gorte, R. J., Myers, A. L., Grey, C. P. and Corbin, D. R. Adsorption of Hydrocarbons HFC-134 and HFC-134A on X and Y Zeolites: Effect of Ion-Exchange on Selectivity and Heat of Adsorption. J. Phys. Chem. B. 1999. 103: 8283-8289. 120. Fitch, A. M., Jobic, H. and Renouprez. A. Localization of Benzene in Sodium-Y Zeolite by Powder Neutron Diffraction. J. Phys. Chem. 1986. 90: 1311-1318. 121. Turro, N. J., Gould, L. R., Liu, J., Jenks, W. S., Staab, H. and Alt, R., Investigation of the Influence of Molecular Geometry on the Spectroscopic and Photochemical Properties of α-Oxo[1.n]paracyclophanes (Cyclophanobenzophenones). J. Am. Chem. Soc. 1989. 111: 6378-6383. 122. Nakayama, T., Sakurai, K., Ushida, K., Kawatsura, K. and Hamanoue, K. Dual Phosphorescence of Benzophenone at 77 K in the Mixed Solvent of 2,2,2-Trifluroethanol and Water. Chem. Phys. Lett. 1989. 164: 557-561. 123. Nakayama, T., Sakurai, K., Ushida, K., Hamanoue, K. and Otani, A. Effects of Water on the Phosphorescence Spectra of Aromatic Carbonyl Compounds. Part 1. Dual Phosphorescence of Benzophenone at 77 K in 2,2,2Trifluroethanol and Water. J. Chem. Soc. Faraday Trans. I. 1991. 87: 449454. 109 124. Shailaja, J., Lakshminarasimhan, P. H., Pradhan, A. R., Sunoj, R. B., Jockusch, S., Karthikeyan, S., Uppili, S., Chandrasekhar, J., Turro, N. J. and Ramamurthy, V. Alkali Ion-Controlled Excited-State Ordering of Acetophenone Included in Zeolites: Emission, Solid State NMR, and Computational Studies. J. Phys. Chem. A., 2003.107: 3187-3198. 125. Sarivastva, S., Tourd, E. and Tascano, J. P. Structural Differents between ππ* and nπ* Acetophenone Triplet Excited States as Revealed by Time-Resolved IR Spectroscopy. J. Am. Chem. Soc. 1998. 120: 6173-6174. 126. Scaiano, J. C., Stewart, L. C., Livant, P. and Majors, A. W. Transient Spectroscopy and Kinetics of the Reactions of Mesocyclic Diamines with tert-Butoxyl and with Ketone Triplets. Effects of Ring Conformation. Can. J. Chem. 1984. 62: 1339-1348. 127. Alvaro, M., García, H., García, S. and Fernández, L. Charge Transfer Complexes between Methylviologen and Aromatic Donors within Faujisite Y: Influence of the Alkaline Metal Counter Cations. Tetrahedron Lett. 1996. 37: 2873-2876. 128. Yoon, K. B., Huh, T. J. and Kochi, J. K. Shape Selective Assemblies of Charge-Transfer Complexes as Molecular Probes for Water Adsorption in Zeolites. J. Phys. Chem. 1995. 99: 7042-7053. 129. Treacy, M. M. J., Higgins, H. B. and Ballmoos, R. V. Collection of Simulated XRD Powder Patterns for Zeolites. 3rd ed. New York: Elsevier Science; 1996. 130. Heydorn, P. C., Jia, C., Herein, D., Pfänder, N., Karge, H. G. and Jentoft, F. C. Structural and Catalytic Properties of Sodium and Cesium exchanged X andY Zeolites, and Germanium-substituted X Zeolite. J. Molecular Catalysis A: Chemical. 2000.162: 227-246. 131. Sherry, H. S. The Ion-Exchanged Properties of Zeolites. I. Univalent Ion Exchange in Synthetic Faujasite. J. Phys. Chem. 1966. 70: 1158-1168. 110 132. Nam, S. S., Kim, H., Kishan, G., Choi, M. J. and Lee, K. W. Catalytic Conversion of Carbon Into Hydrocarbons Over Iron Supported on Alkali IonExchanged Y-Zeolite Catalysts. J. Appl. Catal. A: General. 1999. 179: 155163. 133. Xie, J., Huang, M. and Kaliagiune, S. Characterization of Basicity in Alkali Cation Exchanged Faujisite Zeolites: An XPS Study Using Choloroform as Probe Molecule. Appl. Surf. Sci.1997. 115: 157-165. 134. Lei, X. and Turro, N. J. Photochemical Hydrogen Abstraction by Aromatic Carbonyl Compounds in Zeolite Slurries. J. Photochem. Photobiol. A: Chem. 1992. 69: 53-56. 135. Lam, E. Y. Y., Valentine, D. and Hammond, G. S. Mechanism of Photochemical Reactions in Solution. XLIV: Photodimerization of Cyclohexenone. J. Am. Chem. Soc. 1967. 89: 3482-3487. 136. Valentine, D, Turro, N. J. and Hammond, G. S. Thermal and Photosensitized Dimerizations of Cyclohexadiene. J. Am. Chem. Soc. 1964.86: 5202-5208. 137. Hasting, D. J. and Weedon, A. C. Origin of Regioselectivity in the Photochemical Cycloadditon Reactions of Cyclic Enones with Alkenes: Chemical Trapping Evidence for Structures, Mechanism of Formation, Fates of 1,4-Biradical Intermediates. J. Am. Chem. Soc. 1991. 113: 8525-8527. 138. Andrew, D., Hastings, D. J., Oldroyd, D. L., Rudolph, A., Weedon, A. C., Wong, D. F. and Zhang, B. Triplet 1,4-Biradical Intermediates in the Photocycloaddition Reactions of Enones and N-Acylindoles with Alkenes. Pure Appl. Chem. 1992. 64: 1327-1334. 139. Maradyn, D. J. and Weedon, A. C. The Photochemical Cycloadditon Reaction of 2-Cyclohexenone with Alkenes: Trapping of Triplet 1,4-Biradical Intermediates with Hydrogen Selenide. Tetrahedron Lett. 1994. 35: 81078110. 111 140. Srinivasan, R. and Carlough, K. H. Mercury (3P1) Photosensitized Internal Cycloadditon Reactions in 1,4- 1,5- and 1,6-Dienes. J. Am. Chem. Soc. 1967. 89: 4932-4936. 141. Gleiter, R. and Sander, W. Light-Induced [2+2] Cycloaddition Reactions of Nn-Conjugated Dienes - The Effect of Through-Bond Interaction. Angew. Chem. Int. Ed. Engl. 1985. 24: 566-568. 142. Takashashi, M., Uchino, T. Ohno, K. and Tamura, Y. Intramolecular [2+2] Photocycoadditon of 2-(Alkenyoxy)cyclohex-2-enones. J. Org. Chem. 1983. 48: 4241-4247. 143. Wolff, S. and Agosta, W. C. Regiochemical Control in Intramolecular Photochemical Reactions of 1,5-hexadien-3-ones and 1-Acyl-1,5hexadienes. J. Am. Chem. Soc. 1983. 105: 1292-1299. 144. Ramamurthy, V., Lei, X. G., Turro, N. J., Lewis, T. J. and Scheffer, J. R. Photochemistry of Macro Ketones within Zeolites: Competition between Norrish Type I and Tpye II Reactivity. Tetrahedron Lett. 1991. 32: 76757678. 145. Ramamurthy, V, Eaton, D. F. and Caspar, J. V. Photochemical and Photophysical Studies of Organic Molecules Included within Zeolites. Acc. Chem. Res. 1992. 25: 299-307. 146. Liu, X, Iu, K. K. and Thomas, J. K. Photophysical Properties of Pyrene in Zeolites. J. Phys. Chem. 1989. 93: 4120-4128. 147. Kärger, J., Pfeifer, H., Rauscher, M. and Walter, A. Self-Diffusion of nParaffins in NaX Zeolites. J. Am. Chem. Soc. Faraday Trans. I. 1980. 76: 717-737. 112 148. Ramamurthy, V., Lakshiminarasimham, P., Grey, C. P. and Johnston, L. J. Enegry Transfer, Proton Transfer and Electron Transfer Reactions within Zeolites. J. Chem. Soc. Chem. Commun. 1998. 2411-2548. 149. Ramamurthy, V. Organic Molecular Assemblies in the Solid State. Whitesell, J. K. ed. Chichester: Wiley; 1999. 150. Turro, N. J. and Garcia-Garibay, M. Photochemistry in Organized and Constrained Media. Ramamurthy, V. ed., New York: VCH; 1991. M+, C14H14 m/z 182 113 Appendix 1: MS spectrum of 1,2-diphenylethane (DPE) (10). M+, C15H16O- H2O m/z 194 114 Appendix 2: MS spectrum of 2,3-diphenylpropan-2-ol (DPP) (11). M+, C16H18O2-H2O-2CH3 m/z 195 115 Appendix 3: MS spectrum of 2,3-diphenylbutan-2,3-diols (DPB) (12). 116 Appendix 4: 1H NMR spectrum of 2,3-diphenylbutan-2,3-diols (DPB) (12). M+, C12H16O2 m/z 192 Appendix 5: MS spectrum of CH dimer, HT (16). 117 M+, C12H16O2 m/z 192 118 Appendix 6: MS spectrum of CH dimer, HH (17). M+, C12H16O2 m/z 192 119 Appendix 7: MS spectrum of CH dimer (18) or (19) (Peak 1 in Figure 4.14). M+, C12H16O2 m/z 192 120 Appendix 8: MS spectrum of CH dimer (18) or (19) (Peak 3 in Figure 4.14). M+, C10H14O3 – COCH3 m/z 139 121 Appendix 9: MS spectrum of cyclohexene-cyclobutene adduct (P1) (Peak 1 in Figure 4.20). M+, C10H14O3 – COCH3 m/z 139 122 Appendix 10: MS spectrum of cyclohexene-cyclobutene adduct (P2) (Peak 2 in Figure 4.20). Appendix 11: MS spectrum of cyclohexene-cyclobutene adduct (P3) (Peak 3 in Figure 4.20). 123 M+, C10H14O3 m/z 182 Appendix 12: MS spectrum of cyclohexene-cyclobutene adduct (P4) (Peak 4 in Figure 4.20). 124 M+, C10H14O3 m/z 182 Appendix 13: MS spectrum of cyclohexene-cyclobutene adduct (P5) (Peak 5 in Figure 4.20). 125