3 Q‘ ‘5‘351‘32 a: ' «I h gnaw“ a" . J"). . yr 23%.. If? u. .. A . aw??? . .. .5 . , m& L um 7.. Anumuai. 49.9"“? at . . ..~ ‘ #1,. MW . . ., x , .mz. , w em», 1. Eur mm». £33 . a... g.» fun? 1: «mmfanuum . Sixtflfififi ufifiyuus? 4 . . v. EN". .3915... .. : h..,p.m.f. fiiuflag .. . .wzrm4é. V... . a}... V “.v 3.... r. 5.... \. :1” 3...“. .1 a G? t, \~ . 3&4. 4. :3... 4f . 4.? .4 v... mam. 155..."; V «X Luau , ‘ THESIS 2 mo: This is to certify that the dissertation entitled Design, Synthesis and Catalytic Applications of Mesoporous Silica Molecular Sieves presented by Nenzhong Zhang has been accepted towards fulfillment of the requirements for Ph .0. degree in ChemlStl‘y \J / ”/ Lzyzflyg 1/(zs onazaxnxAAk—- M r professor Date 5/7/99 MSU is an Affirmative Action/Equal Opportunity Institution 0-12771 LIBRARY Mlchlgan State University PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE moo mm.“ DE! DESIGN, SYNTHESIS AND CATALYTIC APPLICATIONS OF MESOPOROUS SILICA MOLECULAR SIEVES BY Wenzhong Zhang A DISSERTATION Submitted to Michigan State University In partial fulfillment of the requirements For the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1999 DESIGN. f M] The dim snmulazed great mNPOl’E: ill [hr of large substran much larger pen sits in the inn reactions. Hexag Wfitonnecrix ii: We branching, and scienriiicali) Whit sites in: in mesostrucrure 311d shape of [hi mmmg siraiecc Wific pore COE {OUHEZTOH and h\{ In adding extremely jmmm ABSTRACT DESIGN, SYNTHESIS AND CATALYTIC APPLICATIONS OF MESOPOROUS SILICA MOLECULAR SIEVES BY Wenzhong Zhang The discovery of the Mobil M418 family of mesoporous molecular sieves has stimulated great interest in surfactant-directed assembly of inorganic mesostructures. The mesopores in these molecular sieves provide new opportunities for catalytic conversion of large substrates in the liquid phase. Even though mesoporous molecular sieves exhibit much larger pore sizes than conventional zeolites, accessibility to the catalytically active sites in the framework of the mesopores is still very crucial for diffusion limited reactions. Hexagonal and cubic mesostructures with uni-dimensional and 3-dimensional pore-connectivity, respectively, show similar catalytic reactivities, despite differences in pore branching, most likely both are diffiision limited. Therefore, it is practically useful and scientifically significant to find some ways to improve the accessibility to the active catalytic sites inside the mesopores. Regardless of the type of inorganic precursors used in mesostructure synthesis, the pore size and connectivity of the framework take the size and shape of the surfactant micelles filling in the structural channels. Therefore, an promising strategy for the design and synthesis of mesoporous molecular sieves with specific pore connectivity and accessibility would be to control the surfactant micelle formation and hydrolysis of the inorganic precursors. In addition to framework mesopores, complementary textural mesopores are extremely important for enhancing the accessibility in catalysis. Our first objective was to tailor W” claimant mlartl)" “JG nucleation l {166i ll) dill led to mm" m to use 0 tailor the f enhibittng Lt the flak} su: mesoscaled the complcr accessibilit} peroxidation the catalysts. thingAl-sub: e“hinted cat Complementa the We Chant What I J and concentra: ,t - 4118.” o i ‘n: the 85 Slim-"n9 “am hfu; tailor both the framework and the textural porosity of HMS silicas assembled through an electrically neutral S°I° mechanism. The approach involved control of the solvent polarity, along with the use of organic auxiliary agent (i.e. mesitylene) to control particle nucleation rates and pore size. Increasing the polarity of the synthesis medium caused faster hydrolysis of silicon alkoxide in the presence of the alkyl amine surfactant, which led to meso-scaled fundamental particles and high textural porosity. Our second objective was to use organic assembly modifiers such as tartaric, oxalic, gluconic and citric acids to tailor the formation of MCM-41 through a 3+1" pathway. Well-ordered MCM-41 exhibiting a flake—shaped particle morphology with pore channels running orthogonal to the flaky surface was obtained. Also, MCM—4l with textural pores in association with meso-scaled fundamental particles was prepared under careful control. The importance of the complementary meso-scaled inter-particle (textural) pores for improving the accessibility of the framework pores in catalysis was demonstrated for the liquid phase peroxidation of 2,6-di-tert-butyl phenol in the presence of Ti-substituted derivatives as the catalysts. Also, liquid phase alkylation of 2,4-di-tert-butyl phenol was carried out using Al-substituted derivatives as the catalysts. By facilitating accessibility, significantly enhanced catalytic activity was observed for both Al-HMS and Al-MCM-4l with complementary textural mesopores and meso-scaled fundamental particles comparable to the pore channel lengths. Variation of surfactant packing parameter g (= V/aol) by changing the structure and concentration of surfactant has been reported in literature as an effective approach to altering the assembled mesostructures. Little is known about chemical modification of surfactant head group using simple chemicals rather than co-surfactants to ultimately tailor the r molecular successful distension. hinthem ammonium relatixely tncorprrnti introductio H mlCt‘ltés a M's-‘OIAWTOL‘ dlmfilhiom hexagonal WIRE“ If; mediated Milena. h‘ 1:03..“ C. lit , m4'veth}ler tailor the molecular sieves. The concept of altering the structures of mesoporous silica molecular sieves by organic promoters in electrostatic S'T assembly has been successfully illustrated. Mesoporous silica molecular sieves with well ordered 3- dimensional hexagonal, cubic, and uni-dimensional hexagonal phases were hydrothermally assembled in the presence of organic modifiers using cetyltrimethyl ammonium bromide as the structure director and sodium silicate as the silicon source at a relatively low surfactant to silicon ratio (i.e. 0.11). In addition, the concept of incorporating structural order into mesoporous silica molecular sieves through the introduction of electrostatic interactions at the interfaces of electrically neutral surfactant micelles and neutral inorganic precursors has also been successfully demonstrated. Mesoporous silica molecular sieves with well ordered hybrid domains consisted of 3- dimensional hexagonal and spherical cubic (3-d-hex-cubic) phases, an uni-dimensional hexagonal phase, and wormhole structural motifs were assembled using alkyl polyethylene oxide surfactants and tetraethyl orthosilicate through a new counterion mediated (N°M*X‘)I° pathway. Our approach explicitly illustrates that silica mesostructures can be altered using both organic and inorganic modifiers such as hydroxyl carboxylic acids for quaternary ammonium surfactant and metal salts for alkyl polyethylene oxide surfactants, respectively. The modifiers most likely bind to the surfactant head group to cause a significant change in the effective-surfactant-head- group-area, which alters the packing parameter g and, ultimately, the inorganic mesostructure. To my grandparent and parants, my wife and my son. I Wt support and scheme» example for met the ye. if. dellM‘ committee. lam [bfi kindness Wang. Dr. \ Pauly for the lam. Oblam XXV assjs’ifllce ire and SEM rm; Mir a Meme {Or} 5011 Dell W ACKNOWLEDGMENT I would like to thank Professor Thomas J. Pinnavaia for his continuously solid support and invaluable advice over the years in both sciences and life. His attitude and seriousness toward the pursuing of the truth and beauty of chemistry set a perfect example for me. I strongly believe that I will greatly benefit from my experiences gained over the years in his lab. I would also like to thank Professor M.G. Kanatzidis, Professor J.F. Harrison and Professor S. Garrett for their guidance in serving on my program committee. I am so lucky to be a member in such a nice research group. I am very grateful for the kindness and help from my colleagues. I would especially thank my friends Dr. Zhen Wang, Dr. Yu Liu, Dr. Seong-Su Kim, Dr. Jialiang Wang, Dr. Tie Lan, and Mr. Thomas Pauly for their generous help. I am also very grateful for Dr. J. Wong and Dr. M. Froba for their measurement to obtain XANES spectra of Ti-substituted mesoporous molecular sieves. Very helpful assistance from Dr. J. Heckman Jr. and Dr. S.L. Flegler to obtain many attractive TEM and SEM images is greatly appreciated. My deepest gratitude goes to my parent, my wife and son for their inspiration and patience for years to allow me work during weekends and holidays. I am obligated to my son, Dejia, who sacrificed his enjoyable kid’s time to allow me to stay in the lab rather than play with him. vi LIST OF TABl LIST OF FlGl' CHAPTER l f Abfiil’m TABLE OF CONTENTS LIST OF TABLES ............................................................................................................. iv LIST OF FIGURES ............................................................................................................. v CHAPTERI SUPRAMOLECULAR ASSEMBLY OF MESOPOROUS MOLECULAR SIEVES AND THEIR APPLICATION IN CATALYSIS ............................................................................................... 1 1.1 Introduction ................................................................................................. 1 1.2 Structural Characteristics of Mesoporous Silica Molecular Sieves Assembled through Neutral and Electrostatic Pathways ............................. 5 1.3 Synthesis of Mesosporous Silica Molecular Sieves Using Primary Amine Surfactant as the Structure Directors ............................................... 9 1.4 Catalysis .................................................................................................... 12 1.4.1 General Application of Mesoporous Molecular Sieves ................ 12 1.4.2 Peroxide Oxidation ........................................................................ 13 1.4.3 Ring Opening Polymerization ....................................................... 17 1.4.4 Selective Catalytic Reduction (SCR) of NO ................................. 19 1.4.5 Acid Catalysis ................................................................................ 20 1 .5 References ................................................................................................. 24 CHAPTER 2 MESOPOROUS TITAN OSILICATE MOLECULAR SIEVES PREPARED AT AMBIENT TEMPERATURE BY ELECTROSTATIC (S+I’, S+X'l+) AND NEUTRAL (Solo) ASSEMBLY PATHWAYS : A COMPARISON OF PHYSICAL PROPERTIES AND CATALYTIC ACTIVITY FOR PEROXIDE OXIDATIONS .......................................................................................... 27 Abstract ................................................................................................................. 27 vii l I l .l 2.3 ltl 2.4 C. 35 Re Clltl’lER3 TA Ol NE Abstract 3.1 lntt 3.2 Ext 3.3 Re- 3.4 Dis 3.5 Rel GWEN [\1 SIE Ab‘lmcr 4'1 lntr 4.3 EXp 43 Re» 4.3,. 2. 1 Introduction ............................................................................................... 28 2.2 Experimental ............................................................................................. 29 2.2.1 Materials ........................................................................................ 29 2.2.2 Characterization ............................................................................ 31 2.2.3 Catalytic experiments .................................................................... 33 2.3 Results and Discussion .............................................................................. 34 2.3.1 Synthesis ....................................................................................... 34 2.3.2 Characterization ............................................................................ 35 2.3.3 Catalytic results ..... , ....................................................................... 48 2.4 Conclusions ............................................................................................... 52 2.5 References ................................................................................................. 54 CHAPTER 3 TAILORING THE FRAMEWORK AND TEXTURAL MESOPORES OF HMS MOLECULAR SIEVES THROUGH AN ELECTRICALLY NEUTRAL (S°I°) ASSEMBLY PATHWAY ........................................... 57 Abstract ................................................................................................................. 57 3. 1 Introduction ............................................................................................... 59 3.2 Experiment ............................................................................ . ................... 61 3.3 Results ....................................................................................................... 63 3.4 Discussion ................................................................................................. 74 3 .5 References ................................................................................................. 8 1 CHAPTER 4 IN-SITU ALUMINATION OF MESOPOROUS HMS MOLECULAR SIEVES SILICA FOR USE AS ALKYLATION CATALYSTS ............. 82 Abstract ................................................................................................................. 82 4. 1 Introduction ............................................................................................... 83 4.2 Experimental ............................................................................................. 85 4.3 Results and Discussion .............................................................................. 88 4.3.1 Catalyst Synthesis ......................................................................... 88 4.3.2 Catalyst Characterization .............................................................. 89 viii 44 4.5 CHAPTER f AIM (J1 I.) (II J) (J! .L LII 'JI 4.3.3 Catalytic Results for 2,4-Di-tert-phenol Alkylation .................... 104 4.4 Conclusions ............................................................................................. 107 4.5 References ............................................................................................... 109 CHAPTER 5 TOWARD MCM-4l WITH MESO-SCALED CHANNEL LENGTHS AND TEXTURAL PORES FOR CATALYSIS ..................................... 111 Abstract ............................................................................................................... 111 5. 1 Introduction ............................................................................................. 1 13 5.2 Experimental ........................................................................................... 1 16 5.3 Results and Discussion ............................................................................ 120 5.3.1 Effect of Organic Acid Addition on Silica Mesostructure Formation ............................................ . ....................................... 120 5.3.2 Siliceous MGM-41 Type Meso-structure Formation .................. 123 5.3.3 Catalyst Preparation .................................................................. 133 5.3.4 Catalytic Alkylation of 2,4-di-tert-butylphenol with Cinnarnyl Alcohol ........................................................................................ 141 5.4 Conclusions ............................................................................................. 145 5.5 References ....................................................................................... . ....... 147 CHAPTER 6 INCORPORATION OF STRUCTURE ORDER INTO MESOPOROUS SILICA MOLECULAR SIEVES ASSEMBLED VIA A METAL SALT MEDIATED N°(M*X')I° PATHWAY ............... 149 Abstract ........................................................................................................................... 149 6. 1 Introduction ............................................................................................. 150 6.2 Experimental ........................................................................................... 152 6.3 Results and Discussion ............................................................................ 153 6.3.1 Materials Characterization .......................................................... 153 6.3.2 Discussion ................................................................................... 168 6.4 Conclusion ............................................................................................... 173 6.5 References .............................................................................................. 174 Table 1.1 Table 1.2 Table 1.3 Table 1.4 Table 2.1 Tahle 2.2 Table 2.3 Table 3.1 Table 4.1 Table 43 Table 5.1. T.1 table 5.3. Table 1.1 Table 1.2 Table 1.3 Table 1.4 Table 2.1 Table 2.2 Table 2.3 Table 3.1 Table 4.1 Table 4.2 Table 5.1. Table 5.2. LIST OF TABLES Catalytic Peroxidation Activity of Ti-substituted (2 mol%) Mesoporous Molecular Sieve Silica .......................................................... l4 Lactide polymerization over heterogeneous catalysts ............................... 18 Comparison of the apparent and intrinsic first-order rate constants and the over all diffusivity for NO SCR reaction over Fe3+ exchanged A1- HMS and Al-MCM-41 .............................................................................. 20 Cumene Conversions (%) at 300 0C.... ...................................................... 23 Properties of Ti-substituted Mesoporous Silicas Prepared by Ambient Temperature Synthesis Using Different Templating Pathways ................ 35 Ti K-edge XANES Results for Ti-Containing Materials .......................... 45 Catalytic Activity of Ti-Substituted (2 Mol %) Molecular Sieves Prepared by Ambient Temperature Synthesis Using Different Templating Pathways ................................................................................ 49 Physical Parameters for Calcined (650 °C) HMS Molecular Sieve Silica Prepared by 8010 Assembly in HzO-rich and EtOH-rich Medium .. 75 Physical and Chemical Properties for Al-substituted Mesoporous Molecular Sieves ..................................................................................... 102 Alkylation of 2,4-di-tert—butylphenol with Cinnarnyl Alcohol Catalyzed by Al-substituted Mesoporous Molecular Sieves ................... 105 Physical and chemical properties of Al-MCM-4l mesoporous molecular sieves ...................................................................................... 140 Alkylation of 2,4-di-tert-butylphenol with Cinnarnyl alcohol catalyzed by Al-substituted mesoporous molecular sieves ..................................... 143 Figare 1.1 Figure 1.2 figure 13 figure 1.4 Figure 1.5 Figure 1.6 figure 2.1 Figare 2) Figure 1.1 Figure 1.2 Figure 1.3 Figure 1.4 Figure 1.5 Figure 1.6 Figure 2.1 Figure 2.2 Figure 2.3 Figure 2.4 Figure 2.5 LIST OF FIGURES Schematic showing of interfacial interactions for surfactant micelles in cooperatively assembly ............................................................................... 2 X-ray diffraction patterns for mesoporous molecular sieves assembled via electrostatic (ST) and neutral (S°I°) pathways 6 N2 adsorption and desorption isotherms for mesostructured molecular sieves assembled via S+l~ and S°I° pathways ............................................... 7 TEM images showing regular order of mesopores for (a) MOM-41 and (c) MGM-48, but wormhole like channels for (b) KIT-1 and (d) HMS ........................................................................ 8 TEM image showing wormhole-like framework mesopores and meso—scaled textural pores indicated by the arrows ........... 10 N2 adsorption and desorption isotherms for HMS prepared in (A) EtOH-rich medium (H20 : EtOH = 35 : 65 by vol.); (B) HzO-rich medium (H20 : EtOH = 90 : 10 by vol.) ....................... 11 Powder XRD patterns for calcined pure and Ti-substituted (2 mol %) silica molecular sieves. (A) MCM-41 and Ti-MCM-41 prepared by S‘“X'I+ assembly and (B) MCM-4l and Ti-MCM-41 prepared by S+l~ assembly, and (C) HMS and Ti-HMS prepared by S°I° assembly ............ 37 29Si-MAS NMR spectra of as-synthesized pure and Ti-substituted (2 mol %) silica molecular sieves: (A) MCM-41 and Ti-MCM-41 prepared by STXT’ assembly, (B) MCM-41 and Ti-MCM-41 prepared by STI' assembly, and (C) HMS and Ti-HMS prepared by S°I° assembly .................................................................................................... 39 UV-VIS spectra of calcined Ti-containing materials ................................. 41 Ti K-edge XANES spectra of reference compounds containing Ti in tetrahedral and octahedral oxygen environments: (a) anatase, (b) rutile, (c) MgTiO3, (d) neptunite, and (e) spirotitanate ........................................ 44 Ti K—edge XANES spectra of calcined (a) Ti-HMS prepared by S°I° assembly, (b) Ti-MCM-4l prepared by STXT’ assembly, (c) Ti-MCM- 41 prepared by STI’ assembly, and (d) TS-l .............................................. 47 xi finn3l Epan finm33 W h d me34 T a: d p m Rants T fiflm36 X Hfimt7 gufe .11 Figure 3.1 Figure 3.2 Figure 3.3 Figure 3.4 Figure 3.5 Figure 3.6 Figure 3.7 Figure 4.1 N2 adsorption-desorption isotherms for HMS silicas assembled at ambient temperatures from dodecylamine and TEOS: Curves A and B are for derivatives obtained from a water-rich (water : ethanol = 90 : 20 (v/v)) and ethanol-rich solution (water : ethanol = 35 : 65 (v/v)); curves C and D are for derivatives prepared from the same water-rich and ethanol-rich solutions, but in the presence of mesitylene (Mes/8° = of 1.0). The HK values are the Horvath-Kawazoe pore diameters. ...... 64 N2 adsorption-desorption isotherms for HMS silicas prepared from tetradecylamine and TEOS in water-rich and ethanol-rich solution. The labeling of the isotherms is the same as in Figure 3.1 ....................... 65 N2 adsorption-desorption isotherms for a HMS silica prepared from dodecylamine in water-rich solution in the presence of mesitylene (Mes/S = 4.5): (A) after calcination at 650 °C with retention of framework and textural mesopores and (B) after calcination at 1000 °C with collapse of framework pores but with retention of textural pores. Insert: Horvath-Kawazoe pore size distribution after calcination at 650 oC ........................................................................................................ 67 TEM images of calcined (650 °C) HMS molecular sieve silicas assembled from dodecylamine : a. mesoscale fundamental particles obtained from water-rich solution; b. the spheroid fundamental particles and macroscale beads-on-a-string texture obtained from ethanol-rich solution .................................................................................. 69 TEM images of HMS silica prepared from dodecylamine in water-rich solution in the presence of mesitylene (Mes/S = 4.5) and calcined at 1000 °C to collapse the framework pore structure; (a) Low magnification image showing the retention of the textural mesopores and, (b) High magnification image showing the absence of framework pores .......................................................................................................... 71 X-ray powder diffraction pattern of calcined (650 °C) HMS silicas assembled from dodecylamine amine in water-rich and ethanol-rich solution with or without mesitylene as an auxiliary structure director ..... 72 Schematic representation of the structure-directing effects of mesitylene on HMS assembly. In the water-rich media, the mesitylene “dissolves” in the hydrophobic central core of the micelle leading to pore expansion. In the ethanol-rich environment, mesitylene preferentially “adsorbs” to interfacial head groups, thus increasing effective head group size and subsequently decreasing the pore size ....... 77 N2 adsorption and desorption isotherms for calcined (650 °C) Al-HMS molecular sieves with different textural porosity. Each sample was xii fimm42 fimm43 P35341651 Figure 4.2 Hym43 Figure 4.4 Figure 4.5 Figure 4.6 Figure 4.7 Figure 4.8 Figure 5.1 prepared at ambient temperature using dodecylamine as the structure director, TEOS as the silicon source, and using Al(sec-OBu)3 as the alumination agent, but the alumination agent was added to the reaction mixture 0, 0.33, and 20 hrs after mixing the TEOS and the surfactant. The aluminum content in each case is 2.0 mole%. The Horvath- Kawazoe (H-K) framework pore size, and the ratio of textural to framework mesoporosity (Vtx/Vfr) is provided for each sample ............... 90 N2 adsorption-desorption isotherms for calcined (650 0C) Al-HMS molecular seives prepared as described in the caption to Figure 4.1, except that the alumination agent was sodium aluminate ......................... 91 N2 adsorption-desorption isotherms for calcined (650 0C) 2% Al-HMS and calcined (540 oC) Al-MCM-41 with zero-textural porosity. The inset shows Harvoth-Kawazoe pore size distributions for these two samples. This Al-HMS was prepared using alumnum sec-butoxide as the alumination agent, which was added to the reaction mixture 20 minutes after mixing the TEOS and the surfactant solution ..................... 93 XRD patterns of calcined 2%Al-HMS molecular sieves prepared as described in the caption to Figure 4.1 ....................................................... 95 XRD patterns of 2%Al-HMS prepared as described in the caption to Figure 4.1, except that NaAlO2 was the Al-source ................................... 96 XRD patterns of calcined 2% Al-HMS and Al-MCM-41 with zero- textural porosity as described in the caption to Figure 4.3 ........................ 97 TEM images showing meso-scaled textural pores as well as worm- hole like framework meSOpores for calcined 1.5-tex-Al-HMS and 0- tex-Al-HMS ............................................................................................... 99 27Al—NMR spectra showing predominantly tetrahedral Al-sites (63 ppm) and a small amount (~10%) of octahedral Al-sites (18 ppm) in as-made Al-HMS prepared using Al(sec-OBu)3 as the alumination agent. Calcined forms of Al-HMS, as well as calcined Al-MCM-41, show 70% tetrahedral and 30% octahedral Al-sites ................................ 100 X-ray diffraction patterns for calcined (540 0C) cubic silica meso- structures formed at 150°C with a reaction composition of 1.0Si : 0.11CTAB : 0.39Na2O : 0.28H+ : 130H2O. The diffraction patterns for MCM-48 prepared by following Landry’s method as well as the diffraction patterns for a smaple prepared in the absence of both EtOH and salicylic acid are included ................................................................. 121 xiii hmm52 finn53 Hamil HpmSj Eflm5f Emits 5.: 5 Fig-lie 53 Figure 5.2 Figure 5.3 Figure 5.4 Figure 5.5 Figure 5.6 Figure 5.7 Figure 5.8 Figure 5.9 X-ray diffraction patterns for calcined (540 °C) 3-d hexagonal silica meso-structures formed under the same conditions as noted in Figure 5.1 with the promotion of citric, oxalic and tartaric acids ....................... 122 N2 adsorption and desorption isotherms for calcined (540 0C) siliceous MCM-41 prepared under the promotion of gluconic, lactic, oxalic, and citric acids with a H+INaOH ratio of 0.75 showing framework mesoporosity as well as clay—like textural porosity similar to that of montrnorillonite. The Mobil MCM-4l sample was prepared using sulfuric acid to adjust the pH (i.e. ~10.5) at the very beginning of the synthesis by control the over-all HTINaOH ratio to 0.625 ...................... 124 TEM images showing (A and B) interwoven tubular MCM-41 formed upon use of sulfuric acid; (C and D) thin sheet MCM—41 (formed upon use of citric acid) with pore channels running orthogonal to sheet surface (the thickness of the sheet estimated to be ~100nm according to the penetration depth of electron beam accelerated at 120 kV); Both samples were prepared with an over all H+lNaOH ratio of 0.75 ............ 127 X-ray diffraction patterns for MCM-41 molceular sieves as noted in Figure 5 .3 ................................................................................................. 128 N2 adsorption and desorption isotherms for MCM-41 molecular sieves prepared using tartaric acid as the structure promoter under various conditions showing the controlled formation of significant amount of HMS-like textural pores as well as clay-like textural pores. Horvath- Kawazoe pore size distribution plots are included in the inset ............... 130 X—ray powder diffraction patterns for MCM-41 molecular sieves as noted in Figure 5.6. Notably, the MCM—41 with high textural porosity exhibits well resolved (100), (110) and (200) reflections ....................... 131 TEM image of MCM-41 showing meso-scaled fundamental particles with inter particle voids as textural pores and regular hexagonal framework mesopores running orthogonal to thin sheet surfaces exemplified by a sample prepared using tartaric acid as the structure promoter .................................................................................................. 132 N2 adsorption and desorption isotherms for 8.8mol%Al-substituted MCM-41 molecular sieves (calcined at 540 0C) with clay-like and HMS-like textural porosity prepared using tartaric acid as the structure promoter and aluminum nitrate as the follow-up in situ alumination reagent in the presence of surfactant nriccelles in the pores. (The inset is the corresponding Horvath-Kawazoe plots showing sharp narrow distribution of framework mesopores) .................................................... 133 xiv fgmeSlO finnill fined] Eflmél fisne63 c Site 64 Figure 5.10 27Al—NMR spectra for as-made and calcined Al-MCM-41 molecular sieves prepared using tartaric acid as the structure promoter with HMS-like and clay-like flaky textural pores ........................................... 136 Figure 5.11 Temperature programmed desorption (TPD) of pyridine pre-adsorbed on Al-MCM-4l catalysts and purged at 150 0C in N2 stream for 1 h before the TPD started. This TPD profile was carried out on a thermal gravity analyzer ....................................................................................... 138 Figure 6.1 TEM images of calcined (600 oC) mesoporous silica molecular sieves prepared in water using the alkyl polyethylene oxide type surfactant Brij 56 (C16H33(EO)10H) as the structure director and TEOS as the silicon source in the presence of CoC12 salt as the structure promoter under the same reaction conditions (45 °C, 40 h reaction time) except that the surfactant to Co“ ratios are different. The surfactant to silicon ratio was 1 to 24 for both samples, but the surfactant to metal ratio was 8.5 for the sample denoted as MSU-C (top photo) and 2.0 for the sample denoted as MSU-H (bottom photo). These images are provided as negatives for easier viewing of the pores as black spots. Pore to pore distance is constantly ~13.7 nm through the specimen for MSU-C. The insets are the electron diffraction patterns for areas imaged. .................. 154 Figure 6.2 (a). Selected area TEM image of MSU-C; (b). Electron diffraction pattern of MSU-C; (c). Optic diffraction pattern obtained from the Fourier-transformation of image (a); ((1). Image obtained by reverse- Fourier-transformation of the center tetragonal pattern in the optic diffraction pattern (c); (e). Image obtained from the total reverse- Fourier-transformation of optic diffraction pattern (c); (f). Image obtained from the reverse-Fourier-transformation of 7 diffraction spots selected from pattern (c) in such a way that these 7 spots resembling 7 spots in the real electron diffraction pattern (b). ..................................... 155 Figure 6.3 A TEM segment for MSU-C showing a different orientation of the structure. The optic diffraction pattern (obtained by Fourier- transformation of arrow indicated area in the image) is attributable to both a (01-10)-3-d-hexagonal orientation and a (210) cubic orientation. These patterns are very similar to the diffraction pattern for meso- structured silica films with pores running orthogonal to the film surface, as reported by Brinker et a1. Notably, the pore to pore distance is ~13.7 nm. ............................................................................... 156 Figure 6.4 X-ray diffraction patterns for the calcined MSU-C and MSU-H samples that were used to obtain the TEM images shown in Figure 6.1. Both d-spacings, observed and calculated according to an uni- XV Figure 6.5 Figure 6.6 figure 6.7 Figure 6.8 Figure 6.9 81 Ol tit 311 lo; 311 dir by 111C 8111 C01 131. Diff Spa Figure 6.5 Figure 6.6 Figure 6.7 Figure 6.8 Figure 6.9 Figure 6.10. dimensional hexagonal phase for MSU-H, are provided for comparison. ............................................................................................. 159 N2 adsorption and desorption isotherms for calcined MSU-C and MSU-H. Pore necking hysterisis loop for MSU-C is indicated by the step at desorption branch, which is ~0.2 unit of relative pressure behind the adsorption branch. MSU-H exhibits almost identical adsorption and desorption branches signify the uniformity of uni- dimensional pores. ................................................................................... 160 TEM images showing the wormhole (top) and 3-d-hex-cubic (lower) meso-structures assembled at Co2+ : Brij56 molar ratios of 1:13 and l : 8.5 respectively. TEOS was the silicon source and the silicon to surfactant ratio was 17 to 1 ..................................................................... 161 X-ray diffraction pattern of the calcined (600 °C) meso-structured silica molecular sieves assembled using Brij 56 as the structure director and TEOS as the silicon source at a silicon to surfactant ratio of 15 to 1 under the same reaction conditions (45 °C, 40 h reaction time), except that the surfactant to Li+ ratio was varied as shown. ........ 163 N2 adsorption and desorption isotherms of as—made meso-structured silica molecular sieves as described in Figure 6.7. The surfactant in the mesopores was removed by ethanol extraction (i.e. refluxing in ethanol for 2 h) prior to the adsorption measurement. Again, the hexagonal silica exhibits no hysterisis loop, as expected for uniform uni- dimensional pores, whereas the wormhole silicas show pore necking hysterisis loops. ....................................................................................... 164 X-ray diffraction patterns for calcined (600 °C) mesoporous silica molecular sieves prepared using the mixed alkyl polyethylene oxide surfactant Tergitol 15-S-12 (€12-16H25-33(EO)3-10H) as the structure director and TEOS as the silicon source under the same reaction conditions (45 °C, 40 h reaction time and a surfactant to metal cation ratio of approximately 24 to 1), but different cation valences for the promotional metal salts. The X-ray patterns show a decrease of d- spacing from 6.4 to 4.65 nm upon replacing Co2+ and Mn2+ with Cr3+.. 165 N2 adsorption and desorption isotherms for mesoporous molecular sieves described in Figure 6.9. The Cr3+ mediated structure shows a pore size 2.0 nm in comparison with a pore size of 3.3 nm for the structures mediated by Co + and Mn“. ................................................... 166 xvi Figure 6.11. Figure 6.12. Scheme 1 .‘ Figure 6.11. Figure 6.12. Scheme 1 SEM (top) and TEM (bottom) images of mesoporous silica prepared using Tergitol 15-S-12 in the presence of Cr3+ as a structure promoter, showing micrometer sized intergrown spherical particles with flatten surfaces on the spheres. ........................................................................... 167 ”Si NMR spectra for as-made HMS(S°1°), MSU-H, MCM-41(S+X’I*) and MCM-41(S+I'). MSU-H was synthesized using Brij 56 as the structure director in the presence of CoCl2. HMS and MCM—415 were prepared as reported. ............................................................................... 169 Metal salt mediated formation of mesostructured silicas ................ . ....... 170 xvii SupramOI ' 1.1.1ntr0ducti A major 310-611 research-e of 11413 matert KurOda and h” quatem’an mm more \ersatile .\ surfactants 18' r a eiten ed the 5° assembly meehar X‘: Cl'. Br' and further extended substrate.4 It me3090mm silica Another 1 urOl’C‘CUial' siei'es IT : tilt the preset Chapter 1 Supramolecular Assembly of Mesoporous Molecular Sieves and Their Application in Catalysis 1.1. Introduction A major advance in the design of mesoporous molecular sieves was provided by Mobil researchers in reporting ‘ the supramolecular surfactant assembly of a broad family of M41S materials with uniform pore sizes in the mesopore range 2.0-10.0 nm. Also, Kuroda and his co-workers 2 described a structurally related mesoporous silica by quaternary ammonium surfactant rearrangement of a layered precursor (kanemite). The more versatile Mobil approach was based on the supramolecular assembly of cationic surfactants (S+) and anionic inorganic precursors (I'). Stucky and his co-workers 3 greatly extended the STI' electrostatic pathway of Mobil to include a charge-reversed S'l+ assembly mechanism, as well as counterion-mediated STX'I+ and S'M+I' pathways, where X' = Cl', Br' and M+ = Na”, KT. More recently, counterion-mediated STXT' pathway was further extended to prepare well ordered cubic mesoporous silica films on glass substrate.4 It was also extended to protonated nonionic surfactants to synthesis mesoporous silica with pore size up to 30 nm.5 Another pathway has been developed for the preparation of mesoporous molecular sieves based on the hydrolysis of an electrically neutral inorganic precursor (1°) inthe presence of a neutral amine (S°) surfactant as the predominate structure directing agent. This S°I° pathway was first used to prepare a mesoporous silica molecular sieve and a Ti-substituted analog. A small amount of protonated amine was used as aco-s protonated cu Electrostatic it forces at the ~ An equivalent. nonionic poly e: in electrically rt mesoporous Mitre 11 SCh used as a co—surfactant in the original synthesis, 6 but subsequent studies 7 showed that the protonated co-surfactant component was not needed to achieve framework assembly. Electrostatic forces do not play an important role in S°I° assembly. Instead, the assembly forces at the surfactant-inorganic precursor interfaces are based on hydrogen bonding. An equivalent H-bonding pathway, denoted N°I° has also been demonstrated for nonionic polyethylene oxide surfactants and 1° precursors. 8 The most recent development in electrically neutral assembly was a preparation of an entire family of ultrastable vesicle mesoporous Supramolecular Assembly Pathways to MeSOporous Molecular Sieves MSU ORG H-Bondlng (neutral assembly) Burpuoa 0mm Charge matchln (electrostatic pathways) Figure 1.1 Schematic showing of interfacial interactions for surfactant micelles in cooperatively assembly molecular sic 531‘ pathil}: Pathwt'S 11“] workers for the aeoordinate ct. metal center. A structures cont: schematically 11 Other 1 templating area not intrinsicall Pit-formed liq concentration I 0f the surfaces altering of the PdeucLs. Hos difflCllll)’ in I PTECUrsors 11m SUFfact Extensiyeh d a “Nantes. i n ser— 4 4C1” ' 33 ~ 1 .3th Skaters molecular sieves using neutral gemini surfactants as the structure director. 9 The S°I° and N°I° pathways are complementary to those electrostatic templating pathways. These pathways also are distinguishable from the S-I pathway developed by Ying and her co- workers for the synthesis of meso-structured transition metal oxides. 10 In the S-I pathway a coordinate covalent bond is formed between the surfactant (usually, an amine) and the metal center. All aforementioned cooperative assembly pathways afford as-made meso- structures contain liquid crystal micelles in the framework pores. All the pathways were schematically illustrated in Figure 1.1. Other than the cooperative assembly mechanisms, surfactant liquid crystal templating mechanism have been proposed too.”ll Although this templating pathway is not intrinsically different from cooperative assembly pathways, it relies very much on pre-formed liquid crystal phases to form meso—structured molecular sieves under a high concentration of surfactant. Diffusion of inorganic precursor molecules to the interfaces of the surfactant head groups and water molecules in the liquid crystal micelles without altering of the liquid crystal phase is crucial for obtaining structural order in the final products. However, the inefficient use of relatively expensive surfactants and the difficulty in maintaining the liquid crystal phases upon addition of the inorganic precursors limits the wide-spread use of this templating pathway. Surfactant-directed assembly of mesostructures by sol-gel methods has been extensively exploited. Regardless of the type of inorganic precursors used to form meso- structures, in general, the pore size and connectivity take the size and shape of the surfactant liquid crystal phase in the structural channels of as-made materials. Therefore, a good strategy for the design and synthesis of meso-structures would be the over-all control of it framework p< Ill LILlL alumina 5‘“ molecular sic metal oxides mesostructurt primary 211111 zirconia was bitWeen surt could not be SU'U'CTUTEd tu: nonionic pl' mommy}. 5 metal OXIde‘s removal of St L‘nlik meSOmeUS S has been int,- Walls are Sill toward Shirl}, limited X‘l‘ai, o“, i: control of micelle formation that determines the final liquid crystal phases in the framework pores. In addition to mesoporous silica molecular sieves, structurally stable mesoporous ”"4 many other metal oxides and metallic platinum alumina, 8'” niobia, ‘0 titania, molecular sieves ‘5 have been reported. Interestingly, most of these stable mesoporous metal oxides have made use of electrically neutral or nonionic surfactants. For example, '6 and niobium oxide 10 have been prepared using mesostructured zirconium oxide primary amine surfactants as structure directors. The surfactant in mesostructured zirconia was easily removed by ethanol extraction because of weak hydrogen bonding between surfactant molecules and framework walls, however the surfactant in niobia could not be simply extracted by ethanol because of the presence of dative bonds. Meso- structured tungsten and iron oxide, as well as other metal oxides were assembled using nonionic pluronic surfactants.l4 Nevertheless, neutral surfactants produced more structurally stable mesoporous transition metal oxides. In contrast, most mesostructured metal oxides prepared using charged surfactants as the structure directors collapsed upon removal of surfactants. ”"8 Unlike crystalline microporous zeolites, it has been recognized that the mesoporous silica molecular sieves are built of amorphous silica walls. '9 Although, effort has been invested in the synthesis of crystalline walled mesoporous molecular sieves, the walls are still amorphous for most of the mesoporous molecular sieves. An advance toward synthesis of fully crystalline mesoporous materials was claimed for materials with limited X-ray order, as indicated by limited number of reflections.14 Intrinsically, the quasi-crystalline walls of these materials is alike of a mesostructured SnO2, which is a consequence of particle assembly rather than a surfactant directed assembly.20 Actually, the assem'r relative or relatively 1 worm-hole patterns. TI ordered rr characterist neutral part 1.2. Structi thittugh Ni Four 0116 is an 0 tCTABr a, prE‘Cllfsor rI (i.e. Sufiacra 41Prepared 1‘ the Silic ”Mariano; “Orin-hater, the assembly forces at the surfactant-inorganic precursor interfaces greatly determines the relative order of meso-structures. Longer range electrostatic interaction force affords relatively higher structural order. The shorter range H-bond interaction force affords worm-hole-like structural motifs that usually exhibit single-line X-ray diffraction patterns. The key-role of longer range electrostatic interaction in the formation of higher ordered meso-structures is illustrated in the following section comparing four characteristic mesoporous silica molecular sieves assembled through electrostatic and neutral pathways. 1.2. Structural Characteristics of Mesoporous Silica Molecular Sieves Assembled through Neutral and Electrostatic Pathways Four characteristic mesoporous molecular sieves are compared below. The first one is an ordered cubic MCM-48 assembled using cetyltrimethyl ammonium bromide (CT AB) as the structure director (8*) and tetraethyl orthosilicate (TEOS) as the inorganic precursor (I) under basic conditions at relatively high concentrations of the surfactant (i.e. surfactant to silicon ratio at 0.5). 2'22 The second one is an ordered hexagonal MCM- 41 prepared using cetyltrimethyl ammonium as the structure director and sodium silicate as the silicon precursor through typical ST assembly pathway at relatively low concentration of the surfactant (i.e. surfactant to silicon ratio at 0.25). 17 The third one is a worm-hole-like HMS formed by a supramolecular assembly pathway with dodecylamine as the structure director that is predominately S°I° in character. 7 The fourth one is a disordered KIT -1 silica assembled via the 8+1" pathway using CTAB, sodium silicate and EDTA as the reagents as reported by Ryoo et al. 23 These four mesoporous molecular sieves were chosen because they are structurally distinctive. Flgllf; order reflectr. Cubic order 1‘ reflections. is, 12001 retlectii. smeared hump intrinsic shon- KTT-l. electrOs ptochSOf mtgh density caused 1 EDTA salts- conditions. 5‘ inevitably 0 electrostatic in shortened. Apparen mesoporous 1 prepared via a1 pathway exhibi mesoporosity ' y. a N2 pore-fitting Piessure 0f 0 3 isothenn for Figure 1.2 shows the X—ray diffraction patterns of these four samples. Higher order reflections for hexagonal MGM-41 and cubic MCM-48 are unambiguously shown. Cubic order is indicated by the presence of (211), (220), (321), (400), (420), and (332) reflections, whereas hexagonal order is indicated by the presence of (100), (110) and (200) reflections. In contrast, HMS and KIT -1 both show broad first reflection with smeared humps at higher 26 angles. The Lack of structural order for HMS indicates the intrinsic short- range interaction property of H-bonding in the assembly mechanism. For KIT-1, electrostatic interactions between the surfactant head group and the inorganic precursor might be weakened due to a decrease in the surfactant head group charge density caused by association with 5* = cur-tamcnypr s = CuHsNflz EDTA salts. Under these conditions, structural disorder inevitably occurs as the 51' assembled, 150 °C MGM-48 electrostatic interaction range is Intensity shortened. Apparently, those three S‘tammbled,100°c MOM-41 mesoporous molecular sieves ' . or ’ . prepared vra an electrostatic 8+1~ s “”mu‘” '1- HMS pathway exhibit only framework Set assembled. 100°C KIT -1 mesoporosity, as indicated by the 4 8 12 16 20 ZGXdeg rees) N2 pore-filling step at a relative Figure 1.2 X-ray diffraction patterns pressure of 0,2 to 0.4. (Fig. 1.3) for mesoporous molecular sieves assembled via electrostatic (ST) and The isotherm for HMS neutral (S°I°) pathways. characteristic significant N; figure l.3.ll1r and 33 A, re differences ar size for MC structurally si' 15 rported t0 1 sizes for 31C) n——‘ ‘IA'IIMA AA-__I__.I I- _ 1. Al characteristically exhibits both framework and textural mesoporosity with two significant N2 uptake steps at relative pressures of 0.3 and 0.7. As shown by the inset in Figure 1.3, the average pore sizes of HMS, MCM-41 and KIT-1 molecular sieves (29, 30, and 33 A , respectively) are significantly larger than that of MCM-48 (26 A). Structural differences are associated with differences in framework pore size. The similarity in pore size for MGM-41 and KIT-1 may indicate that MCM-41 and KIT-l tend to be structurally similar, but different from bi-continuous cubic MCM-48, even though KIT-1 is rported to possess interconnected pore channels analogous to MGM-48. Larger pore sizes for MCM-48 (Ia3d) so far have not been reported. Although the X-ray diffraction 2000‘ A a. '— (D “ m ‘ B 1500- o i A '3 a. r a) 10 20 so 40 so so 70 HMS (S "“12" '2 WAAAA“‘-- AA‘ 7 o 100 - Av!!! vV‘VIlllsV'V v v v v vv-vwwwc" .g 0 a 1"" I KIT-1 (8+1) \\\“ < r r - a w w E _ -/,,‘IV’;"'" .ooeoe e o o e o 9 990.10...” E .wc-a non-41$ r) = ' — 500 i 9’". . 'orueueueeeeee eee .. «one: to emcee O 0‘" .0. ‘s > . .- MOM-48 (3‘1) o" N «‘9‘, z ys‘ 0 "'l"'l"'l"'l"' 0 0.2 0.4 0.6 0.8 1 PIPo Figure 1.3 N2 adsorption and desorption isotherms for mesostructured molecular sieves assembled via ST and S°I° pathways. Figure 1.4 TEM images showing regular order of mesopores for (a) MCM-4l and (c) MCM-48, but wormhole like channels for (b) KIT-1 and (d) HMS. pattern does pore size dis: fundamental . TEN-l tfigrue 1.4). ' mmdmm pore channels pore channels. structural dirt} electrostatic as Structures asse limited to the . matfirials. 13' Synthesis Sufiactam as The fir SUUCIUI‘C direCI pattern does not explicitly shown higher order reflections, KIT-l exhibits a quite narrow pore size distribution which is about 3 times narrower than that of HMS. This verifies the fundamental differences in assembly mechanisms for HMS and KIT-l. TEM images show the structural versatility for these four molecular sieves. (Figure 1.4). The image for MCM-4l explicitly shows a typical hexagonal arrangement of uni-dimensional pores. The image for MCM-48 shows triangle arrangement of the pore channels running along the cubic (111) direction. Instead of very regulme arranged pore channels, HMS and KIT -1 both exhibit worm-hole-like channels. It is hard to tell the structural difference by TEM for HMS and KIT-1. Nevertheless, the long range electrostatic assembly affords meso-structures with relatively higher order than meso- structures assembled from electrically neutral surfactants. However, the order is still limited to the order of pore channels, which is far from the atomic order in crystalline materials. 1.3. Synthesis of Mesosporous Silica Molecular Sieves Using Primary Amine Surfactant as the Structure Directors The first example of HMS silica was prepared at ambient temperature in the presence of a 13.5:1 molar mixture of dodecylamine and dodecylarnmonium ion as the structure directing co-surfactants. 6 The product formed under these reaction conditions exhibited only one resolved XRD reflection, which precluded the assignment of a long range structure. Selected area electron diffraction studies provided evidence for the occasional occurrence of very small domains of hexagonal symmetry, but the vast majority of the sample was highly disordered and lacking in a long range regular SII'UCIUI'C. Si ommmgt suucture mmsue mmgam 953! O I. . ffifi’ ug-q 1“ In" 5.. .1 2’. HIV. - Subsequent studies revealed that equivalent HMS silicas could be prepared by omitting the onium ion from the reaction mixture and using only the neutral amine as the structure director. 7 This S°I° pathway afforded silicas with N2 adsorption properties, pore sizes, and XRD patterns virtually identical to the original HMS products formed using a mixture of S° and S+ surfactants. Also, the sparsely occurring small domains of Figure 1.5 TEM image showing wormhole-like framework mesopores and meso- scaled textural pores indicated by the arrows hexagonal order were absent. In fact, hexagonal regions are very rarely formed even when protonated surfactant is present. Instead, the wormhole channel motif with meso- scaled inter particle voids shown in Figure 1.5 is formed almost exclusively 24 even when up to 15% of the amine is protonated. The onium ion can be introduced by adding a protonic acid. Alternatively, the introduction of certain Lewis acid centers, as in the replacement of some Si4+ sites by Al“, Fe3+ or B“, will result in the formation of some protonated amine surfactant during the assembly process in order to balance the resulting 10 framework. Structurally mesoscalcd I frarnework. However, this small electrostatic participation of the surfactant is structurally inconsequential, and neither it alter the wormhole channel motif, nor the meso-scaled textural pores. E '— (D 0': B 3 u o .a h o m 'u < o E 2 o >“ HK=29A z .1 c I I I ' r I I 0 02 0.4 06 08 1 PIP. Figure 1.6 N; adsorption and desorption isotherms for HMS prepared in (A) EtOH- rich medium (H20 : EtOH = 35 : 65 by vol.); (B) HZO-rich medium (H20 : EtOH = 90 : 10 by vol.) Recent studies have shown that particle morphology and textural porosity of HMS can be judicially controlled by choice of solvent polarity. 23 HMS prepared in higher polarity water-rich media (i.e. 90:10 H20 : EtOH medium by volume) shows wormhole- like to sponge-like channels with fractal-like particle texture. This HMS exhibits a Significant amount of textural mesopores in the range of 150 A to 500 A, built from inter- growth of r from the V0 ethanol medi with beads 0 Therefore. re. but a propertr 1.4 Catalyst 1.4.1 Genera As pr have receireg flames-or},- p mflg Chen Complex to [h also afforded an excePliona groups of MC has been i‘IUd: thCmESOPOres Mesop Woun- A rr Mun-On 0f : en uCOUUIEr d1": 14 growth of meso-scaled fundamental particles. These textural mesopores are different from the voids formed by the physical packing of particles. HMS prepared in high ethanol medium (H20 : EtOH = 35 : 65) exhibits much larger fundamental particle size, with beads on string type of morphology and very little textural porosity. (Figure 1.6) Therefore, textural porosity is not an intrinsic property of electrically neutral assembly, but a property generated by the kinetics of the assembly process. 1.4 Catalysis 1.4.1 General Application of Mesoporous Molecular Sieves As potential sorbents, catalyst supports or catalysts, mesoporous molecular sieves have received considerable attention. A high density silanol groups on the walls of the framework provide unique capability for meso-structured silica molecular sieves for grafting chemistry. A very active Ti-MCM-41 was prepared by grafting of a titanocene complex to the meso-structured walls of MCM-4l. 25 A grafted Ru-complex on MGM-41 also afforded a super active oxidation catalyst. 26‘” Pd supported on MCM-41 generated an exceptionally active catalyst for Heck reactions. 28 Grafting chelating ligand to silanol groups of MCM—41 achieved excellent adsorbent for heavy metals in waste water. 29'” It has been studied that most of hazardous heavy metals could be remedied by trapping into the mesopores. Mesoporous molecular sieves are catalytically useful as long as their pores are uniform. A relatively high X-ray order is not necessarily beneficial for the catalytic utilization of mesoporous molecular sieves. Actually, mesoporous molecular sieves still encounter diffusion limitations, particularly for reactions in condensed media. The 12 accessibility. extremely ir The l silicas. prep. CllCllllCéll COD pOl}'mCrlZallt be discussed Prepared lhrr meSOporcs. a faulltate 3CC€ 1‘41 Peron Metal for the Pero: CPO-lilies anj accessibility to the catalytically active centers located in the mesoporous framework is extremely important for diffusion controlled reactions. The uses of chemically modified derivatives of mesoporous HMS molecular sieve silicas, prepared by neutral surfactant assembly pathways, as catalysts for a variety of chemical conversions, including peroxide oxidations of olefins and phenols, ring opening polymerization of lactide dimers, selective reduction of NOx, and cumene cracking will be discussed below. These HMS catalysts are often more active than analogous catalysts prepared through electrostatic assembly mechanisms. The wormhole-like framework mesopores, along with the presence of complementary textural mesopores, most likely facilitate access to the catalytic active centers in the framework. 1.4.2. Peroxide Oxidation Metal-substituted HMS silicas have received considerable attention as catalysts for the peroxide oxidation of aromatics to phenols and quinones and of alkenes to epoxides and diols (see equation 1-3). One of the first reactions to be investigated was OH 0 Eq. (1) O OH HO H E: + + Eq (2) —» ° / Eq. (3) 0/ 45 “c o MMA MPV l3 [he Ti-W correspond” was ofpamt‘ l. an lndusrr havebeen in phenol. mcrh cpoxide. diol : The C; phase oxlda dCSCrlbed in Obtained wit Consideratior for the mesr HUS C3121] 3. Table 1“ ‘ MOI‘~‘C\ular EWA S‘FTEne the Ti-HMS catalyzed oxidation of 2,6—di-tert-butylphenol (2,6—DTBP) to the corresponding mono-and dinuclear quinones using H202 as the oxidant. 6 This substrate was of particular interest, because it was too large to access the framework Ti sites of TS- ], an industrial microporous molecular sieve catalyst. 31 Other oxidation reactions that have been investigated using Ti-HMS as a catalyst includes the conversion of benzene to phenol, methyl metharcylate to methyl pyruvate and styrene to the corresponding epoxide, diol and benzaldehyde. 3'2 The catalytic properties of mesoporous Ti-HMS and of Ti-MCM-41 for the liquid phase oxidations of methylmethacrylate, styrene and 2,6-di-tert-butylphenol are described in Table 1.1.32 Included in the Table for comparison are the conversions obtained with microporous TS-l as the catalyst. As expected based on pore size considerations, the conversions observed for all three substrates are substantially larger for the mesoporous catalysts than for the microporous catalyst. The S°I°-assembled Ti- HMS catalyst exhibited consistently greater reactivity than the two Ti-MCM-4l catalysts Table 1.1 Catalytic Peroxidation Activity of Ti-substituted (2 mol%) Mesoporous Molecular Sieve Silica Catalyst TS-l Ti-MCM-41 Ti-MCM-41 Ti-HMS (S‘T) (S+X'I+) (S°I°) MMA Conv. (mol%) 2.5 4.0 6.2 6.8 Styrene Conv. (mol%) 8.4 10 23 28 2,6-DTBP Conv. (mol%) 5.0 39 22 55 MA is methylmethacrylate. prepared by electrostatic assembly. The superior performance of Ti-HMS is especially pronounced in the case of the large 2,6-di-tert-butylphenol substrate. 14 be anri Iiianiur coordin similuri ngl’t’gd centers. superior aiailihl. inaccess Ti-llCh 15). Th l0 the a efficienc silicas 35 Verify 1h; M0141 active Slit number 01 The differences in catalytic reactivity between Ti-HMS and Ti-MCM-4l cannot be attributed to differences in Ti siting. XANES and EXAFS studies showed that the titanium center adopt primarily tetrahedral coordination in all three catalysts. 32 Also, the coordination environment is very similar for the three catalysts, as judged from the similarities in the EXAFS features. Also, UV-VIS adsorption spectra showed no phase segregation of titania, the spectral features being consistent with site-isolated titanium centers. Because the framework walls of HMS tend to be thicker than MCM-4l, the superior reactivity of Ti-HMS cannot be due to an enhancement in the fraction of Ti available for reaction on the pore walls. Thicker walls should bury more titanium at inaccessible sites within the walls. The most distinguishing feature between Ti—HMS and Ti-MCM-41 is the greater textural (inter particle) mesoporosity for Ti-HMS (cf. Figure 1.5). This complementary textural mesoporosity facilitates substrate transport and access to the active sites in the framework-confined mesopores, thus enhancing catalytic efficiency compared to MCM-4l meso-structures with little or no textural porosity. Sayari and his co-workers ”'36 have investigated Ti- and V-substituted HMS silicas as liquid phase oxidation catalysts for large organic molecules. Their results verify that the activity of HMS derivatives is typically higher than the corresponding MCM-41 analogs. He also has emphasized the importance of the accessibility of Ti- active sites in determining reactivity toward large molecules. He further noted that a number of oxidation reactions which occured readily over small pore TS-l catalysts, did not take place in the presence of larger pore Ti-B or ultra large pore Ti-MCM-41 and Ti- HMS. All these observations suggest that differences in surface hydrophilicity and Ti redox potential also play a role in determining the reactivity of tetrahedral Ti sites in these frameworks. On the other hand, TS-l is not known to catalyze the oxidation of 15 JCCIOUC at T acetone oxi. Kilnl HUS and butylhydropt than Ti-MC.\| N prepared temperature. air were p0 subsequently l rt-stcsHi r Tuel suggest 16mm] Poms: limits] meso; N331d$0l13ti0n m1Sand HM for Meteor; Process is can texrural Porog cam-V315 can acetone at rates that are competitive with benzene hydroxylation, yet Ti-HMS catalyzes acetone oxidation readily.37 Kaliaguine and his co-workers 38 have compared the catalytic reactivity of Ti- HMS and Ti-MCM-41 silicas for the epoxidation of oe-pinene with tert- butylhydroperoxide as the oxidant. The conversions over Ti—HMS were somewhat lower than Ti-MCM-41, although epoxide selectivities were similar. Also, Gontier and Tuel 39‘” prepared a series of Ti-HMS using reaction times as short as 15 min at ambient temperature. Tetrahedral Ti loadings up to 2 wt % and thermal stabilities up to 650°C in air were possible without forming extra-framework titania. 39 However, they subsequently found no substantial difference in catalytic reactivity between Ti-HMS and Ti-MCM-4l for the oxidation of aniline. 40 These results of Kaliaguine and Gontier and Tuel suggest that the Ti-HMS derivatives used in these studies not possess the high textural porosity needed for facile access to framework Ti sites. In fact, the absence of textural mesoporosity in the Ti-HMS catalyst used by Gontier and Tuel was confirmed by N2 adsorption studies. 4' These reports verify that the interparticle mesoporosity of Ti- HMS and HMS molecular sieves, in general, is dependent on the reaction conditions used for framework assembly. In general, textural porosity is formed when the assembly process is carried out in a water-rich solvent. Alcohol-rich solvents tend to eliminate textural porosity. Enhanced reactivity for large molecule conversions over HMS catalysts can be expected only when the textural and framework mesoporosities are comparable in magnitude. 16 1.43 Ri achieved promotir the total; potential Sm IV l-Sl lactide di low polyt‘ Sr umey Swim-(:3 directing 5 indicated . a textum Ta ellllzttion .-' a SMOpg Prepared those User Tr. COm'miO] and a be 1.4.3 Ring Opening Polymerization The small crystallite domain sizes and high textural mesoporosity that can be achieved for HMS derivatives through S°I° assembly may be especially beneficial in promoting polymerizations and other bulky conversions where diffusion effects can limit the catalytic effectiveness of larger particle mesostructures. In order to demonstrate the potential utility of mesostructures for polymerization reactions, it has been shown 42 that Sn(IV)-substituted HMS is remarkably effective for the ring opening polymerization of [- lactide dimer to poly (l—lactic acid), abbreviated PLA, with a high molecular weight and low polydispersity. Sn-HMS containing 1 mol % tin(IV) was prepared at ambient temperature by S°I° assembly in ethanolzwater (3:1v/v) using a 100:1 molar mixture of Si(OC2H5)4 and Sn(iso-C3H7)4 as the inorganic precursors and dodecylamine as the structure directing surfactant. The N2 adsorption isotherm for the calcined (550 °C) mesostructure indicated a BET surface area of 886 mzlg, an average framework pore size of 2.7 nm, and a textural (inter-particle) mesoporosity in excess of the framework mesoporosity. Table 1.2 reports the conversions of L-lactide dimer to FLA at 130 °C (see equation 4). Included for comparison purposes are the conversions for pure HMS silica, a Sn-doped silica gel (1.0 mol% Sn) and pure SnOz. The latter two catalysts were prepared by hydrolysis of the corresponding alkoxides under condition analogous to those used to form Sn-HMS. The polymerization product obtained from Sn-HMS exhibited the highest conversion as determined by lHNMR (82%), the largest average molecular mass (36000) and a low polydispersity (1.1). In the case of pure tin oxide as the catalyst, the 17 conversion lower mole case of Sn-t centers in t Lewis-acid : vs l conversion was substantial (73%), but the polymerization product had a much lower molecular mass (17800) and a high polydispersity (1.7). The low activity in the case of Sn-doped silica gel may indicate the lack of a suitable dispersion of active metal centers in the host silica. Sn-HMS, however, clearly combines the reactivity of tin Lewis-acid sites with the selectivity of a regular mesopore structure in affording PLA 0 o 0 Sn-HMS II II CH3CH(OH)CO(CH(CH3)CO),,H Eq. (4) o o 130 °C Lactide Dimer PLA Table 1.2 Lactide polymerization over heterogneous catalysts“ Catalyst Conversion (%) PLA Molecular mass Sn-HMS 82 36 000 HMS 0 -- Sn-doped Silica 22 3 200 Sn02 73 17 800 “Reaction conditions: 2.00 g ( 13.9 mmol) lactide dimer; 0.1 g catalyst (except for SnOz, where 0.001 g was used); T = 130 °C; reaction time 72 h. All catalysts were calcined at 550 °C prior to use. with a reasonably high molecular mass and low polydispersity. It appears that the ordered pore structure improves average molecular mass and polydispersity values in comparison to homogeneous catalysts by imposing steric constraints on the propagating PLA chains and minimizing ‘back-biting’ and intermolecular transesterification reactions. 18 1.4.4. 5 MCM- equahc dil‘t‘ere. conten' distribi were it 0.75 c mesOpt 0f 00 ll Sufiace Fain: down “1‘ of “Jude 1.4.4. Selective Catalytic Reduction (SCR) of NO Yang et al., 43 recently investigated the activity of Fe3+ exchanged forms of A1- MCM—4l and Al-HMS for the selective catalytic reduction (SCR) of NO by NH3 (see equation 5). In order to minimize the number of factors that may cause significant differences in the catalytic properties between Al-MCM-4l and Al-HMS, the aluminum contents of both molecular sieves were controlled to around 8%. Also, the pore size distribution for the two supports was controlled to around 28 A, the BET surface areas were in the range 800-850 m2/g, and the framework pore volumes were similar (0.62 and 0.75 cm3/g for Al-HMS and A1-MCM-41, respectively). However, the textural mesoporosity of Al-HMS was comparable to the framework mesoporosity, whereas little or no textural porosity was present for Al-MCM-41. Although HMS and MCM-41 are similar both in terms of chemical composition, surface area, and pore volume, Fe/Al-HMS showed considerably higher activities than Fe/Al-MCM-4l. The main differences between HMS and MGM-41 are the crystal domain sizes and the unique textural porosity of HMS. The crystal domain size for HMS was of the order 150 A, whereas that of MCM-41 was larger by two orders of magnitude, as judged by XRD line widths and TEM. The comparison of the estimated apparent and intrinsic rate constants given in Table 1.3 show that the reaction for the Fe/Al-MCM-41 sample was severely limited by pore diffusion, but not for HMS molecular sieve. The Thiele effectiveness factor for the HMS was nearly 1.0, whereas that for the MGM-41 were 0.53 at 350 °C and 0.44 at 400 °C. An overall activation energy of 6.5 kcal/mol for Fe-HMS was in the range for SCR 19 withou attribut Thus. I Al-HM Table . the ow MCM- Catalyst \ FéfAl-l without diffusion limitation. The high effectiveness factor for the HMS catalyst is attributable to the small domain size and short diffusion path in the framework channels. Thus, the advantage for the Fe3+ exchanged mesoporous molecular sieves, in particular Al-HMS, for the SCR reaction is clearly demonstrated. NO + NH3 03‘3”“ a» N2 + H20 Est-(5) 350 °c Table 1.3 Comparison of the apparent and intrinsic first-order rate constants and the over all diffusivity for NO SCR reaction over Fe3+ exchanged Al-HMS and Al- MCM-4l'l' Catalyst Apparent k (8") Intrinsic k (s’l)£1 Diffusivity (cmzls);l 350 °C 400 °C 350 °C 400 °C 350 °C 400 °C Fe/Al-HMS 56.5 87.5 56.6 87.6 3.95x10'7 4.0x10'7 Fe/Al-MCM-41 30.0 38.5 éThe intrinsic rate constant and diffusivity are assumed to be the same for both mesoporous catalysts. 1.4.5. Acid Catalysis The replacement of silicon by trivalent elements in HMS materials has been investigated by several groups in an effort to improve Bronsted and Lewis acidity for catalytic applications. 4448 Although Al-MCM-41 derivatives can be prepared directly by electrostatic assembly pathways, the as-synthesized materials are structurally sensitive to calcination. 49 Also, if sodium silicate is used in the synthesis, a post-synthesis treatment with NH4NO3 is needed to remove residual Na+ ions from the exchange sites. 50 It has been reported by Corma and his co-workers 5' that Al-MCM-4l undergoes framework 20 dealumin. dealumin; dealumin: proton CXt In pathway molecular noted earl of structur frameworl létmhedral interaction Surfactant PWSCn'es r Tu: pathway it of Walsh incOIT‘Orat: lmponam‘l. damaging Man an dealumination upon calcination to remove the surfactant. Hitz and Prins 52 verified the dealumination of surfactant-filled Al-MCM—41 upon calcination, but they showed that dealumination could be minimized by first removing up to 73% of the surfactant by proton exchange in ethanol. In contrast to the electrostatic assembly pathways to Al-MCM-41, the S°I° pathway to Al-HMS derivatives offers a convenient route to acidic mesoporous molecular sieves with retention of framework aluminum and other trivalent ions. As noted earlier, the replacement of Si4+ by Al3+ requires the protonation of one equivalent of structure-directing amine surfactant for every equivalent of Al3+ incorporated into the framework. However, because only up to 15 mole % of the silicon can be replaced at tetrahedral sites, the majority of the framework is assembled through H bonding interactions of the amine surfactant with the silica framework. Consequently, most of the surfactant (>90%) can be efficiently removed by simple solvent extraction and this preserves the tetrahedral siting of aluminum centers in the framework. Tuel and Gontier 44 have recognized the potential importance of the S°I° assembly pathway for the preparation of HMS derivatives containing trivalent framework elements of catalytic significance. They found that A134”, Ga”, Fe3+ and 83* all could be incorporated into the framework of HMS silica at Si/M3+ ratios as low as 10. More importantly, the neutral amine surfactant could be removed by solvent extraction without damaging the framework. Also, the remaining small concentration of charge balancing primary alkylammonium ions associated with the aluminum sites could be removed either by calcination or by ion exchange without collapsing the framework or removing the aluminum from the framework. Similar results were obtained for other M3”- substituted derivatives. Thus, the structural degradation that normally occurs upon 21 calcining . through S surfactant. protonated ~ Mo's mesoporOUs pathway. L', synthesis (ti ethanol-Mate surfactant). 1 identical to 1» HMS defiml ”“3: 540 e cOml‘iifed t0 alUtninum cc h 13 unlikel} HMS, Al-Mt Walls in a“ 1 materials ma compared m The Shon int may Well rm. may be the n calcining M3+-substituted mesostructured silicas to remove surfactant can be avoided through S°I° assembly and the subsequent solvent extraction of the vast majority of the surfactant. The heat of combustion associated with the removal of the remaining protonated surfactant is not sufficient to significantly alter the framework structure. Mokaya and Jones ”'46 also have recognized the advantages of preparing mesoporous silica and aluminosilicate molecular sieves through a S°I° assembly pathway. Using a synthetic methodology equivalent to that originally reported for HMS synthesis (i.e., alkyl amines as the surfactant, alkoxides as the inorganic precursors, ethanol/water as a solvent, ambient temperature assembly, and solvent extraction of the surfactant), they prepared a series of derivatives with XRD and textural properties identical to HMS materials. However, they designated their products as MS rather than HMS derivatives. Interestingly, their Al3+-substituted products with Si/Al ratios in the range 5-40 exhibited greater Bronsted acidity and catalytic activity for cumene cracking compared to Al-MCM-41, amorphous aluminosilicate gel and zeolite HY of similar aluminum content. 47 In addition, the Al-HMS catalysts were less prone to deactivation. It is unlikely that the siting and intrinsic acidity of tetrahedral aluminum centers in Al- HMS, Al-MCM-41, and aluminosilicate gels differ significantly, because the framework walls in all three materials are amorphous. However, access to the acidic sites in these materials may differ greatly. What is particularly distinctive of most Al-HMS materials compared to Al-MCM—41 and amorphous gels is the uniformity of the framework pores. The short intersecting framework pores, in addition to the small crystallite domain size, may well facilitate access to the active sites in HMS structures. This improved access may be the main reason for the superior reactivity of Al-HMS materials as acid catalysts 22 relative to Al-MCM-4l, which have very long channel lengths, and amorphous gels with irregular, highly constrained pore-structures. Very recently, van Bekkum and his co-workers 48 have reported modifying Al- HMS and Al-MCM-41 (Si/Al E 30) to contain entrapped unit cells of ZSM-5. These modified mesostructures were prepared by ion exchanging the aluminosilicates with tetrapropylarnmonium cations as MFI structure directors and subsequently digesting the mesostructures in glycerol at 120°C for 24 h. The resulting nanoporous aluminosilicate products, designated NPA-l and NPA-2, respectively, exhibited XRD patterns Table 1.4 Cumene Conversions (%) at 300 °C Catalyst Al-MCM-4l PNA-l Al-HMS PNA-2 ZSM-S 20 min 14.7 41.3 24.8 47.6 95.1 3 hr. 13.6 37.5 26.8 42.4 93.7 characteristic of Al-MCM-41 and Al-HMS, but reflections characteristic of ZSM-5 were absent. An FI'IR band at 550-560 cm‘1 in both products was considered to be indicative of unit cells of ZSM—5 highly dispersed in the mesostructure framework. Further evidence for the presence of entrapped ZSM-S unit cells was provided by the acid catalytic activity of NPA-l and NPA-2 toward cumene cracking. As can be seen from the cumene conversions listed in Table 1.4, the acid catalytic activity of the modified mesostructures is substantially greater than the parent mesostructures. It is especially 23 noteworthy that PNA-2, obtained by modification of Al-HMS, is more active than the PNA-l product derived from Al-MCM-41. The superior activity of Al-HMS and NPA-2 relative to Al-MCM-41 and NPA-l may again be attributed to the intrinsically more efficient access to framework catalytic sites for HMS derivatives. 1.5. References l. 15. 16. Beck, J .S.; Vartuli, J .C.; Roth, W.J.; Leonowicz, M.E.; Kresge, C.T.; Schmit, K.D.; Chu, C.T.-W.; Olson, D.H.; Sheppard, E.W.; McCullen, S.B.; Higgins, J .B.; Schlenker, J .L J. Am. Chem. Soc. 1992, 114, 10834. Inagaki, S.; Fukushima, Y.; Kuroda, K. Chem. Commun. 1993, 680. Huo, Q.; Margolese, D.I.; Ciesla, U.; Feng, P.; Gier, T.E.; Sieger, P.; Leon, R.; Petroff, P.M.; Schiith, F.; Stucky, G.D. Nature 1994, 368 , 317. Lu, Y.; Ganguli, R.; Drewien, A.C.; Anderson, M.T.; Brinker, C.J.; Gong, W.; Guo, Y.; Soyez, H.; Dunn, B.; Huang, M.H.; Zink, J .1. Nature 1997, 389, 364. Zhao, D.; Feng, J.; Huo, Q.; Melosh, N.; Fredrickson, G.H.; Chmelka, B.F.; Stucky, 6D. Science 1998, 2 79, 548. Tanev, P.T.; Chibwe, M.; Pinnavaia , TJ. Nature 1994, 368, 321. Tanev, T.P.; Pinnavaia, T.J. Science 1995, 267, 865. Bagshaw, S.A.; Prouzet, E.; Pinnavaia, T.J. Science, 1995, 269, 1242. Kim, S-S.; Zhang, W.; Pinnavaia, T.J. Science 1998, 282, 1302. Antonelli, D.M.; Ying, J.Y. Angew. Chem, Int. Ed. Engl., 1996, 35, 426. Attard, G.S.; Glyde, J .C.; Goltner, C.G. Nature 1995, 378, 366. Bagshaw, S.A.; Pinnavaia, T.J. Angew. Chem. Int. Edn. Eng]. 1995, 34, 2014. Antonelli, D.M.; Ying, J.Y. Angew. Chem. Int. Edn. Engl. 1995, 34, 2014. Yang, P.; Zhao, D.; Margolese, D.I.; Chmelka, B.F.; Stucky, G.D. Nature 1998, 396, 152. Attard, G.S.; Bartlett, P.N.; Coleman, N.R.B.; Elliott, J .M.; Owen, J .R.; Wang, J .H. Science 1997, 278, 838. Huang, Y.; McCarthy, T.J.; Sachtler, W.M. Appl. Catal. A. 1996, 148, 135. 24 ii in 31 ii Huo. P.; Fi 1176. Yoda. Chen. Seven hlonn Stucltj liF.S Zhang Ryoo. Zhang Masch Kim. S Liu. 1.. hlehnc Merci: Feng, 1997, , ' Notari . Zhang Chem. Reddg Redd} Saran Rfiddt 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. Huo, Q.; Margolese, D. I.; Ciesla, U.; Demuth, D. G.; Feng, P.; Gier, T. E.; Sieger, P.; Firouzi, A.; Chmelka, B. F.; Schiith, F.; Stucky, G. D. Chem. Mater. 1994, 6, 1176. Yada, M.; Machida, M.; Kijima, T. Chem. Commun. 1996, 769. Chen, C.-Y.; Li, H.—X.; Davis, ME. Microporous Mater. 1993, 2, 17. Severin, K.G.; Abdel-Fattah, T.M.; Pinnavaia, TJ. Chem. Commun. 1998, 1471. Monnier, A.; Schuth, F.; Huo, Q.; Kumar, D.; Margolese, D.; Maxwell, R.S.; Stucky, G.D.; Krishnamurty, M.; Petroff, P.; Firouzi, A.; Janicke, M.; Chmelka, B.F. Science 1993, 261, 1299. Zhang, W.; Pinnavaia, T.J. Catal. Lett. 1996, 38, 261. Ryoo, R.; Kim, J .M.; Shin, C.H.; Lee, J.Y. J. Phy. Chem. 1996, 100, 17718. Zhang, W.; Pauly, T.R.; Pinnavaia, T.J. Chem. Mater. 1997, 9, 2491. Maschmeyer, T.; Rey, F.; Sankar, G.; Thomas, J .M. Nature 1995, 378, 159. Kim, S-S.; Zhang, W.; Pinnavaia, T.J. Catal. Lett. 1997, 43, 149. Liu, L.J.; Yu, W.Y.; Li, S.G.; Che, C.M. J. Organ. Chem. 1998, 63, 7364. Mehnert, C.P.; Ying, J.Y. Chem. Commun. 1997, 2215. Mercier, L.; Pinnavaia, TJ. Adv. Mater. 1997, 9, 500. Feng, X.; Fryxell, G.E.; Wang, L.-Q.; Kim, A.Y.; Liu, 1.; Kemner, K.M. Science 1997, 276, 923. Notari, B. Adv. Catal. 1996, 41, 253. Zhang, W.; Froba, M.; Wang,J.; Tanev, P.T.; Wong 1.; Pinnavaia, T]. J. Amer. Chem. Soc. 1996, 118, 9164. Reddy, K.M.; Moudrakovski, I.; Sayari, A. Chem. Commun. 1994, 1059. Reddy, J .S.; Sayari, A. Chem. Commun. 1995, 2231. Sayari, A. Chem. Mater. 1996, 8, 1840. Reddy, J.S.; Liu P.; Sayari, A. Appl. Catal. A 1996, 148, 7. Zhang, W.; Wang, J.; Tanev, P.T.; Pinnavaia, T]. J. Chem. Soc., Chem. Comm. 1996, 979. On, T.D.; Kapoor, M.P.; Joshi, P.N.; Bonneviot, L.; Kaliaguine, S. Catal. Lett. 1997, 44, 171. 25 39. 40. 41. 42. 43. 45. 46. 47. 48. 49. 50. 51. 52. Gontier, S.; Tuel, A. Zeolites 1995, 15, 601. Gontier, S.; Tuel, A J. Catal. 1995, 157, 124. Tuel, A.; Gontier, S. Chem. Mater. 1996, 8, 114. Abdel-Fattah, T.M.; Pinnavaia, T.J. Chem. Commun. 1996, 665. Yang, R.T.; Pinnavaia, T.J.; Li,W.; Zhang, W. J. Catal. 1997, I 72, 488. Yue, Y.; Sun, Y.; Xu, Q.; Gao, Z. Appl. Catal. A: 1998, 175, 131. Mokaya, R.; Jones, W. Chem. Commun. 1996, 981. Mokaya R.; Jones, W. Chem. Commun. 1996, 983. Mokaya, R.; Jones, W. J. Catal. 1997, 172, 211. Kloetstra, K.R.; van Bekkum, H.; Jansen, J .C. Chem. Commun.1997, 2281. Corma, A. Chem. Rev. 1997, 97, 2373. Luan, 2.; Cheng, C.F.; He, H.; Klinowski, J. J. Phys. Chem. 1995, 99, 10590. Corma, A.; Fomés, V.; Navarro M.T.; Pérez-Pariente, J. J. Catal. 1994, 148, 569. Hitz, S.; Prins, R. J. Catal. 1997, I68, 194. 26 Mt Temp Pathv Abstrat He: textural tempera incorpor mdwm anionic S‘l‘and ionic sil ms N; (Wheres degree 0 for the r isolated “Wgwj inds‘tterid Cage of GREEN [0 frames Chapter 2 Mesoporous Titanosilicate Molecular Sieves Prepared at Ambient Temperature by Electrostatic (ST, S+X'I+) and Neutral (S°I°) Assembly Pathways : A Comparison of Physical Properties and Catalytic Activity for Peroxide Oxidations Abstract Hexagonal mesoporous titanosilicates with distinguishable framework charges and textural mesoporosity, namely, Ti-MCM-4l and Ti-HMS were prepared at ambient temperature by electrostatic and neutral assembly processes, respectively. Titanium incorporation at the 2 mol % level was accompanied by increases in lattice parameters and wall thickness, but the framework pore size remained unaffected. Cross-linking of anionic framework of as-synthesized Ti-substituted MCM-41 prepared by electrostatic S+I‘ and S+X‘I+ assembly pathways (where St“ is a quaternary surfactant and I” and I+ are ionic silicon precursors) was enhanced significantly by Ti-substitution, as judged by 2‘e’Si MAS NMR. The neutral framework of as-synthesized Ti-HMS formed by $010 assembly (where S0 is a primary amine and 1° is a neutral silicon preursor) exhibited the same high degree of cross-linking as the unsubstituted silica analog. UV—VIS and XANES spectra for the calcined forms of Ti-MCM-41 and Ti-HMS indicated: (i) the presence of site- isolated Ti species in the framework; (ii) predominantly tetrahedral coordination for Ti, along with some rehydated five- and six-coordinated sites; (iii) Ti siting that was virtually independent of the framework assembly pathway. The exceptional catalytic activity in the case of Ti—HMS, especially toward larger substrates, was attributable to the small crystallite size and complementary textural mesoporosity that facilitates substrate access to framework Ti sites . 27 ll lntrt phase p hs'droxy site-isol. coordin: small p( are limi hexagon surfacta POre TS MC aSsemtil ElCCUog‘ r0d-lilte ~)- The: of [he s POSltiwef a neUlra “Cum 5 “1108111. With “Cg frameWO 2.1 Introduction Titanium silicalite-1 (denoted TS-l) is an effective active catalyst for the liquid phase peroxide oxidations of alcohols and alkanes 1, the epoxidation of alkenes 2 and the hydroxylation of aromatics 3. The activity of TS-l arises from the presence of accessible, site-isolated Ti centers in the silicalite framework that are capable of undergoing a facile coordination change and forming an active peroxo-titanium complex 4. Because of the small pore size of the inorganic framework, the substrates that can be oxidized by TS-l are limited to species having kinetic diameters < 6 A. However, the recently reported hexagonal mesoporous silica molecular sieves prepared by electrostatic 5'6 and neutral 7’8 surfactant templating pathways offer promising opportunities for the preparation of large pore TS-l analogs capable of transforming larger organic molecules. MCM-41 silicas normally are prepared by one of two possible electrostatic assembly pathways. The S+I‘ pathway originally utilized by Mobil researchersS involves electrostatic interactions and charge matching between positively charged assemblies of rod-like micelles of quaternary ammonium surfactants (8+) and anionic silicate species (I ). The second pathway, the so-called counterion mediated S+X’I+ pathway 6, makes use of the same surfactant cations under strongly acidic conditions in order to assemble a positively charged silica precursors (1*). In contrast, HMS derivatives were prepared by a neutral S°I° assembly pathway that involves hydrogen bonding interactions between neutral S° primary amine surfactants and neutral I° inorganic precursors (e.g., tetraethyl orthosilicate) 7. The electrostatic assembly pathways afford as-synthesized materials with negatively charged framework walls, whereas neutral assembly yields neutral frameworks and walls that are characteristically thicker than those formed by electrostatic 28 assembly. ' chemical prc Contra The catalyst at 408 K. 1 110141 at greater reac termed on the synthes hydrotherm concerning asStttnbleti In the HMS can compare t1 While a, 2‘2 E“(Per 12'] Matt 13-1 by the me assembly. These fundamental differences give rise to distinguishable physical and chemical properties for MGM-41 and HMS materials. Corma et al. have reported the preparation and catalytic activity of Ti-MCM—41 9. The catalyst was obtained by electrostatic S+I‘ assembly under hydrothermal conditions at 408 K. We reported the ambient temperature preparation of St‘X‘I+ - assembled Ti- MCM-4l and S°I° assembled Ti-HMS catalysts ‘0 and found preliminary evidence for greater reactivity for the Ti-HMS derivative. More recently, several other groups have reported on the preparation and catalytic properties of Ti-MCM—41. ”‘15 In all instances the syntheses were accomplished using exclusively the S+I' assembly pathway 5 under hydrothermal conditions at temperatures above 373K. Thus, relatively little is known concerning the catalytic activities of titanium-substituted mesoporous molecular sieves assembled at ambient temperature by electrostatic assembly pathways. In the present work we investigate the assembly of mesoporous Ti-MCM—41 and Ti- HMS catalysts via electrostatic and neutral assembly processes at ambient temperature, compare the physical properties of these products, and elucidate the differences in their catalytic activities for the peroxide oxidations of large organic molecules. 2.2 Experimental 2.2.1 Materials TS-l with an analytically determined titanium loading of 2 mol % was synthesized by the method of Taramasso et al. 3. Tetraethyl orthosilicate (TEOS), tetraisopropyl orthotitanate (TIPOT) and tetrapropylammonium hydroxide (TPAOH) were used as a 29 source math accon h und “Tilt 3556 solu Stirr hscl; Si: source of silica, titanium and template, respectively. The molar composition of the reaction mixture was 0.022 Ti : 1.0 Si : 0.35 TPAOH : 35 H20. The synthesis was accomplished by placing the reaction mixture in an autoclave and heating at 443 K for 48 h under static conditions. Ti-MCM-41 and Ti-HMS derivatives with equivalent titanium loadings (i.e., ~2 mol %) were prepared using deionized colloidal silica or TEOS as the source of silica, TIPOT as a titanium precursor, and long chain alkylammonium or alkylamine surfactants, respectively, as templates. In contrast to the hydrothermal reaction conditions normally used to prepare Ti-MCM-41 (i.e. autoclaving above 373 K), our Ti-MCM-4l catalysts were prepared by ambient temperature synthesis via electrostatic S+I' and S+X'I+ assembly. In both of these electrostatic pathways cetyltrimethylammonium cations (CTMA+) served as the template. For S+I' assembly, a deionized 34 wt % colloidal silica solution (Aldrich) and TIPOT were added to a solution of template under vigorous stirring and the pH of the reaction mixture was adjusted to 12 with tetramethylarnmonium hydroxide (TMAOH). The molar composition of the reaction mixture was 0.020 Ti : 1.0 Si : 0.50 CTMA+ : 1.0 TMAOH : 160 H20. The preparation of Ti-MCM-41 by the counterion-mediated pathway (S+ X'It) utilized strongly acidic conditions (pH=l.5) in order to generate and assemble positively charged silicon species from TEOS. In a typical preparation the template was dissolved in acidic aqueous solution followed by the addition of a TEOS/TIPOT solution. A higher Ti : Si gel ratio of 1:10 was used for this pathway in order to prepare a product with the desired Ti:Si ratio of 2 : 98. The use of excess Ti for this preparation was necessitated by the high solubility of Ti under strongly acidic conditions. The composition of the reaction mixture was 0.10 Ti : 1.0 Si : 0.20 30 (:1th '. ' aged unde product. Ti-ll‘l 100A) as TEOS/111 vigorous : 0.20 00.1 under vi; comparat omitting thorough? Corn 01 hexadg reference COOTdinat 2.2.2 Ch; The CTMA+ : 1.0 HCl : 160 H20. For both electrostatic pathways, the reaction mixture was aged under vigorous stirring for 24 h in order to obtain the crystalline Ti-MCM-41 product. Ti-HMS was synthesized via a neutral S°I° templating pathway using dodecylamine (DDA) as the surfactant and ethanol (EtOH) as a co-solvent. In a typical preparation the TEOS/1‘ IPOT solution was added to the solution of DDA in water and ethanol under vigorous stirring. The molar composition of the reaction mixture was 0.022 Ti : 1.0 Si : 0.20 DDA : 9.0 EtOH : 160 H20. The reaction mixture was aged at ambient temperature under vigorous stirring for 24 h in order to obtain the crystalline product. For comparative purposes we have also prepared pure silica MCM-41 and HMS samples by omitting the Ti source in the above preparations. All samples were filtered, washed thoroughly with water, dried at ambient temperature and calcined at 923 K for 4 h. Commercially available samples of MgTiO3, anatase, rutile, neptunite and a sample of hexadecaphenyl-octasiloxy-spiro-(9, 9)—titanate (denoted spirotitanate 16) were used as reference materials for XANES analysis. Spirotitanate contains Ti atoms tetrahedrally coordinated to four siloxy oxygens 17. 2.2.2 Characterization The powder X-ray diffraction (XRD) patterns of all samples were measured on a Rigaku Rotaflex diffractometer equipped with a rotating anode and Cu-Ka radiation (2:1.542 A). In general, the diffraction data were collected by using a continuous scan mode with a scan speed of 2 degrees (ZED/min. 31 state the s deter W-l retlec meat transt' Specti Libor 29Si-MAS NMR spectra were measured at 79.5 MHz on a Varian VXR-4OOS solid state NMR spectrometer equipped with a magic angle spin probe. For each measurement the samples were placed in zirconia rotors and spun at 4.2 kHz. The quantitative determination of Q2, Q3 and Q4 sites was accomplished by deconvolution of the spectra. UV—VIS spectroscopic measurements were carried out on a Shimadzu UV-3101PC UV-VIS-NIR Scanning Spectrophotometer equipped with an integrating sphere. A reflection mode with a resolution of 10 nm and BaSO4 reference were used for the measurements. The collected relative reflection intensity (R..=Rsample/Rrefe,ence) was transformed into F(R.,..) by using Kubelka-Munk function F(R..) =(1-R..)2/(2R..) 18. All spectra were plotted in terms of HR...) versus wavelengths. The X-ray absorption spectra were recorded at the Stanford Synchrotron Radiation Laboratory with a SPEAR storage ring operating at 3 GeV and 50-100 mA. Wiggler beam line 102 was used with a Si-(220) double-crystal monochromator. In order to reduce contributions from higher harmonics the second monochromator crystal was detuned to 50% of the maximum intensity. The slit before the monochromator as well as that in front of the first ion chamber was set to 1 mm height. The energy calibration of the monochromator was checked between every two spectra by measuring the Ti-K edge (4966 eV) of a Ti metal foil reference. All spectra were recorded at room temperature using a continuous—scan technique (QEXAFS) 19. The counting time per data point was 0.5 s and the distance between data points was 0.1 eV 20. The pre-edges were normalized for absorbance by fitting the spectral region from 4900 to 4950 eV with a Victoreen function and subtracting it as background absorption. In addition, all spectra were normalized for atomic absorption by using the average absorption coefficient of the 32 region from preedge pe XANES Spr symmetric p preedge an. the micro- 8 using a Ster were measu filled with a N2 ads Omnisorp 3 measuremer diSIIibutiom region from 5050 to 5150 eV. The normalization was performed in order to compare the pre-edge peak intensities, energy positions and peak widths. The normalized Ti K—edge XANES spectra in the energy range of 4960 to 4980 eV were fitted using a series of symmetric profile functions. Lorentzian and Gaussian functions were used for fitting the pre-edge and the ascending absorption edge features, respectively. The XANES data of the micro- and mesoporous titanosilicates were collected in the fluorescence-yield mode, using a Stem-Heald type detector filled with Ar gas 21, whereas all reference compounds were measured in a transmission mode using two ion chambers. The first chamber was filled with a mixture of He and N2 (ratio 2:1) and the second only with N2. N2 adsorption-desorption measurements were carried out at 77K on a Coulter Omnisorp 360CX Sorptometer using a continuous flow measurement mode. Prior measurements samples were outgassed at 423K and 10'6 Torr for 12 h. The pore size distributions were calculated by the method of Horvath-Kawazoe 22. 2.2.3 Catalytic experiments The catalytic performance of all samples was tested for the liquid phase peroxide oxidation of: (i) methyl methacrylate (MMA) to methyl pyruvate, (ii) styrene to benzaldehyde, and (iii) 2,6-di-tert-butyl phenol (2, 6-DTBP) to quinones. All reactions were performed under vigorous stirring in a three-neck glass flask equipped with condenser and thermometer. The oxidation of MMA and styrene was canied out using 10 mmol of substrate, 30 mg of catalyst, and 10 ml of acetonitrile as a solvent. The amount of 30 wt % H202, the reaction temperature, and reaction time were as follows: (i) 40 mmol H202, 321 K and 6 h, respectively for MMA oxidation, and (ii) 20 mmol 33 how: mmc prod” a F11 melt and: 2.31 H202, 321 K, and 3 h, respectively for styrene oxidation. For 2,6-DTBP oxidation, however, 10 mmol of substrate, 100 mg of catalyst, 10 ml of acetone as a solvent, and 30 mmol 30 wt % H202 were used, and the reaction was carried out at 335 K for 2 h. The products were analyzed by means of a GC equipped with a SPB-20 capillary column and a FID. The reaction products were confirmed by GC-MS analysis. The internal standard method was used for quantitative analysis of the products. The conversion of substrate and selectivities to the products were calculated through carbon balance. 2.3 Results and Discussion 2.3.1 Synthesis A major objective of this study was to elucidate the structure-reactivity relationships for mesoporous MCM-41 and HMS titanosilicates prepared at ambient temperature by electrostatic and neutral surfactant templating pathways, respectively. In order to minimize the number of compositional and structural variables for these two classes of materials and to focus on potential differences in local Ti siting by XANES spectroscopy, the ambient temperature synthesis was performed in such a way that the Ti loading for the two classes of materials was 2 mol % and the framework-confined mesopore size was 27-29 A. Cetyltrimethylarnmonium cations were used as templating surfactant for both S+I' and S+X‘I+ electrostatic preparations. The ambient temperature synthesis of Ti-HMS materials by the neutral S°I° approach generally affords materials with larger framework-confined pores than the corresponding S+I‘ or Si‘X'I+ pathways with surfactants of the same alkyl chain lengths. Thus, in order to prepare Ti-HMS with an average pore size of 27 A, we selected a primary amine 34 template witl was accompl 2.3.2 Chara (a) XR, catalysts pre Table 2.1 l Temgratu Tizf template with shorter alkyl chain, namely, dodecylamine. In all cases, template removal was accomplished by calcination in air at 923 K for 4 hr. 2.3.2 Characterization (a) XRD. Table 2.1 summarizes the properties of the Ti-MCM-41 and Ti-HMS catalysts prepared by different templating pathways. Included in the Table for Table 2.1 Properties of Ti-substituted Mesoporous Silicas Prepared by Ambient Temrature Synthesis UsingDifferent Templating Pathways Parametera MCM-41 MGM-41 HMS (s+I-) (s+x-I+) (S°I°) Surfactant C16H33N(CH3)3+ C16H33N(CH3)3+ C12H25NH2 Ti:Si ratio (mol): Initial gel 2.0:98 10:90 22:97.8 Calcined product 22:97.8 25:97.5 2.4:97.6 d100 (A) 38.1 (36.0)b 36.5 (33.0) 40.2 (36.0) Unit cell (A) 44.0 (41.5) 42.5 (38.1) 46.4 (41.5) A (A) 4.1 (8.5) 5.6 (9.6) 0 (3.0) H-K pore size (A) 29 (28) 27 (26) 27 (26) FWT (A) 15 (12) 15 (12) 19 (15) SBET (m2/g) 859 (923) 1354 (1345) 1075 (1108) Vtotal (cm3/g) 0.70 (0.72) 0.92 (0.95) 1.40 (1.42) Vertcm3/g) 0.68 (0.70) 0.90 (0.92) 0.68 (0.70) Vtx (cm3/g) 0.02 (0.02) 0.02 (0.03) 0.72 (0.72) Vth fr 0.03 (0.03) 0.02 (0.03) 1.06 (1.03) a The unit cell parameter = 2d100\/3 and A is the unit cell contraction upon calcination. FWT is the framework wall thickness obtained by subtracting the Horvath-Kawazoe (H- K) pore size from the unit cell parameter. The total liquid pore volume Vtotal was estimated at a relative pressure of 0.95. The volume of framework confined mesopores, V2,, was determined from the upper inflection point of the corresponding adsorption step. The volume for textural porosity, V“, was obtained from the equation Vt, = Via... - V2,. The data in parenthesis are for the pure silica analogs. 35 comparison art parentheses). essentially com; the S‘X‘l+ path achieve a simil. (based on Si) fr 5*X'l” pathway All reactio but Ti-MCNH Pflfameters (cf. Sarnpleg pIEpm could be 3mm framework Sylz henna] [6mm flalnewmk ‘ mcSOPOFOUS my electrostatic [er Surfactant-silk; interacrjOHS in harnewOrk “Q1 framewom Th Stto . my) Suggegr comparison are the structural parameters for the pristine silicates (the values in parentheses). It should be noted that both S+I' and S°I° pathways, allow for the essentially complete incorporation of the Ti in the product at the 2 mol % level, whereas the S+X'I+ pathway required a 4-fold excess of Ti in the reaction mixture in order to achieve a similar Ti loading (2.5 mol%). In addition, the yields of crystalline product (based on Si) for the S+I' and S° I° pathways were more than 85%, whereas that for the S+X‘I+ pathway was only about 50%. All reaction products exhibit similar d100, unit cell parameters and H-K pore sizes but Ti—MCM-4l and Ti-HMS catalysts differ dramatically by their lattice contraction parameters (cf. Table 2.1). The significant lattice contraction exhibited by Ti-MCM-41' samples prepared by the electrostatic S+I' and St‘X'I+ pathways is not surprising and could be attributed to the cross-linking of the significant amount of non-fully condensed framework sylanol groups upon calcination (see below). In contrast Ti-HMS prepared by neutral templating does not exhibit lattice contraction due to its more fully cross-linked framework. We have previously noted 7’23, that the neutral S°I° pathway affords mesoporous materials with more fully cross-linked and thicker framework walls than the electrostatic templating pathways. We have attributed this phenomenon to the absence of surfactant-silica oligomer charge matching and silica-silica oligomer charge repulsive interactions in forming the framework walls. As shown by the data in Table 2.1 the framework wall thickness of all materials increases upon Ti incorporation into the framework. The changes in unit cell size and framework wall thickness (cf. Table 2.1) strongly suggest Ti incorporation in the walls of the mesoporous silicate framework. An 36 F: 76 ”\k T‘ \ \ 35.62. .5 .3 83%... 34...: Ba 9.: Q .Us. 53888 km E Began. .1624... Ba .36: 3. Ba 32:88 .Cbm E 882.. $202-: can _YEUE O3 3.63 3.329: 85m $5 BE 3 @2332??? new 2.5 3523 com mEozam max Swazi mm 23mm— Ameohmouv ®~ 3 or m a on my 3 m a on m. or m o a»..F.P_~—Pphhhblppp upphp-h.—--p_b»». LPL.~l»»-p_- mi: .1205. 51.20.... mixfi 5.5.2....» piano—2-... j Mtsueim 0— 0m u— +m +— Ix +m 37 increa incorp substit :3‘ (b isot r produc larger t increase in the unit cell size parameter has been also noted 24 to occur upon Ti incorporation into ZSM-S frameworks. It is noteworthy that the specific surface areas are very similar for the pure and Ti- substituted samples. On the basis of comparisons of the full N2 adsorption-desorption isotherms (not shown), the most significant difference between the Ti-substituted products prepared by neutral and electrostatic templating pathways is in the substantially larger textural mesoporosity exhibited by the S°I° assembled Ti-HMS derivative (compare Vtx and Vm/Vfr ratios in Table 2.1). Ti-HMS exhibits ratio of Vtx/Vfr mesoporosity of 1.06 whereas Ti-MCM-41 samples exhibit negligible small ratios in the range of 0.02- 0.03. The enhanced interparticle mesoporosity of Ti-HMS, which is a consequence of the smaller particle size afforded by neutral surfactant templating, will be shown later to play an important role in liquid phase catalytic oxidation reactions of large organic molecules. Figure 2.1 illustrates the XRD patterns of the calcined Ti-MCM-41 and Ti-HMS materials together with those for the corresponding pure silica analogs. All materials exhibit well-defined 100 reflections in their XRD patterns. Hexagonal mesostructures with greater long range order were obtained at ambient temperature conditions by the S+X'I+ pathway (Fig. 2.1 A). This is evidenced by the presence of additional relatively narrow 110 and 200 diffraction lines in the XRD patterns of pure and Ti-substituted MCM-4l samples. The diffraction patterns of the calcined Ti—MCM-4l and pure silica analog prepared by S+I' pathway (Fig. 2.1 B) are very similar to these of HMS and Ti- HMS (Fig. 2.1 C). These patterns generally exhibit 100 reflections accompanied with broader unresolved higher order reflections. These results are not surprising in view of the ambient temperature conditions used for the synthesis and the lack of NaOH in our 38 .5583 arm 3 83%... 9.5... Ba m2: Q Ea $3888 .5 3 Base $202-? .85 .152 .6 $3588 Form 3 883:. $202-: .05 .162 A3 .8508 3:60—08 3:3 36 SE S coda—5.0.98-3 can 23 wonmmofiimbm mo 880% £22 wee/758 End co..- cc 7 ow 7 oo ..- co. co. co w- 03. cu w- 2:. cm. om. cow. 2;. cu... cc 7 an. on. . . . . _ . . . b _ _ . . _ p . —hhblhbphl-Ib_blb PbPhPhh_Pbl b Pth 29.362 29.5»: 295002 macaw P-nb—th—bDb—b-h—hbn-Ph .....m .o -..m .m +can d 39 Stl' re silica (BSClll bond: band S+I' reaction mixtures. We observed similar lack of long range order for a series of pure silica MCM-41 materials prepared by ambient temperature synthesis by the S+I' assembly pathway with surfactants of different alkyl chain lengths 23. The lack of long range order exhibited by the HMS and Ti-HMS samples was attributed to the weak H bonding forces that govern the neutral S°I° assembly process and to the corresponding small scattering domain sizes 7. (b) 29Si MAS NMR Figure 2.2 shows the 2981 MAS NMR spectra for as-synthesized samples of pure and Ti-substituted mesoporous molecular sieves prepared by ambient temperature S+I‘, Se‘X‘I+ and S°I° assembly. In general, three bands centered at chemical shifts of —92, -100, and -110 ppm were observed for the MCM-4l derivatives. These bands can be attributed to Si(OSi),((OH)4.x framework units where x=2 (Q2), x=3 (Q3) and x=4 (Q4), respectively. It is noteworthy that all as-synthesized electrostatically- templated pure and Ti-substituted MCM-41 samples exhibit a higher fraction of incompletely cross-linked Q2 and Q3 framework units than the S°I° HMS samples (see Figure 2.2). In addition, the as-synthesized pure and Ti-substituted HMS samples exhibit a much higher average ratio of Q4/(Q3+Q2) units (~2.8) than the as-synthesized pure and Ti-substituted MCM-4l samples prepared by electrostatic templating (~1.0). This implies that neutral templating allows for the preparation of mesoporous molecular sieves with more completely cross-linked frameworks. A noteworthy trend is observed by comparing the spectra of pure and Ti-substituted MGM-41 samples prepared by both electrostatic templating pathways. Significantly, Ti incorporation leads to a dramatic increase of the cross-linking of the MCM-41 framework. This is evidenced by the increase in fraction of Q4 sites at the expense of both Q2 and Q3 framework sites. In the 40 - y...— case of 1015 significant (1 framework. An add conespondi case of HMS, the increase of the fraction of Q4 sites upon Ti incorporation is not as significant due to the much greater degree of cross- linking exhibited by the pure silica framework. An additional consequence of Ti framework incorporation is the broadening of the corresponding Q2, Q3 and Q4 peaks in the 29Si MAS NMR spectra. This broadening 210 e 325 a. HMS, S°I° f b.Ti-MCM-41, S’X'I” c. Ti-HMS, 8° 1“ d. Ti-MCM-41, S t l' e. Anatase f.TS-1 260 I d 8 G u. S ' I c | b I I a N 200 250 300 350 400 450 500 Wavelength (nm) Figure 2.3 UV-VIS spectra of calcined Ti-containing materials. 41 could be attributed to the effect of the Ti sites on the chemical environment of the adjacent Si atoms. (c) UV-VIS Diffuse Reflectance Spectroscopy. The incorporation of titanium into MCM-41 and HMS frameworks was further verified by UV-VIS diffuse reflectance spectroscopy. The corresponding spectra of the Ti-substituted samples are shown in Figure 2.3, along with those for TS-l, HMS and anatase. The spectrum for TS-l shows an absorption band at 210 nm, whereas bulk titania (anatase) shows a band at 325 nm. The band at 210 nm was attributed to ligand-to-metal charge transfer associated with isolated Ti (IV) framework sites in tetrahedral coordination 25. The broad absorption band centered at 325 nm is typical for ligand-to-metal charge transfer occurring in bulk titania. The spectra for the mesoporous Ti-MCM-41 and Ti-HMS derivatives lack the 330 nm band characteristic for segregated titania. This suggests that most of the Ti atoms in our Ti-MCM-41 and Ti-HMS samples occupy site-isolated positions in the silica framework. The spectra for Ti-MCM-41 and Ti-HMS also are clearly different from the spectrum of the microporous TS-l. This is manifested by the much broader character of the absorption bands centered at 220 and 260-270 nm. The possibility of some Ti-O-Ti clustering in the framework cannot be unequivocally precluded as amorphous TiO2/SiO2 gels with various Ti content exhibit absorption bands in the intermediate range of 250- 330 nm 4. However, the 220 nm band clearly signifies that much of the Ti is in site- isolated form. The slight shift and the increase in the width of the 220 nm band may be indicative of a Ti in a distorted tetrahedral environment or of the presence of Ti species in 13.14 an octahydral coordination sphere. However, since there is very little shifting of the 42 band toward lower wavelengths upon thermal dehydration, we favor a distorted tetrahydral environment for most of the Ti sites in Tti-MCM-41 and Ti-HMS. The XANES data provided below verify this assignment. We believe that this distorted environment is a direct consequence of the amorphous character of the pore walls (i.e., a wide variety of Ti-O-Si bond angles). According to several literature studies, 27 the absorption shoulder at 260-270 nm can be attributed to the presence of site-isolated Ti atoms in penta- or octahedral coordination. Similar UV-VIS behavior, indicative of a high fraction of higher coordinated Ti sites, also has been reported to occur upon hydration of TS-l 27. The presence of a significant fraction of Ti sites with coordination numbers higher than four in Ti-MCM-41 and Ti-HMS may be associated with the less crystallographic order in the pore walls, the much higher accessible surface area, and the larger mesopores. These later factors, most likely are responsible for the observed enhanced hydration of Ti sites in MGM-41 and HMS relative to silicalite-1. It is important to note that the UV-VIS spectra of the electrostatically templated Ti- MCM-41 and Ti-HMS prepared by neutral templating are very similar and practically indistinguishable. This implies that the effect of the templating method on the Ti siting in Ti-MCM-41 and Ti-HMS samples prepared under ambient temperature conditions is negligible. Therefore, any differences in the catalytic behavior of these materials may not be attributed to intrinsic differences in Ti-siting. (d) X -Ray absorption. In order to better elucidate the nature of the Ti-sites in mesoporous Ti-MCM—41 and Ti-HMS samples we have performed Ti K-edge XANES measurements 21. For comparison purposes we have also obtained Ti K-edge 43 annualized Absorption l 1 l 1 I l 4.96 5.0 5.04 E (keV) Figure 2.4 Ti K-edge XANES spectra of reference compounds containing Ti in tetrahedral and octahedral oxygen environments: (a) anatase, (b) rutile, (c) MgTiO3, (d) neptunite, and (e) spirotitanate. 44 Spj Table 2.2 Ti K-edg XANES Results for Ti-Containifl Materials ' Sample position normalized FWHMb (eV) height (eV) Spirotitanate 4970.2 0.95 0.87 rutile 4971.7 0.20 l .5 anatase 4972.1 0.15 1.7 M8Ti03 4971.5 0.27 0.8b neptunite 4971 .6 0.33 l .2 TS-l 4970.5 0.52 1.0 (2 % Ti) Ti-MCM-4l 4970.7 0.30 1.3 (S+I', 2 % Ti) Ti-MCM-4l 4970.7 0.3 1 1.3 (S+X'I+,2.5 % Ti) Ti-HMS 4970.8 0.31 1.3 (S°I°, 2.4 % Ti) “The Ti K-edge XANES parameters were obtained by fitting the central pre-edge peak to a Lorentzian Function. bl'he full width at half maximum, FWHM, is for the central pre-edge peak. XANES spectra of reference compounds such as octahedrally coordinated anatase, rutile, MgTiO3, distorted octahedrally coordinated neptunite, tetrahedrally coordinated spirotitanate and TS-l. Figure 2.4 shows the Ti K-edge XANES spectra of the reference compounds. All reference compounds with nearly octahedral Ti site symmetry exhibit multiple, low intensity pre-edge peaks in the region from 4960 to 4980 eV (cf., spectra a, b). Distortion of the octahedral geometry (inversion symmetry) leads to an increase in intensity of the central peak, as shown for neptunite 28 (spectra c and d). In contrast, the spirotitanate with nearly regular tetrahedral geometry exhibits a single, very high 45 intensity pre-edge peak (spectrum e). These results (see Figure 2.4 and Table 2.2) are in good agreement with previous studies. 29'” We may conclude, therefore, that tetrahedral Ti sites give a single, high intensity pre-edge peak, whereas Ti-sites in regular octahedral symmetry afford multiple, low intensity pre-edge peaks. Although tetrahedral and highly distorted octahedral Ti sites exhibit similar single pre—edge peak, the two environments clearly are distinguishable by means of peak intensity and position. The Ti K-edge XANES spectra of the mesoporous Ti-MCM-41 and Ti-HMS samples prepared by electrostatic and neutral templating methods, respectively, are shown in Figure 2.5. Included for comparison is the spectrum for microporous TS-l. The following features are evident: (i) the spectra for all three samples are very similar and contain a sharp pre-edge peak, (ii) the energy position of the pre-edge peak for Ti- substituted MCM-41 and HMS is similar to that for TS-l and spirotitanate, but different from the reference materials containing octahedral Ti (see Table 2.2), (iii) the two Ti- MCM-41 samples and Ti-HMS exhibit weaker pre-edge peaks that are slightly shifted toward higher energies and wider FWHM than those of TS-l and spirotitanate, suggesting in addition to tetrahedral Ti the probable presence of some Ti higher coordination sites (see below); (iv) most significantly, the spectral parameters for calcined Ti-substituted MCM-41 and HMS molecular sieves prepared via different assembly pathways at ambient temperature are very similar and practically indistinguishable. These observations are also supported by our UV-VIS results and show that the Ti siting in mesoporous MCM-41 and HMS materials is independent of the templating pathway. 46 normalized Absorption fl r l r r L l I l l 4.96 5.0 E (keV) Figure 2.5 Ti K-edge XANES spectra of calcined (a) Ti-HMS prepared by S°I° assembly, (b) Ti-MCM-4l prepared by S“X‘I+ assembly, (c) Ti-MCM- 41 prepared by S‘T assembly, and (d)TS-1. 47 Extending somewhat further our interpretation of the XANES data, we note that Farges et al. performed ab initio multiple-scattering calculations of the Ti K-edge XANES spectra for mixtures of reference compounds containing Ti atoms with various coordination. 33 For instance, they calculated that the normalized height of the central pre-edge peak varies from 0.6 to 1.0 for 4~fold Ti, from 0.4 to 0.7 for 5-fold Ti and from 0.05 to 0.27 for 6-fold Ti containing compounds. Besides the differences in intensity, the 4- and 6~fold Ti containing compounds were found to differ by energy position of the pre-edge peak. The energy position of the pre-edge peak for 6cfold Ti containing compounds was found to be shifted on average of 1.5 eV toward higher energies. 33 Thus, structures containing mixtures of 4-fold Ti with 5-fold or 6-fold coordinated Ti sites should exhibit pre-edge peaks with lower intensity that are slightly shifted toward higher energies. Therefore, the coordination state of Ti-sites in our Ti-MCM-41 and Ti- HMS materials may well be a mixture of 4-, 5- and 6-fold coordinations with tetrahedral coordination being in the majority. The higher coordination Ti-sites, are most likely generated through hydration of the tetrahedrally coordinated sites. 2.3.3 Catalytic results Corma et al. have demonstrated 9 that Ti-MCM-41 catalyzes the epoxidation of rather small organic molecules such as hex-l-ene and norbomene. We have reported that Ti-MCM-41 and Ti-HMS are very effective oxidation catalysts for bulky aromatic molecules that can not be converted over TS-l, such as 2, 6-DTBP 10. More recently, several groups have elaborated on the preparation, characterization and catalytic activity of Ti-MCM-41 ”‘15 and Ti-HMS 534. All of the Ti-MCM-4l catalysts investigated to date were prepared by prolonged hydrothermal synthesis above 373 K using the 48 A. N 8.. Mam .m masses 8 2.8038 .80.. a... 8... ON: 2. 0: 2 «OS: 68 .3 0m 8...... cm .3388 m... cm .8 8.8.88 80.23. 0.2.8.. 8.8.8. $22.... a 8 .8288... 982:... .5520: Z .6 .23 .8382. .8 82.2.8.2. a...» 82892808 «02.: 8 882.5 88.2.. 2... 0.82.8 8:88... 8 323828 03.22.88 8 23.8.98 83 0.3828 08:80 9 8.93.3 13.0... 2 >22 a m. m... am 8 o 3 o o 68.22 new. 3 .8 mm 2. S. P ma 8 we 22..-... 5-8.9 .e om mm a... .2. MK 8 8 me .8202... A-~+mv mm a a... .3 me a. o. 8 o... .5202... E m n... v. E 4... .K on .22 as .2... as .25 as .2... .28 .2... .28 8.9.. as .25 .2. .2... 28.9.. .628 as .2... 8 >2). .2. .2... 82.28 3:5 0.>:00 0:0:30 .>:00 0028 BR— ofimxomm OIUSQ .>:00 goo—om .>:00 .9882. NONI 38282.8 mmhabd 82.2.8 0.825 82.28 <22 .9335...— wfiaaifioh 0:88.50 mama: 28.—«Fa muggomfioh 2.2.1.2. .3 2.8.8...— 8>2m 81.882 88 82 N0 08353.6-3 .8 .3233. 23.380 QN «Sun. 49 electrostatic S+I‘ pathway originally demonstrated by the pioneering work of Mobil 5. Here we compare the catalytic activity of Ti-MCM-4l and Ti-HMS samples prepared by ambient temperature synthesis using both S+I' and S+X'I+ electrostatic templating pathways and the neutral S°I° templating pathway, respectively. The catalytic performance of the samples was tested toward olefin and aromatic substrates of different size, such as the relatively small methyl methacrylate (MMA), styrene and the bulky 2,6- di-tert-butyl phenol (2,6-DTBP). The catalytic results are summarized in Table 2.3. The mesoporous Ti-MCM-41 and Ti-HMS exhibit higher catalytic activity than microporous TS—l for all reactions. The difference in catalytic activity between the mesoporous and microporous titanosilicates increases with increasing of the substrate size. The difference in catalytic activity is very small for the relatively small and elongated MMA molecule (2-2.5 times higher conversion for the mesoporous catalysts). However, as the substrate molecule becomes larger and more bulky (styrene and 2,6- DTBP) the difference becomes much more pronounced. Thus, the difference in catalytic activity among Ti-MCM-41, Ti-HMS and TS-l reaches a maximum in the case of 2,6- DTBP (4 - 11 times higher conversion). This result is not surprising given the fact that the Ti-active centers are much more accessible in the larger mesopore size frameworks of Ti-MCM-4l and Ti-HMS. Due to the small micropore size of TS-l (~6 A), the large 2,6- DTBP substrate can not penetrate into the framework pores and therefore can not be transformed (only 4.5 mol % conversion). On the other hand, anatase (bulk titania) was completely inactive for MMA and styrene oxidation. Its catalytic activity for the conversion of the large 2,6-DTBP is also negligible (only 4.5 mol %). 50 The oxidation of MMA proceeds with higher selectivity to methyl pyruvate (MPV) over Ti-MCM-41 and Ti-HMS, but the microporous TS-l affords slightly higher selectivity to epoxide than Ti-MCM-41 and Ti-HMS (epoxide selectivity not included in Table 2.3). The tendency of TS—l to oxidize olefins primarily to epoxides is well documented 35. A similar trend also is observed for styrene oxidation. Generally, Ti- MCM-4l and Ti-HMS afford higher selectivity to diol, whereas more epoxide is produced over the microporous TS-l. The fact that the oxidation of MMA and styrene over Ti-MCM-41 and Ti-HMS proceeds selectively toward the alcohol derivative (through the epoxide intermediate) could be a consequence of the more hydrophilic nature of the amorphous pore walls. Notari et al. have reported that bulk titania or substituted silicas containing Ti-O-Ti bonds are not suitable for catalytic peroxide oxidation reactions because they selectively decompose H202.4 The comparison of the results for the H202 decomposition over our samples reveals a very interesting trend (Table 2.3, last column). In accordance with Notari et al. our anatase sample affords 13 mol % decomposition of H202. Both TS-l and Ti-HMS exhibit lower H202 peroxide decomposition activity, whereas Ti-MCM-41 samples prepared by electrostatic 8+1' and S+X'I+ templating methods exhibit a higher tendency to decompose H202. Especially noteworthy is the higher degree of H202 decomposition by Ti-MCM-41 prepared by the acidic counterion-mediated pathway S+X' 1+. This result is probably related to the pronounced tendency of the acidic S+X'I+ pathway to impede incorporation of Ti atoms in the silicate framework. All of our syntheses involving the counterion-mediated pathway required at least a two to four-fold excess of Ti in the initial gel in order to achieve the desired Ti doping. It may be that the 51 strongly acid mesoporous A furth exhibits acti samples for. pronounced 2.1. Ti-MC little or n characteris meSOporos usually eq These dif‘ molecular szzs d. liquid pha meSOPOIOS large (3er 2'4 Conclu Hex;lg Prepared a: approac h Es {Elma} S°I° strongly acidic conditions used in the synthesis limit the incorporation of Ti atoms in the mesoporous framework. A further comparison of the catalytic data reveals that S°I° -assembled Ti-HMS exhibits activity superior to the electrostatically S+I' and S+X'I+-assembled Ti-MCM-41 samples for all reaction systems. This enhanced activity for Ti-HMS is especially pronounced in the case of the large 2,6-DTBP. As shown by the data presented in Table 2.1, Ti-MCM-41 samples possess predominantly framework-confined mesoporosity and little or no significant textural mesoporosity. 10 In contrast, Ti-HMS samples characteristically exhibit significant complementary textural or interparticle mesoporosity. The ratio of textural to framework-confined mesoporosity of Ti-HMS is usually equal or higher than 1, whereas that of Ti-MCM-4l is usually close to zero. These differences in catalytic behavior of Ti-HMS and Ti-MCM-41 mesoporous molecular sieves are not due to differences in Ti siting, as judged from UV-VIS and XANES data. We conclude, therefore, that the superior catalytic activity of Ti-HMS in liquid phase oxidations is most likely due to the presence of complementary textural mesoporosity that facilitates access of the framework-confined mesopores, especially by large organic substrates. 2.4 Conclusions Hexagonal mesoporous titanosilicates containing 2 mol % Ti - substitution can be prepared at ambient temperature synthesis via three different surfactant templating approaches (S+I‘, S+X‘I+, and S°I°). Electrostatic S+I' assembly of Ti-MCM-41 and neutral S°I° assembly of Ti-HMS allow for almost complete incorporation of Ti in the 52 mesoporous framework (at 2 mol % doping level), whereas electrostatic assembly of Ti- MCM-41 under acidic conditions via a counterion-mediated S+X‘I+ pathway impedes Ti incorporation. The incorporation of Ti in the anionic framework of MCM-4l materials and in the neutral framework of Ti-HMS mesopore frameworks is accompanied by an increase of the lattice parameter and the framework wall thickness. The framework mesopore size of the MCM-4l and HMS samples remains unaffected by the framework incorporation of Ti atoms. The degree of framework cross-linking of the electrostatic S+I' and S+X'I+-templated MGM-41 samples is significantly improved by Ti-substitution as judged from the 29Si MAS NMR data. The UV-VIS and XANES results for calcined forms of Ti-MCM-41 and Ti-HMS provide evidence that: (i) Ti species are incorporated and site-isolated into the framework structure; (ii) Ti atoms are predominantly in a tetrahedral coordination, but there is the possibility of some Ti sites in five- or six- coordinated sites . (Higher coordinated Ti-sites most likely are generated by rehydration of some of the tetrahedrally coordinated sites), and (iii) the Ti siting for the S+I' and S+X' 1+ Ti-MCM-41 and for S°I° Ti—HMS are practically indistinguishable. This implies that the assembly pathway method has little or no effect on the Ti siting, at least when the Ti- MCM—4l and Ti-HMS samples are prepared at ambient temperature conditions. A comparison of the catalytic results for the liquid phase peroxide oxidations of methyl methacrylate (MMA), styrene and 2,6—di-tert-butylphenol (2,6-DTBP) reveals that Ti-MCM-4l and Ti-HMS exhibit catalytic activities higher than TS-l for all reactions. The difference in activity between the mesoporous molecular sieve catalysts and microporous TS-l increases with increasing substrate size. The difference in activity is small for the relatively small MMA molecule, but as the substrate becomes larger, as 53 in styrene a pronounced. centers emb Due to the . pnettnte int li-lhlS ex electrostatn especially l lit-NBS do be attnbut thicker fra. sites Can it Stmctural mesoporo facilitates e“harming 2'5 Refer in styrene and, especially 2,6-DTBP, the difference in activity becomes much more pronounced. This result is not surprising given the much more accessible Ti-active centers embedded in the larger mesopore size frameworks of Ti-MCM-4l and Ti-HMS. Due to the small micropore size of TS-l (~6 A) the large 2,6-DTBP substrate can not penetrate into the framework pores and therefore can not be transformed. S°I° -assemb1ed Ti-HMS exhibits greater catalytic activity than the Ti-MCM-41 materials assembled by electrostatic S+I‘ and S+X'I+ pathways The superior performance of Ti-HMS is especially pronounced in the case of the large 2,6-DTBP. On the basis of UV-VIS and XANES data, these differences in catalytic behavior of Ti-HMS and Ti-MCM-41 cannot be attributed to differences in Ti coordination environment. Bercause Ti-I-IMS has thicker framework walls than Ti-MCM-4l, an enhancement in the fraction of surface Ti sites can not be responsible for the greater activity of Ti-HMS. The most distinguishing structural feature between Ti-HMS and Ti-MCM-41 is the greater inter-particle (textural) mesoporosity for Ti-HMS. These complementary textural mesoporosity most likely facilitates substrate transport and access of the framework-confined mesopores, thus enhancing the catalystic efficiency of Ti-HMS. 2.5 References 1. a. Huybrechts, D. R. C.; De Bruycker, L.; Jacobs, P. A. Nature 1990, 345, 240; b. Esposito, A.; Neri, C. ; Buonomo, F. US Patent 1984 , 4,480,135. Clerici, M. G.; Romano, U. US Patent 1990, 4,937,216. Taramasso, M.; Perego, G.; Notari, B. US Patent 1983, 4,410,501. Notari, B. Stud. Surf. Sci Catal. 1988, 37, 413. a. Kresge, C. T.; Leonowicz, M. E.; Roth W. J .; Vartuli, J. C.; Beck, J. S. Nature 1992, 359, 710; (b) Beck, J. S.; Vartuli, J. C.; Roth, W. J.; Leonowicz, M. E.; 9‘5“?!” 54 Margo Firouz: l 176. 7. Tanev, 8. aAttar. S. A.; P 9. Corma. 1994.14 10a Time Pinnava Materiaj Eds.).11 11.Franke‘ Related Weitkan Elsevjer 1241 Sanka DentA. F.;SankE 1331mm,) 14, Remy. 1 Synthesis PTCSS. 15. 00mm! 3 16' Zeitler, V_ 17. HUtsyhOuSe Kresge, C.T.; Schmitt, K. D.; Chu, C. T-W.; Olson, D. H.; Sheppard, E. W.; McCullen, E. B.; Higgins, J. B.; Schlenker, J. L. J. Am. Chem. Soc. 1992, 114, 10834. 6. a. Huo, Q.; Margolese, D. I.; Ciesla, U.; Feng, P.; Gier, T. E.; Sieger, P.; Leon, R.; Petroff, P. M.; Schuth, F.; Stucky, G. D. Nature 1994, 368, 317; b. Huo, Q.; Margolese, D. I.; Ciesla, U.; Demuth, D. G.; Feng, P.; Gier, T. E.; Sieger, P.; Firouzi, A.; Chmelka, B. F.; Schuth, F.; Stucky, G. D. Chem. Mater. 1994, 6, 1176. 7. Tanev, P. T.; Pinnavaia, T. J. Science 1995, 267, 865. 8. a. Attard, G. S.; Glyde, J. C.; Goltner, C. G. Nature 1995, 378, 366; b. Bagshaw, S. A.; Prouzet, E.; Pinnavaia, T.J. Science 1995, 269, 1242. 9. Corma, A.; Navarro, M. T.; Perez-Pariente, T. J. Chem. Soc. Chem. Commun. 1994, 147. 10. a. Tanev, P. T.; Chibwe, M.; Pinnavaia, T. J. Nature 1994, 368, 321; b. Pinnavaia, T. J.; Tanev, P. T.; Jialiang, W.; Zhang, W. in “Advances in Porous Materials” MRS Symp. Proc. Ser. (Sh. Komameni, J. S. Beck and D. M. Smith, Eds.), 1995, Vol. 371, p. 53, MRS, Pittsburgh. 11. Franke, 0.; Rathousky, J .; Schulz-Ekloff, G.; Starek, J .; Zukal, A. in “Zeolite and Related Microporous Materials: State of the Art 1994, Stud. Surf. Sci. Catal.“ (J. Weitkamp, H. G. Karge, H. Pfeifer and W. Holderich, Eds.), 1994, Vol. 84, p. 77, Elsevier Sci. BV, Amsterdam. 12. a. Sankar, G.; Rey, F.; Thomas, J. M.; Greaves, G. N.; Corma, A.; Dobson, B. R.; Dent, A. J. J. Chem. Soc., Chem. Commun. 1994, 2279; b. Maschmeyer, T.; Rey, F.; Sankar, G.; Thomas, J. M. Nature 1995, 378, 159. 13. Blasco, T.; Corma, A.; Navarro, M. T.; Pariente, J. P J. Catal. 1995, 156, 65. 14. Reddy, J. S.; Dicko, A.; Sayari, A. in “Third International Symposium on Synthesis of Zeolites and Expanded Layered Compounds” Anheim, 1995, in press. 15. Gontier, S.; Tuel, A. J. Catal. 1995, 157, 124. 16. Zeitler, V. A.; Brown, C. A. J. Am. Chem. Soc. 1957, 79, 4618. 17. Hursthouse, M. B.; Hossain, M. A. Polyhedron 1984, 3, 95. 55 i1 18.1(0m l9.Ftuh1 20. Frah 211511: 22. Hor 23. Tan 24. van 25. Boc C at 26. (30.! 27.21. ( l9! Sm 38. Ca 29. W. 30. Be 31.Y; 33- G: 33. F C 18. Kortum, G. "Reflectance Spectroscopy", Springer-Verlag, New York, 1969. 19. Frahm, R. Nucl. Instrum. Meth. Phys. Res. 1988, 270, 578. 20. Frahm, R.; Wang, J. Jpn. J. Appl. Phys. 1993, 32, 188. 21. Lytle, F. W.; Greegor, R. B.; Sandstrom, D. R.; Marques, D. R.; Wong, J.; Spiro, C. L.; Huffman, G. P.; Huggins, F. E. Nucl. Instrum. Meth. 1984, 226, 542. 22. Horvath, G.; Kawazoe, K. J. J. Chem. Eng. Jpn. 1983, I6, 470. 23. Tanev, P. T.; Pinnavaia, T. J. Chem. Mater. 1996, 8, 2068. 24. van Koningsveld, H.; Jansen, J. C.; van Bekkum, H. Zeolite 1990, 10, 235. 25. Boccuti, M. R.; Rao, K. M.; Zecchina, A.; Leofanti, G.; Petrini, G. Stud. Surf Sci. Catal. 1989, 48, 133. 26. Gontier, 8.; Tue], A., Zeolites 1995, 15, 601. 27. a. Geobaldo, C. F.; Bordiga, S.; Zecchina, A.; Giamello, E. Catal. Lett. 1992, 16, 109; b. Bonneviot, L.; Trong On, D.; Lopez, A. J. Chem. Soc., Chem. Commun. 1993, 685; c. Trong On, D.; Denis, 1.; Lortie, C.; Cartier, C.; Bonneviot, L. Stud. Surf. Sci. Catal. 1994, 83, 101. 28. Cannillo, E.; Mazzi, F.; Rossi, G. Acta Cryst. 1966, 21, 200. 29. Waychunas, G. A. Am. Min. 1987, 72, 89. 30. Behrens, P.; Felsche, J .; Vetter, S.; Schulz-Ekloff, Jaeger, N. I.; and Niemann, W. J. Chem. Soc., Chem. Coomun. 1991, 678. 31. Yarker, C. A.; Johnson, P. A. V.; Wright, A. C.; Wong, J.; Greegor, R. B.; Lytle, F. W.; Sinclair, R. N. J. Non—Cryst. Solids 1986, 79, 117. 32. Greegor, R. B.; Lytle, F. W.; Sandstrom, D. R.; Wong, J.; Schultz, P. J. Non- Cryst. Solids 1983, 55, 27. 33. Farges, F.; Brown Jr., G. E.; Navrotsky, A.; Gan, H.; Rehr, J. J. Geochimica Cosmochimica Acta 1995, submitted. 34. Sayari, A.; Karra, V. R.; Reddy, J. S.; Moudrakovski, I. L. in “Advances in Porous Materials” MRS Symp. Proc. Ser. (Sh. Komameni, J. S. Beck and D. M. Smith, Eds.), 1995, Vol. 371, p81, MRS, Pittsburgh. 35. Kumar, S. B.; Mirajkar, S. P.; Pais, G. C. G.; Kumar, R; Kumar, R. J. Catal. 1995, 156, 163. 56 Abstract Wat framework. neutral (8" orthosilicatu = 90 : 10 textural p0; mitture, w; Chapter 3 Tailoring the Framework and Textural Mesopores of HMS Molecular Sieves Through an Electrically Neutral (S°I°) Assembly Pathway Abstract Water : ethanol solvent mixtures of differing polarity have been used to tailor the framework and textural mesopores of HMS molecular sieve silicas through an electrically neutral (S°I°) assembly pathway (8° = dodecyl or tetradecylamine; I° = tetraethyl orthosilicate). Mesostructure assembly from a water-rich solvent mixture, water : ethanol = 90 : 10 (v/v), afforded wormhole-like framework structures with a complementary textural pore volume equal in magnitude to the framework pore volume. An ethanol-rich mixture, water : ethanol = 35 : 65 (v/v), also formed wormhole-like frameworks, but the textural porosity was less than 20% of the framework pore volume. HMS derivatives with high textural porosity were comprised of mesoscale fundamental particles that aggregate into larger particles. In contrast, HMS mesostructures with low textural porosity were assembled into much larger aggregates of macroscale spheroid to disk- shaped fundamental particles. The differences in particle textures were attributed to differences in I° hydrolysis rates and S°I° nucleation and growth rates in the two solvent systems. The presence of mesitylene in the reaction mixtures resulted in an expansion of the framework pores under water-rich conditions. Pore contraction, however, was observed with mesitylene present under ethanol-rich conditions. This versatile structure- modifying prOperty of mesitylene in S°I° assembly is explained by the solvent-dependent binding of the aromatic molecules to two structurally distinct and size-altering 57 “dissolved” and “adsorbed” states at the centers and interfacial surfaces of the surfactant micelle, respectively. Thus, both the framework and the textural pores of HMS silica can be readily tailored to the needs of a particular materials application through S°I° assembly by a judicious choice of an appropriate solvent and an auxiliary structure modifier. 58 (to 110 by me 5+ am Sill. silt. Slllr elec ”II-pr 3.1 Introduction The discovery of Mobil MCM-41 mesoporous molecular sieves 1 has stimulated great interest in the surfactant-directed assembly of mesostructures by sol-gel methods. Thus far, three general assembly pathways have emerged. First, the electrostatic charge matching pathway between cationic or anionic surfactant micelles and charged inorganic precursors. Among these, the so-called S+I‘ and S‘I+ pathwayszr3 represent the most commonly encountered electrostatic assembly pathways. Counterion mediated S+X‘I+ and S'M+I' assembly processes, where surfactant and inorganic reagents are brought together at the micelle interface through triple ion interactions, represent extensions of the charge matching pathway3. The second pathway pairs neutral amine surfactants (S°) or nonionic polyoxyethylene surfactants (N°) with neutral inorganic precursors (1°) through hydrogen bonding at the S°I° or N °I° interface4'3. The third general assembly route to mesostructures exploits dative bond formation between donor groups on the surfactant and a metal acceptor centers in the inorganic precursor9. The original synthesis of hexagonal MCM-4l silicas was accomplished through a S+I‘ assembly mechanism using quaternary ammonium ions as the surfactant and silicate anions as the inorganic precursor]. Related mesoporous structures, designated HMS silicas, have been obtained through S°I° assembly, wherein S° is an alkylarnine and P is a silicon alkoxide5. In general, surfactant removal from the neutral frameworks HMS silicas can be achieved by simple solvent extraction, whereas the displacement of the electrostatically bound surfactants from the anionic framework of as-synthesized MCM- 41 requires proton exchange or destruction of the surfactant by combustion“). More importantly, there are significant structural differences between MCM-41 and HMS 59 silicas. MCM—41 derivatives typically exhibit three or more X-ray diffraction lines indicative of long range hexagonal channel packing. In contrast, HMS silicas show one or, at most, two broad X-ray peaks as a consequence of a small crystallite domain size and/or a much lower degree of channel packing order. Although they are distinguishable with regard to structural ordering, MCM-41 and HMS molecular sieves both exhibit a sharp step in their nitrogen adsorption isotherms, corresponding to the presence of a regular mesoporous framework. Owing to the very small elementary particle size of many HMS derivatives, they can exhibit complementary textural mesopores, in addition to framework pores. The textural pore volumes for HMS can be up to 1.5 or more times as large as the framework pore volumes, whereas MCM- 41 exhibits very little textural mesoporosity“. The textural mesopores are important because they greatly facilitate mass transport to the framework mesopores. For this reason the catalytic reactivity of HMS is usually superior to MCM-41, especially for conversions involving large substrates in a liquid reaction medium where reaction rates are diffusion limited”. Another potential benefit of S°I° assembly is the possibility of conducting mesostructure synthesis in media of diverse polarity. Unlike their ionic counterparts, 8° and 1° reagents are generally soluble in a wide range of solvents. Thus, solvation effects on the rates of hydrolysis and assembly might be an effective means of controlling structure. One of the objectives of the present study was to examine the possibility of tailoring both the framework and textural mesopores of HMS silicas by controlling the polarity of the reaction medium in which the assembly process is carried out. Our 60 approach It rich and the We directing : equilibriur nneeHesl3 structure 1 be of gr. materials 3.2 Expe approach makes use of two water : ethanol compositions to control polarity, one water- rich and the other ethanol-rich. We also have investigated the role of mesitylene as an auxiliary structure- directing agent in both solvent systems. By using solvation effects to shift the equilibrium between structurally distinct binding states of mesitylene in the surfactant micellesl3‘15, we are able to effect an expansion or contraction of the framework pore structure of HMS. The ability to control both framework and textural mesoporosity can be of great value in designing HMS materials as catalysts, adsorbents and sensor materials. 3.2 Experiment HMS molecular sieves were prepared by S°I° assembly pathways in water : ethanol solvent mixtures of differing composition and polarity. In both reaction media tetraethyl orthosilicate (TEOS) served as the neutral silica precursor and dodecylamine and tetradecylamine were the neutral structure director. In a typical synthesis the surfactant was dissolved in ethanol, and then the desired amount of water was added under vigorous stirring to obtain a homogeneous solution. TEOS was added to the surfactant solution and the mixture was allowed to react under stirring at ambient temperature for about 20 h. The reactions were carried out in an open beaker in a well ventilated hood to allow for some evaporation of solvent and concentration of the solid reaction products. When mesitylene was used as an auxiliary structure director, it was added to the surfactant solution and stirred for 15 minutes before the addition of TEOS. All of the HMS reaction products were filtered and dried in air. Although the surfactant 61 HE: can be readily removed from HMS mesostructures by solvent extraction, there was no need to recover the surfactant from the small quantities of products formed in the present work. Consequently, the as synthesized products were directly calcined at 650 0C in air for 4 h to simutaneously remove the surfactant and dehydroxylate the framework. For the purposes of probing the effect of solvent polarity on textural porosity, it was desirable to form HMS silicas from water—rich and ethanol-rich solutions with equivalent framework porosities. To achieve equivalent framework pore structures, we used reagent concentrations in the ethanol-rich system that were twice the concentrations for the water-rich medium. For HMS assembly in the relatively low polarity, ethanol - rich reaction medium, where water : ethanol volume ratio was 35 : 65 , the molar composition of the reaction mixture was 1.0 TEOS : 0.25 Surfactant : 18 EtOH : 34HzO. For assembly in a relatively high polarity, water - rich solvent mixture, namely, 90 : 10 (v/v) water : ethanol, the molar composition was 1.0 TEOS : 0.25 Surfactant : 10 EtOH : 130 H20. Powder X-ray diffraction patterns were measured using Cu-Ka radiation (£21542 A) and a Rigaku Rotaflex diffractometer equipped with a rotating anode operated at 45 kV and 100 mA. The scattering and receiving slits were 1/6 degree and 0.3 degree, respectively. N2 adsorption and desorption isotherms at -196 0C were obtained on a Coulter Omnisorp 360CX Sorptometer operated under continuous adsorption mode. Pore size distributions were calculated from the N 2 adsorption branch using the Horvath—Kawazoe model. 16 62 lOOCX accelerat estimate high ma about 30 suspensi minutes. sectionir 3-3 Resu l [Cmperat mesosm The Othe 3 ethanol StruCture and the e 35 an aUX Fl Silicas respectivg su . “Clam Transmission electron microscopy (TEM) studies were carried out on a JEOL lOOCX instrument using an electron beam generated by a CeB6 filament and an acceleration voltage of 120 kV. The resolution of the instrument was about 6 A, as estimated by indirect measurement of the spherical aberration constant 17 under medium— high magnification (i.e., 100,000X). Therefore, it was possible to resolve pores above about 30 A. The specimens were prepared by dipping a carbon coated copper grid into a suspension (0.1 wt %) of mesoporous material in ethanol that was pre-sonicated for 10 minutes. Attempts to use thin-sectioned specimens were abandoned, because thin sectioning caused damage and loss of texture-pore information. 3.3 Results Two versions of HMS silicas were prepared through S°I° assembly at ambient temperature in reaction media that differed in solvent polarity. In one reaction system the mesostructures were formed from a “water—rich” solution of 90 : 10 (v/v) water : ethanol. The other reaction medium was a less polar “ethanol-rich” solution of 35 : 65 (v/v) water : ethanol. Two S° surfactants, namely, dodecylamine and tetradecylamine, were used as structure directors. The reaction stoichiometries were the same for both the water-rich and the ethanol-rich reaction systems (S°/I° = 0.25). When mesitylene (Mes) was present as an auxiliary structure director, the Mes/8° molar ratio was 1.0 or 45. Figures 3.1 and 3.2 provide N2 adsorption-desorption isotherms for the HMS silicas obtained from dodecylamine and tetradecylamine as structure directors, respectively. The adsorption properties of the mesostructures assembled from the two surfactants under the same reaction conditions are qualitatively equivalent. As can be 63 Figure 3,] Figure 3.1 2000 H U] c ? N2 volume adsorbed (cc/g, STP) p—L 2 ? ‘ l , I 500‘ :BOH-rich-t-MES J' 1 HK=21A D 0 0.2 0.4 0.6 0.8 1 PIPo N2 adsorption-desorption isotherms for HMS silicas assembled at ambient temperatures from dodecylamine and TEOS: Curves A and B are for derivatives obtained from a water-rich (water : ethanol = 90 : 20 (v/v)) and ethanol-rich solution (water : ethanol = 35 : 65 (v/v)); curves C and D are for derivatives prepared from the same water-rich and ethanol-rich solutions, but in the presence of mesitylene (Mes/8° = of 1.0). The HK values are the Horvath-Kawazoe pore diameters. 64 :0 l‘\“\ IGI-IclU-ho Gin-nth! -‘ '6 fiflm37 s-=c H NH: 14:! I°=T EOS Hao-rleh + MS N2 volume adsorbed (cc/g, STP) o' ' '05 ' 'oii . 'ole' ' 'o‘.s' i 1 PlPo Figure 3.2 N2 adsorption-desorption isotherms for HMS silicas prepared from tetradecylamine and TEOS in water-rich and ethanol-rich solution. The labeling of the isotherms is the same as in Figure 3.1. 65 seen fr water-r positio (S°=C have adsor; the m1 medic TEM BSOC seen from the isotherms labeled A and B in both Figures, the structures obtained from the water-rich and ethanol-rich solutions give stepped-shaped isotherms at P/Po < 0.4. The positions of the steps correspond to Horvath-Kawazoe pore sizes of 28-29A (S°=C12H25NH2) and 33-35 A (5° = C14H29NH2). Although the mesostructures obtained from the water—rich and ethanol-rich media have equivalent frameworks, the textural mesoporosity, as evidenced by the N2 adsorption/desorption in the region P/Po > 0.4, depends dramatically on the polarity of the medium used for assembly. A textural pore volume even larger than the framework pore volume can be obtained from the water-rich system, whereas the ethanol-rich medium generates a mesostructure with little or no textural pores. As will be seen from TEM images presented below, the high textural mesoporosity for the water-rich system is associated with the presence of extremely small fundamental particles. We consider next the effect of mesitylene on the framework pore structure and textural mesoporosity of HMS silicas assembled from water-rich and ethanol-rich solutions. The structure mediating properties of mesitylene is manifested in the adsorption-desorption curves labeled C and D in Figures 3.1 and 3.2. The presence of mesitylene at Mes/S° = 1 in the water-rich systems causes the adsorption step to be significantly shifted to higher relative pressure. The shift in the step position corresponds to a 3-5 A increase in HK pore size. In the ethanol-rich medium, however, the presence of mesitylene results in a shift of the step position to lower relative pressures, corresponding to a 5-7 A decrease in HK pore diameter. 66 2 E=~C> C noncomtm o .200 Ehm figure dodecy 15004 I C12o Horvath~szoe Plot N 0.3 . 2 ”‘5 o z .1 g ‘ - 38 A 1000-_ 0.1 2 N2 volume adsorbed (cc/g, STP) Orrrlrrrlrirlrrrlrrr 0 0.2 0.4 0.6 0.8 1 PlPo Figure 3.3 N2 adsorption-desorption isotherms for a HMS silica prepared from dodecylamine in water-rich solution in the presence of mesitylene (Mes/S = 4.5): (A) after calcination at 650 °C with retention of framework and textural mesopores and (B) after calcination at 1000 0C with collapse of framework pores but with retention of textural pores. Insert: Horvath-Kawazoe pore size distribution after calcination at 650 °C. 67 depend solvent increas. a mino under e from 3 ethanol 3.3, the pressur Obtains b}. the 10 Although mesitylene can substantially expand or contract the framework pores depending on the polarity of the reaction medium, it does not alter the key role of the solvent in regulating the textural porosity. As will be shown below, mesitylene actually increases the textural mesoporosity under water - rich assembly conditions, but it has only a minor influence on the extremely low textural pore volume of HMS when assembled under ethanol - rich conditions. To further probe the influence of mesitylene on the framework pores assembled from a water-rich environment, we repeated HMS assembly in 90 : 10 (v/v) water : ethanol at a much higher Mes/8° ratio of 4.5. As shown by the N2 isotherms in Figures 3.3, the adsorption step due to framework pore filling is further shifted to higher relative pressure as a consequence of a HK pore size (38A) that is 9A larger than the value obtained in the absence of mesitylene. The insert to Figure 3.3 shows the half width of the HK pore distribution to be ~10 A, a value typical of HMS materials. On the basis of the hysteresis loop at higher relative pressure, it appears that the textural pore volume is substantially increased by ~ 50 % from ~550 ml STP/g in the absence of mesitylene to ~850 ml STP/g by the presence of mesitylene. In order to verify that the textural pores do indeed arise from interparticle voids and not from a large pore component of the framework, the HMS silica assembled from S° = C12H25NH2 at 90: 10 (v/v) H2O : ethanol and Mes/8° = 4.5 was calcined at 1000 °C to collapse the framework pores. The N2 adsorption/desorption isotherms are given by curve B of Figure 3.3. Note that the framework pores indeed have collapsed, as signified by the loss of the pore filling step, but a substantial fraction (>50%) of the textural 68 Huma4 Figure 3.4 TEM images of calcined (650 °C) HMS molecular sieve silicas assembled from dodecylamine : a. mesoscale fundamental particles obtained from water—rich solution; b. the spheroid fundamental particles and macroscale beads-on-a-string texture obtained from ethanol-rich solution. 69 porosity i: structure. Tl Structure : - and ethz solvent 5) the space is no app; more wor T: reaction reaction fundame] aggregate are Sllgg porosity is retained. Thus, the textural porosity cannot be a consequence of framework structure. The TEM images shown in Figure 3.4 provide insights into the framework structure and textural mesoporosity associated with HMS materials assembled from water - and ethanol - rich media. Regular framework pores are readily observed, regardless of solvent system used to prepare the products. It is quite clear that the pores originate from the space initially occupied by uniform supramolecular assemblies of surfactant, but there is no apparent long range order to the pore arrangement. Instead the pore packing motif is more wormhole - like , perhaps, even sponge-like, in character. The most important distinguishing feature between the water - and ethanol - rich reaction products is the particle texture. As shown in Figure 3.4a, the water - rich reaction mixture yields irregularly-shaped mesoscale fundamental particles. These fundamental particles aggregate into larger particles. The coastline - like silhouette of the aggregates, which persists over length scales differing by at least an order of magnitude, are suggestive of a fractal texture. Textural mesopores comparable in size to the mesoscale fundamental particles are clearly evident. In contrast, the ethanol-rich medium forms spheroid to disk-like fundamental particles that are typically 100 nm or larger in size (see Figure 3.4b). These macroscale fundamental particles are linked into larger aggregates with a beads-on-a—string texture. The interparticle voids are much larger, typically in the macropore range (>50 nm). Figure 3.5 provides TEM images for a HMS derivative assembled from water-rich solution in the presence of a relatively high concentration of mesitylene (Mes/8° = 4.5) 70 F‘é’uIe 3 Figure 3.5 TEM images of HMS silica prepared from dodecylamine in water-rich solution in the presence of mesitylene (Mes/S = 4.5) and calcined at 1000 °C to collapse the framework pore structure; (a) Low magnification image showing the retention of the textural mesopores and, (b) High magnification image showing the absence of framework pores. 71 FigUre 3 Figure 3.6 Intensity _ d=484 s~=c H NH, 123 |°=TEOS rip-rich; Mods -—..4.s l \ Hzo-rlch; MedS°=1.0 tip-rich; MedS°=0 l EON-rich; Meds “=0 d=34.0 A EtOH-rich; MedS °=1.0 O 5 1O 15 20 26) (degrees) X-ray powder diffraction pattern of calcined (650 °C) HMS silicas assembled from dodecylamine amine in water-rich and ethanol-rich solution with or without mesitylene as an auxiliary structure director. 72 but calcin- obtained a obtained at destroyed results desc the origin 0 Figt assembled showed qt relatively 1 near 50° 2 rich polar PositiOns POladty 0‘ but calcined at 1000 0C to collapse the framework mesopore structure. Image (a), obtained at low magnification, shows the retention of the textural pores. Image (b), obtained at higher magnification, shows that the framework mesopores indeed have been destroyed by thermal treatment. This result, which is consistent with the N2 adsorption results described earlier (cf, Figure 3.3), further identifies the interparticle voids as being the origin of the textural mesopores. Figure 3.6 provides the X-ray powder diffraction patterns for HMS derivatives assembled from C12H25NH2 as the structure director. Structures formed from C14H29NH2 showed qualitatively equivalent diffraction features. The patterns all contain a strong, relatively broad reflection at 2.0 - 30° 20 and a very weak broad shoulder in the region near 50° 20. The qualitative form of the patterns is not affected by the water- or ethanol- rich polarity of the assembly medium or by the presence of mesitylene. However, the positions of the intense reflection and the weak broad shoulder are dependent by the polarity of the reaction medium and by the presence of mesitylene. Table 3.1 summarizes the basal spacings, HK pore sizes, N2 BET surface areas (8351), total liquid pore volumes (Vt), liquid framework pore volumes (Vfr), the ratio of textural to framework pore volumes (Vtx/Vfr), and the bulk densities for our HMS materials. The basal spacings represented by the strong diffraction line are correlated with the HK pore sizes, even though the framework lacks regular long-range order. The BET surface areas are in the range 900-1464 m2/g. The incorporation of mesitylene into the synthesis of HMS from a water - rich medium increases the HK pores size and decreases the surface area of the mesostructure. Conversely, mesitylene decreases the pore sizes and increases the surface areas of HMS assembled from ethanol-rich solution. 73 framework i ln contrast. eqUit'alent l differences SUbstantiall Ethanol-rte” 3.4 Disc“S Th asstUnbly 1 HMS silic , which g ludicious 10 ("I") it TEM ima; are 011 a COUVCTSElt It is especially noteworthy from the results in Table 3.1 that the total pore volumes are much larger for HMS derived from a water - rich medium (~1.3 - 1.7 cc/g) than an ethanol -rich medium (~0.5 - 0.7 cc/g). Yet, the framework pore volumes are confined to the approximate range 0.5 - 0.7 cc/g. The difference between the total and framework pore volumes is expressed as the textural pore volume, Vtx. For HMS assembled from a water-rich medium, the textual pore volume which can be up to 1.6 times as large as the framework volume, depending on the amount of mesitylene used as a structure modifier. In contrast, the textural pore volume for HMS obtained from an ethanol-rich medium is equivalent to only a small fraction (6 - 18 %) of the framework volume. Finally, the differences in total pore volumes is manifested in the bulk densities, which are substantially lower for HMS derived from a water-rich medium (0.16 - 0.35 g/cc) than an ethanol-rich medium (0.53 - 0.67 g/cc). 3.4 Discusion The results of the present work demonstrate several important advantages of S°I° assembly for the preparation of mesoporous metal oxide molecular sieves. In the case of HMS silicas we have shown through N2 adsorption studies that the textural mesoporosity , which greatly facilitates access to the framework mesopores, can be controlled by a judicious choice of solvent (£12, Figures 3.1 and 3.2). A water-rich solvent, such as 90 : 10 (v/v) water : ethanol, promotes the formation of mesopore fundamental particles sizes. TEM images show that the interparticle voids responsible for the textural mesoporosity are on a length scale comparable to the fundamental particles (c_f., Figure 3.4a). Conversely, a solvent of lower polarity, namely, 35 : 65 (v/v) water : ethanol, minimizes 74 .. ruin: - .Ceu 1‘0 ICNLICCI _ c. : Edema o_om >9 conceded wmoEw o>o_m 53090:. mEI 60 0mm» noEo‘mo no“ mcoCoEmch‘moncl g 03m _ |.E 23:088. .IOEHONI .2 33 oo H on new or u 00 9o; wcozaom notigm 0cm cocémI “.0 20060068 of. 75 mod mod 3.0 8.0 t F F mm Em C. Townes. Bottom no.0 0 F0 rod 0o.0 omm on flow ouomhos. ”notioum ofio 0v; No.0 no; 000 me who o.vuom\mo_2 EEONI «Izommzu 0N0 00.0 no.0 00. F 5N0 0v «do 0. Fuomhos. £0:-O~I no.0 3.0 no.0 om. F moor v0 «.3 ouowhos. EotémI no.0 o0.0 $0 5.0 33 _.m 0.3 0. Fuomhofi Eocigm no.0 o F .0 oo.0 oo.0 050w om Yam chowaos. EociOfl ofio 0o; no.0 mo; $0 on 0.3 o.vuow\mo_2 EotémI «322320 5.0 5; mod no; 000 No 0.9 0.Fuom\mos_ EEONI 00.0 0 P... mod on; mac? 0w 5. 3 Howie—2 Eo:-OmI 83+ who 98 0M8 20 9on 30 mEauos. x390 =>\5> :> .> hmmm xi 0 5:0on ow Eaton. cociQm one 20:63.. 5 2063mm o_om 3 09390 32.6 o>o_m 8.30206 mi: A00 88 350.3 .2 929530 33095 .P 2an the textur There is n water : er hysteresis pores even Figure 33 application The likely is detc the reaction mesostructur Virtually win rapid nuclea flmdamenral 3.4a). HOWE and mesosri temperature mesostructuI in SlZe, The imerpafllcle The r manly Of (h faults in Tab the textural porosity by forming much larger fundamental particles (9f, Figure 3.4b). There is no doubt that the textural mesoporosity for HMS assembled in 90 : 10 (v/v) water : ethanol arises from interparticle pores. The retention of a well expressed hysteresis loop in the N2 adsorption/desorption isotherm show the retention of textural pores even after the framework pores have been collapsed by calcination at 1000°C (c_f., Figure 3.3). Thus, the textural porosity can be tailored to a particular materials application by simply controlling the polarity of the solvent for supramolecular assembly. The relationship between fundamental particle size and solvent polarity most likely is determined by the relative rates of 1° hydrolysis and supramolecular assembly in the reaction medium. A water - rich medium leads to rapid nucleation of the mesostructure. For assembly in 90 : 10 (v/v) water : ethanol, 3 solid product is formed virtually within minutes of mixing the reagents at ambient temperature. This leads to rapid nucleation and the formation of irregularly shaped fundamental particles. The fundamental particles further aggregate into a self-similar agglomerates @11, Figures 3.4a). However, in the lower polarity 35 : 65 (v/v) water : ethanol solvent, 1" hydrolysis and mesostructure assembly is relatively slow, requiring several hours at ambient temperature for the onset of product formation. The slower nucleation and growth of the mesostructure results in spheroid to disk-shaped fundamental particles 100 nm or larger in size. These fundamental particles interpenetrate to form even larger aggregates with interparticle voids beyond the mesopore range (c_f., Figure 3.4b). The role of mesitylene as an auxiliary structure director is highly dependent on the polarity of the medium in which mesostructure assembly is carried out. As shown by the results in Table 3.1 for HMS assembly in the water-rich medium, mesitylene enlarges the 76 ,5 v A“ I .,o I =0 "-2: r‘ :3: I}: ‘9 3 Q: I * Flgllre 3.7 HMS assembi central COre ol mesitylene head 0 cFOUp SlZe Micellar Binding States of Mesitylene in S°I° assembly Sfliatemcggunt 1|!th ? Sigicate-surfacunt "Sb 51 P3 (‘89s ; g o 4 /0~ v -‘ HI} H “N"? "u [no \0 ‘Sio § 03(0‘35: QN If“ I ”at“ H ©9© H H0 H s 0 Q Q“ Ll1111'.v f ”0"” ‘V OHQHQN N G SQ“ N N O H H H H u a». M n MES H© 1" “Lb- "‘O 0N W wvwvaO 0" 0N W WNO H H H moon 31 NM T 0 H .I-I‘PNJJ dchmedium a H a: g H20/&0H=35:6s H i ° H H H %H ©§©g rte HHo H H). H H '1 H 11%“ MES InH20 nch medium H2CNE10H=9010 0 \; / 0 \ ’9‘ j ./ 0 Si — 510 § 0 \ .o“ H\ 0H 9 OSi\ Q H "N H H a Q “H N H 0’ T "N H .n“\“ osr ace ‘ H H ® H 0N N“ H H H H ONH :0 H 4% "NJ! H in "ii" HNQH Silicate-mflactant micelle expansion Figure 3.7 Schematic representation of the structure-directing effects of mesitylene on HMS assembly. In the water-rich media, the mesitylene “dissolves” in the hydrophobic central core of the micelle leading to pore expansion. In the ethanol-rich environment, mesitylene preferentially “adsorbs” to interfacial head groups, thus increasing effective head group size and subsequently decreasing the pore size. 77 frameworl with the e 0n the Oil framework The is known I “dissolved" the hydroph between the the surfactar states of me Figure 3.7 fc HMS asseml “dissolved“ 1 Of the micell mesOStI'uCtur in the medir Hl'dmgen be with the pol; adsorbed Stat. the sizc of ”1‘ decrease in th. hydrophobic l framework pores and increases the textural mesoporosity. This latter result is consistent with the effect of mesitylene on the pore size of MCM-41 prepared by S+I' assembly 1. On the other hand, for assembly in the ethanol-rich medium, mesitylene reduces the framework pore size and decreases still further the already low textural pore volume. The role of mesitylene on the framework pore structure is intriguing. Mesitylene is known to bind to surfactant micelles in at least two binding states, namely, in a “dissolved” state at the hydrophobic center of the molecule, and in an “adsorbed” state at the hydrophilic micelle / solvent interface 13'15. It is reasonable to expect the equilibrium between these structurally distinct states to depend on the polarity of the solvent in which the surfactant micelles are formed. In accordance with the structurally distinct binding states of mesitylene in surfactant micelles, we propose the mechanism illustrated in Figure 3.7 for the expansion or contraction of the framework pores by mesitylene during HMS assembly. In water-rich solvent, in which the solubility of mesitylene is low, the “dissolved” binding state should be favored over the adsorbed state. In this case the size of the micelle will be increased and this will be manifested as an enlarged pore in the mesostructure. But in the lower polarity ethanol-rich solvent, the solubility of mesitylene in the medium will be increased and this will favor binding at the “adsorption” site. Hydrogen bonding between the p-electrons of the aromatic and water dipoles associated with the polar head groups at the micelle-solvent interface is believed to stabilize the adsorbed state of mesitylene 13. Binding of mesitylene at the micelle surface increases the size of the effective polar head group. The increase in head group size leads to a decrease in the radius of curvature and to a reduction in micelle size in order to retain hydrophobic interactions between the surfactant chains at the micelle center. 78 Consequer mesostruc l In t that certaii textural me tetradecylar sensitive to the present mesoporosit Figures 3.1 . W e framework ‘1 Cltannels w; diffraction F the results 0 EMS mater Obtained {Or Consequently, a framework with a smaller pore size is formed in the assembled mesostructure. In our earlier studies of HMS assembly in 50:50 (v/v) water : ethanol we noted that certain alkylamine surfactants (e.g., dodecylamine) gave derivatives with high textural mesoporosity, whereas others afforded relatively low textural pore volumes (e.g., tetradecylamine) ‘3. It is now apparent that the degree of textural mesoporosity is quite sensitive to both the solvent polarity and the nature of the S° surfactant. The results of the present work show that dodecylamine and tetradecylamine afford high textural mesoporosity when the solvent is highly polar, as in 90 : 10 (v/v) water : ethanol (c_f., Figures 3.1 and 3.2). We have previously described as-synthesized HMS materials as neutral framework analogs of hexagonal MCM—41 5. Evidence for the hexagonal ordering of channels was obtained from selected area electron diffraction 4. The X-ray powder diffraction patterns were attributed to a very small scattering domain size. It is clear from the results of the present study, however, that the broad diffraction lines characteristic of HMS materials are not due exclusively to a small scattering domain size. The patterns obtained for the fine grained HMS particles are indistinguishable from those obtained for the much larger spheroid to disk-like fundamental particles (of, Figure 3.6). Consequently, framework disorder, in addition to small scattering domain sizes, plays an important role in broadening the diffraction lines. On the basis of the TEM images obtained in the present work (pf, Figures 3.4 and 3.5), we find no evidence for hexagonal long range channel packing order. Even “disordered" hexagonal channel packing, which has been previously documented for MCM-4l prepared by S+I' assembly, 19 appears to 79 be difficu the reactio -like HM.“ W0 obtained b) like framet electrostatic ethylenediar structures is nearly unifor 0n the basis Work and the HMS materi; quantitative d these and Ollie be difficult to achieve by S°I° assembly. Additional work is progress to better describe the reaction conditions that determine the assembly of hexagonal vs. wormhole or sponge - like HMS framework structures. Wormhole motifs have been observed for silica and alumina mesostructures obtained by N010 assembly, where N ° is a polyoxyethylene surfactant 6‘3. Also, sponge- like framework structures have been described for meSOporous silicas obtained by electrostatic S+I' assembly in the presence of a structure disrupter (e.g., ethylenediaminetetraacetate) 20. Distinguishing between wormhole and sponge-like pore structures is not a straight forward matter. “Wormholes” imply channel structures with nearly uniform diameters, whereas “sponges” imply substantially reticulated structures. On the basis of the relatively narrow HK pore size distributions observed in the present work and the presence of channel-like voids in the TEM images, we prefer describing HMS materials as wormhole frameworks. Future work will require more incisive quantitative descriptions based in part on the X-ray and neutron scattering properties of these and other highly disordered framework structures. 80 'N‘) 3.5 References 1. 99°.“ng 10. 11. 12. 13. 14. 15. l6. 17. 18. 19. 20. Beck, J. S.; Vartuli, J. C.; Roth, W. J .; Leonowicz, M. E.; Kresge, C. T.; Schmitt, K. D.; Chu, C. T-W.; Olson, D. H.; Sheppard, E. W.; McCullen, B.; Higgins, J. B.; Schlenker, J. L. J. Am. Chem. Soc. 1992, 114, 10834. Huo, Q.; Margolese, D. I.; Ciesla, U.; Demuth, D. G.; Feng, P.; Gier, T. E.; Sieger, P.; Firouzi, A.; Chmelka, B. F.; Schuth, F.; Stucky, G. D. Chem Mater. 1994, 6 , 1176. Huo, Q.; Margolese, D. I.; Ciesla, U.; Feng, P.; Gier, T. E.; Sieger, P.; Leon, R.; Petroff, P. M.; Schtith, F.; Stucky, G. D., Nature 1994, 368, 317. Tanev, P. T.; Chibwe, M.; Pinnavaia, T. J. Nature 1994, 368, 321. Tanev, P. T.; Pinnavaia, T. J. Science 1995, 267, 865. Bagshow, S. A.; Prouzet, E.; Pinnavaia, T. J. Science 1995, 269, 1242. Bagshaw, S. A.; Pinnavaia, T.J. Angew. Chem. Int. Ed. Engl. 1996, 35, 1102. Prouzet, E.; Pinnavaia, T. J. Angew. Chem. Int. Ed. Engl. 1997, 36, 516. Antonelli, D. M.; Ying, J. Y. Angew. Chem. Int. Ed. Engl., 1996, 35, 426. Whitehurst, D. M. US. Patent, 1992, 5,143,879. Schmidt, R. E.; Hansen, W.; Stocker, M.; Akporiaye, D.; Ellestad, O. H. J. Am. Chem. Soc. 1995, 117, 4049. Zhang, W.; Froba, M.; Wang, J .; Tanev, P.; Wong, J .; Pinnavaia, T. J. J. Am. Chem. Soc. 1996, 118, 9164. Mukerjee, P.; Cardinal, J. R. J. Phys. Chem. 1978, 82, 1620. Eriksson, J .E.; Gillberg, G. Acta Chem. Scand, 1966, 20, 2019. Gordon, J. E.; Robertson, J. C.; Thome, R. L. J. Phys. Chem, 1970, 74, 957. Horvath, G.; Kawazoe, K. J. J. Chem. Eng. Jpn. 1983, I6, 470. Spence, J .C.H., "Experimental High-Resolution Electron Microscopy ", 1988, P264, Oxford University Press, New York. Tanev, P.T.; Pinnavaia, T.J. Chem. Mater. 1996, 8, 2068. Chen, C. Y.; Xiao, S. Q.; Davis, M. E. Microporus Mater. 1995, 4, l. Ryoo, R.; Kim, J. M.; Ko, C. H.; Shin, C. H. J. Phy. Chem. 1996, 100, 17718. 81 Chapter 4 In-situ Alumination of Mesoporous HMS Molecular Sieves Silica for Use as Alkylation Catalysts Abstract The in-situ alumination of mesoporous HMS molecular sieves assembled via an electrically neutral assembly mechanism using primary amine surfactant as the structure director, tetraethylorthosilicate (TEOS) as the silicon source and NaAlO2 or Al(sec-OBu)3 as the alumination agent was conducted at ambient temperature. The majority of Al-sites (i.e. >90%) were found in a tetrahedral environment by 27Al-NMR. The textural porosity was not alternated by alumination of structurally well formed HMS for Al-loadings up to 8 mol%. Al—HMS with 4 different amounts of textural porosity (O—tex, 0.5-tex, l-tex, 1.5- tex) relative to framework mesoporosity were purposely prepared through control of solvent polarity and the time of introducing the aluminum precursor following the addition of TEOS. The results indicate that the alurrrination of HMS at ambient temperature, using either NaAlO2 or Al(sec-OBu)3, are as effective as the hydrothermal post synthesis alumination of other mesoporous molecular sieves.1 It is not necessary to remove the surfactant prior to effectively incorporate aluminum into framework of mesostructural HMS at ambient temperature. It is also not necessary to use surfactant-free materials to convert sodium-forms of Al-HMS and Al-MCM-41 into their proton forms by ammonium-exchange. The total acidity of Al-HMS catalysts measured by TGA analysis of cyclohexylamine desorption were reasonably equivalent to each other when the Al-loading was ~2 mol%. Al-HMS with higher textural porosity, regardless of the aluminum precursor used, showed significantly higher conversion (up to 65%) of the 82 bulky 2,4-di-tert-butylphenol (2.4-DTBP) reacting with Cinnarnyl alcohol to a bulky flavan (selectivity up to 87%) under very mild reaction conditions (i.e. 60 °C, 6 h). 4.1 Introduction Mesoporous HMS, as a product of electrically neutral assembly, is distinguished from other mesoporous molecular sieves such as MCM-4l by exhibiting significant textural mesoporosity. The complementary textural porosity of HMS, in addition to framework mesoporosity, was observed in the first HMS sample. 2 Improving and tailoring textural porosity was reported more recently. 3 The textural porosity of HMS facilitates diffusion limited catalytic reactions.4 For example, HMS catalysts have better performance than MCM-4l analogous for both 2,6-di-tert-butylphenol (2,6-DTBP) peroxidation 5 and NO reduction by NH3.6 No significant physical and chemical property differences except textural porosity have been found between these two analogs. The better performance for HMS derivatives can not be attributed to different active sites for those mesoporous molecular sieves. For cumene cracking, substantially higher activity of Al-HMS than MGM-41 analogous have been reported.7'8 Higher acidity and smaller domain size of HMS with apparently shorter diffusion channels might account for the better performance of Al-HMS in cumene cracking. Another very important feature of Al-HMS is that the negative charges generated by substitution of Al into the silica framework can be automatically balanced by the hydrolysis of the primary amine to its protonated form. Thus, the H-form Al-HMS could be easily obtained by calcination without tedious ion exchange required for transformation of sodium-form Al-MCM-4l to H-form. And also, the majority of 83 surfactant in Al-HMS (>90%) can be easily removed by simple ethanol extraction. Dealumination upon removal of surfactant by calcination can be purposely avoided through facile ethanol extraction for Al-HMS, but this is more difficult to accomplish for Al-MCM-41 .9" ' Aluminum substituted mesoporous molecular sieves with mild acidity '0 and large pores have been recognized as good catalysts for fine chemical syntheses where large molecular substrates are usually used. ‘2 Al-HMS with high meso-textural pores is apparently better than other Al-substituted mesoporous molecular sieves in terms of accessibility to framework active sites. Therefore, it is very important for catalysis to have HMS derivatives that are rich in textural porosity and have meso-scaled fundamental particles that limit the pore channel lengths to being comparable to the fundamental particle size itself. One of the objectives of this report is to prepare Al-HMS derivatives with large textural porosity, as well as a small (<150 A) fundamental particle size. The importance of these physical properties to catalysis were verified in the alkylation of 2,4- di-tert-butylphenol (2,4-DTBP) with cinnamyl alcohol to a flavan. This particular liquid phase reaction was first reported by Corma and his coworkers.l3 It was reported that this reaction was not possible to take place inside the microporous channels of H-Y because the molecular size of the substrates prevented them from accessing acidic centers in the smaller channels. Even though the conversion of 2,4-DTBP catalyzed by Al-MCM—41 was high, a very significant amount of it was dealkylated to 4-tert-butylphenol (4-TBP, ~25% yield) rather than to flavan.13 In our study, Al-HMS with meso-scaled fundamental particles and high textural porosity has shown remarkable selectivity (i.e. 86%) to flavan 84 at a significantly higher conversion of 2,4-DTBP (i.e. 65%) without detectable amounts of 4-TBP in the reaction products. Another objective is to prepare Al-HMS with cost-effective aluminum sources such as NaAlO2, because all of the Al-HMS derivatives reported so far were synthesized using aluminum alkoxide, which is expensive and sensitive to moisture. 4.2 Experimental Our approach to the synthesis of Al-HMS with high. textural porosity and a small fundamental particle size using either aluminum alkoxide or NaAlO2 was based on assembly of HMS framework in a relatively high polarity medium 3 and a follow-up alumination step. A water medium with 10 vol% ethanol as co-solvent in the presence of mesitylene (MES) as a textural porosity promoter was used. The synthesis was carried out using dodecylamine as the neutral surfactant and tetraethyl orthosilicate (TEOS) as the neutral inorganic precursor. The aluminum precursor was introduced into the synthesis mixture in a follow—up manner. Three different Al-HMS derivatives were prepared by control the time of introducing the aluminum precursor. First, add aluminum precursor solution (Al(sec-OBu)3 in sec-butanol or N aAlO2 in water) immediately after addition of TEOS to surfactant solution; Second, add aluminum precursor solution 20 minutes after addition of TEOS; Third, add aluminum precursor solution 20 h after addition of TEOS. The reaction mixture was aged under stirring at ambient temperature for 20 h after addition of aluminum precursor. The molar composition of our synthesis mixture was: 0.02~0.08 Al : 1.0 Si : 0.25 Surf. : 1.12 MES : 5.0 EtOH : 130 H20. 85 For comparison, a zero-textural porosity Al-HMS was synthesized using aluminum alkoxide as the alumination agent in a more non-polar medium comprised of EtOH/H2O with a volume ration of 50/50. The molar composition of this synthesis was 0.02 Al : 1.0 Si : 0.25 Surf. : 13 EtOH : 42 H2O The product for each synthesis was filtered and dried in air. The samples prepared using aluminum alkoxide as the aluminum source were calcined at 650 0C for 4 h without prior NHf-ion-exchange. The as-made samples prepared using NaAlO2 as the aluminum source were ion—exchanged at ambient temperature for 4 h using 0.1 M NH4NO3 solution with a solution to solid ratio of 50 ml/g. The obtained product was dried in air and calcined at 650 °C for 4 h. The Al-MCM-4l was prepared using 27wt% sodium silicate solution as the silicon source, and cetyltrimethyl ammonium bromide as the structure director. The pH was initially adjusted to 10.5 using sulfuric acid and adjusted again to 10 after the reaction mixture was heated at 100 0C for 1 day. Subsequently, the reaction mixture was heated at 100 °C for another day. Then, A12(N03)3 aqueous solution was added at ambient temperature under stirring to the reaction mixture, which was statically aged at 100 0C for another 2 days. Aluminum nitrate was chosen as the aluminum precursor over other aluminum sources such as aluminum sulfate and sodium aluminate. The purpose for this choice is to avoid the undesired acidic sites generated by residual anions (i.e. 8042') in the case of aluminum sulfate and the structural alternation caused by pH increase in the case of sodium aluminate. The molar composition of the reactant in this synthesis was: 0.02 Al : 1.0 SiO2 : 0.25 Surf. : 0.39 Na2O : 0.24 H2SO4 : 83 H20 86 The product was filtered and washed thoroughly with water. The filter cake was dispersed in 0.1 M NH4NO3 solution for 4 h at ambient temperature to have it ion-exchanged. Then it was filtered and dried in air before calcination at 540 °C for 6 h. The powder X-ray diffraction pattern of the samples was recorded on a Rigaku Rotaflex Diffractometer using Cu K, radiation 0c: 1.542 A). N2 adsorption and desorption isotherms were measured on an ASAP 2000 sorptometer at -196 0C. The sample was evacuated under 10'5 torr at 150 °C over night before the measurement. 27A1 NMR spectra were recorded on a Varian VXR-4OOS spectrometer with a 7-mm zirconia rotor, and a spinning frequency of 4 kHz. TEM images were taken on a JEOL 100CX with a Ce85 gun that was operated at an acceleration voltage of 120 kV. The specimen was loaded onto a holey carbon film that was supported on a copper grid by dipping the grid into a solid sample suspension in ethanol. The total acidity of a catalyst was measured by means of TGA analysis of cyclohexylamine desorption in the temperature region 240 to 420 °C. Cyclohexalymine was pre-adsorbed on the catalyst and the sample was then purged with N2 at 200 °C for 1 h before temperature programmed desorption was started. Alkylation of 2,4-di-tert-butylphenol (2,4—DTBP) with cinnamyl alcohol was carried out in a 100 ml round bottom flask. A 1.0mmol quantity of 2,4-DTBP, 1.0mmol cinnamyl alcohol and 50 ml of isooctane as the solvent were used. A 500mg quantity of catalyst was added when the reaction mixture was heated to 60 °C. The reaction was kept at 60 °C under magnetic stirring for 6 h. After reaction, the catalyst was filtered and washed with dichloromethane. The dichloromethane was separated from the filtrate by distillation at ~30 °C under reduced pressure so that heat-induced side reactions could be 87 prevented. The reaction products were qualitatively confirmed using a GC-MS. They were quantitatively analyzed by an internal standard method using a GC with a FID detector and a 15 m long SPB-l type capillary column. In order to obtain the response factor of the flavan relative to internal standard that was 2,4-di-tert-butylbenzene, pure flavan (i.e. ~95%) was purified from a large batch reaction using a TLC approach. The conversion was calculated by dividing the amount of 2,4-DTBP consumed by the amount initially present. The selectivity toward flavan was calculated by dividing the amount of falvan formed by the amount of 2,4-DTBP consumed. The yield of flavan was obtained by dividing the amount of flavan formed by the amount of 2,4-DTBP added. The conversion of cinnamyl alcohol was virtually complete in most reactions because '4 such as cinnamyl alcohol alone could transform into many undesirable products cinnamyl aldehyde, 1,5-di-phenyl pentadiene, and polymeric aldehyde. The latter products formed a tar on the catalyst, which required calcination in air to regenerate the catalyst. 4.3 Results and Discussion 4.3.1 Catalyst Synthesis In order to obtain acidic sites for catalysis, aluminum is usually incorporated into the framework of microporous zeolite or mesoporous silica molecular sieves under hydrothermal conditions in situ.15'16 Alumination of mesoporous silica MCM-41 by sodium aluminate in a previous report showed that the aluminum species were predominantly incorporated into the framework near the pore openings, if the alurrrination was carried out in the presence of surfactant molecules inside the mesoporous channels.17 88 This is a particularly significant problem for well-ordered large particle sized MCM-41. The interpretation for this result might be as following. First, the surfactant molecules in the uni-dimensional channels of MCM—41 limit the diffusional migration of aluminum species to deep inside the channels. Second, the interfacial electrostatic charge matching ‘8 between framework and surfactant micelles generated a large electrostatic repulsion to negatively charged aluminum species, therefore, the aluminum was predominantly incorporated at the opening of the pores. In contrast, the framework of HMS assembled through an electrically neutral mechanism was easily aluminated at ambient temperature even in the presence of surfactant micelles in the pores. A reasonable interpretation for this might be that the diffusion resistance was minimized for aluminum species to migrate into the meso-scaled pores with relatively short lengths in meso-scaled fundamental particles to become incorporated into the neutral framework. For sodium forms of mesoporous molecular sieves, previous reports very much rely on ammonium exchange of the (calcined) surfactant free materials and further calcination to generate the H-forrn.'9'20 However, we found that it is not necessary to remove the surfactant before ammonium exchange to obtain effective catalysts. Sodium form of Al-HMS and Al-MCM-41 were transformed into their ammonium forms easily through NHi-exchange at ambient temperature in the presence of surfactant molecules without noticeable NH4+ diffusion limitations. 4.3.2 Catalyst Characterization Figure 4.1 and 4.2 show N2 adsorption-desorption isotherms of Al-HMS prepared using Al(sec-OBu)3 and NaAlO2, respectively. The textural porosity increases with the 89 2000 ~ 2%AI-HMS L Time of AI(OBu)3 h Addition (h) A 20 n. 1500 r V . :7, . H-K (A) "‘ ' 55 t 1.5 > g . 33A 3 1000 " ”3 m 2 10 0 33A 5 O > ZN LlllllklllLlllllL O 0.2 0.4 0.6 0.8 1 Relative Pressure PlPo Figure 4.1. N2 adsorption and desorption isotherms for calcined (650 0C) Al-HMS molecular sieves with different textural porosity. Each sample was prepared at ambient temperature using dodecylamine as the structure director, TEOS as the silicon source, and using Al(sec-OBu)3 as the alumination agent, but the alumination agent was added to the reaction mixture 0, 0.33, and 20 hrs after mixing the TEOS and the surfactant. The aluminum content in each case is 2.0 mole%. The Horvath-Kawazoe (H-K) framework pore size, and the ratio of textural to framework mesoporosity (Vthfr) is provided for each sample. 90 I 2%AI-HM S 2000 F ' Time of NaAlO2 & ' Addition (h) «'72 0'5 2 1500 i m 3 w 1000 E 2 O > ZN e e e e :rgoeerezeee soc t" W 0.5 I; l 1 l l l l l l l l l l l l l I O 0.2 0.4 0.6 0.8 1 Relative Pressure PlPo Figure 4.2. N2 adsorption-desorption isotherms for calcined (650 °C) Al-HMS molecular seives prepared as described in the caption to Figure 4.1, except that the alumination agent was sodium aluminate. 91 delay-time for introduction of the aluminum precursor after addition of TEOS. It is obvious to conclude from Figure 4.1 and 4.2 that the textural pore volume depends very closely on the time of the aluminum precursor addition to the synthesis medium. The highest textural pore volume relative to framework pore volume, was achieved from an in-situ alumination of siliceous HMS with the surfactant micelle still in the mesopore channels. Al-HMS with high textural porosity as high as ~l.5 times of framework mesoporosity (denoted 1.5-tex-Al-HMS) was obtained when the alumination agent was added '20 h after the hydrolysis of the TEOS in the presence of the surfactant at ambient temperature. Due to the fact that most of the framework skeleton of HMS is formed within first 20 minutes after addition of TEOS to the surfactant solution, structural alternation would be minimized if aluminum precursor was added 20 minutes after addition of TEOS. Actually, a Al-HMS with the textural porosity equivalent to the framework mesoporosity, (denoted 1.0-tex-Al-HMS) was obtained at this alumination time. Interestingly, when the aluminum precursor was added immediately after TEOS in an aluminum-silicon co-incorporation manner, the Al-HMS exhibits much lower textural porosity with smaller meso-scaled textural pores indicated by the uptake of N2 at relatively smaller P/Po (i.e. 0.7). The ratio of textural to framework pore volume (VuNfr) was ~0.5, hence, denoted 0.5-tex-Al-HMS. In addition, the step in the isotherm indicative of N2 filling of framework mesopores at P/Po ~0.3 was smeared, and the framework pore volume and pore size was decreased. This suggests that aluminum incorporation at the initial stage of the assembly deteriorate the formation of framework, especially for neutral S°I° assembly. This result is in agreement with Al-MCM—41 which exhibits a much broader (100) reflection, and weakened higher order reflections, in comparison with 92 Figure 4.3. 7001 600: p: : (I) . 5 500: 3 . 3 I E 400: h . o .. m .- 8 3001 a, . E . a . 3 200‘. > - s N V Z t 1 00-. g \“\\1)).V.V.vxv;v_v_v_v_ - 0 20 40 60 30 100 ‘ (A) o I I I l I I T r r I U l I I I r I I j 0 0.2 0.4 0.6 0.8 1 Relative Pressure PIPo N2 adsorption-desorption isotherms for calcined (650 °C) 2% Al-HMS and calcined (540 °C) Al-MCM-41 with zero-textural porosity. The inset shows Harvoth-Kawazoe pore size distributions for these two samples. This Al-HMS was prepared using alumnum sec-butoxide as the alumination agent, which was added to the reaction mixture 20 minutes after mixing the TEOS and the surfactant solution. 93 siliceous MCM-4l.2' Our observations on the effect of co-assembly of aluminum with silicon in HMS is in agreement with the results of Al-HMS prepared by Jones and his co— worker.7 They mixed aluminum alkoxide and TEOS before adding the mixture to the surfactant solution. The total pore volume of their mesostructure was significantly reduced from 1.68 mllg for the pure silica analogous to 0.65 ml/g by incorporation of just ~2.5 mol% Al into the framework. Interestingly, the Al-HMS prepared by Jones et al did exhibit higher acidity than Al-MCM-41 analogous although textural porosity was not observed. We verified their result by preparing a Al-I-IMS in a relatively less polar solvent medium with a EtOH/H2O ratio of 50/50 (see below for discussion about acidity). As showed in Figure 4.3, this Al-HMS exhibits no hysteresis loop for textural pores in the isotherm and so does the Al-MCM-41. Actually, the N2 isotherm for Al-MCM-41 exhibits very sharp framework mesopore filling step at relative pressure of 0.3. Consequently, the pore size distribution is relatively sharp and narrow also. In contrast, 0- tex-HMS exhibits a somewhat wider pore size distribution. The differences in framework structure, hexagonal vs. wormhole, are reflected in the widths of the pore size distribution. Figure 4.4 and 4.5 show the powder X-ray diffraction patterns of Al-HMS prepared by addition of Al(sec-Obu)3 and NaAlO2 respectively to the TEOS- dodecylamine hydrolysis mixture after various reaction times. Three Al-HMS samples with different textual porosity prepared using aluminum alkoxide as the aluminum precursor show nearly identical d-spacing at 4.2 nm (Fig. 4.4). Relatively higher d- spacing (52 A) along with larger pore size (41 A) was obtained for Al-HMS prepared using NaAlO2 (Fig. 4.5). This may indicate that sodium cations facilitates the pore size 94 41.7 A 2%AI‘HMS {l s°=c H NH,I°=TEOS, 12 25 2 Calcined at 650 °C E / in c .9 5 Time of AI(OBu)3 addition (h) 24 0.33 0 ..,....,s...,.... 0 5 10 15 20 29 (degrees) Figure 4.4. XRD patterns of calcined 2%Al-HMS molecular sieves prepared as described in the caption to Figure 4.1. 95 51.75. f 2%Al-HMS s°=c H NH,I°=TEOS 12 25 2 Calcined at 650 °C .5 .. c on 0 o 'E <- Time of NaAIO2 addition (h) 24 0 WW], 0 5 10 15 20 20 (degrees) Figure 4.5. XRD patterns of 2%Al-HMS prepared as described in the caption to Figure 4.1, except that NaAlO2 was the Al-source. 96 510‘ ' 2%Al-Loading 41d~ g 310 - m -i C o «H E 210 . Al-MCM-41(O-tex) .1 11d- ' AI-HMS (o-tex) "'l""l""l O 5 10 15 20 29 (degrees) Figure 4.6. XRD patterns of calcined 2% Al-HMS and Al-MCM41 with zero—textural porosity as described in the caption to Figure 4.3. 97 enlargement by exchange into the interface of surfactant and framework wall as proposed 22 It worth pointing out that this kind of pore for larger pore MCM-41 synthesis. enlargement process takes place at elevated temperature for MCM-4l, but it occurs for HMS at ambient temperature. Regardless of the type of aluminum precursors, the sample prepared by the addition of an aluminum precursor immediately after addition of TEOS caused the intense reflection to be broadened. This loss of wormhole framework structure is in agreement with the N2 adsorption-desorption isotherms (cf. Fig. 4.1 and 4.2). The XRD patterns for 0-tex-Al-HMS and Al-MCM-41 are showed in Figure 4.6. O-tex-Al-HMS exhibits equivalent reflection to that of pristine HMS.3 Al-MCM-4l exhibits well resolved hexagonal X-ray reflections. Apparently, the structural order of both HMS and MCM-4l were fully retained since both samples were prepared through a follow-up in-situ alumination. Even though they were assembled from a different mechanism, the total pore volumes for both of them are virtually the same (c.f. Fig. 4.3). As stated earlier, it is very important to have mesoporous molecular sieves with small fundamental particle size so that the porous channels can be shortened and diffusion resistance minimized. It was found that meso-scaled fundamental particles are always associated with high textural porosity in pristine HMS. To confirm this for Al- HMS, we examined three derivatives prepared using either Al-alkoxide or NaAlO2 by TEM. As exemplified in Figure 4.7A, fractal-like particles with a worm-hole-like structural motif prevail for both 1.0-tex-Al-HMS and 1.5-tex-Al-HMS. Irregularly shaped inter-particle voids (~150 to ~500 A) are present between aggregates of the fundamental meso-scaled particles ~100 to ~500 nm in size. These voids are the textural pores responsible for the hysterisis loop observed in the high textural pore materials at partial 98 Figure 4.7. TEM images showing meso-scaled textural pores as well as worm-hole like framework mesopores for calcined 1.5-tex-Al-HMS and 0—tex-Al- HMS. 99 Figure 4.8. 1.5-tex-AI—H MS(650 °C) Al-MCM-41 (540 °C) 1.5-tex-H MS (NaAlO 2) O-tex-AI-H MS (AI(OBu) 3) 1 .5-tex-Al-H MS (AI(OBu) 3) _‘ I I I I I I I l V ' I ' ' I I '7 I I I 400 200 0 -200 -400 ppm 27Al-NMR spectra showing predominantly tetrahedral Al-sites (63 ppm) and a small amount (~10%) of octahedral Al-sites ( 18 ppm) in as-made Al-HMS prepared using Al(sec-OBu)3 as the alumination agent. Calcined forms of Al-HMS, as well as calcined Al-MCM-41, show 70% tetrahedral and 30% octahedral Al-sites. 100 pressures above 0.9. Unlike the textural porosity-rich sample, Al-HMS with zero-textural porosity shows sub-micro-sized beads on string type of spheroidal particles (Fig. 4.7B). Although the particle size for a 0.5-tex-Al-HMS is about 2 times larger (not shown) than 1.0-tex-Al-HMS, but they are still about 5 times smaller than that of O-tex-Al-HMS. 27Al-NMR spectra are shown in Figure 4.8. Obviously, the vast majority of aluminum species in as-made Al-HMS are in a tetrahedral coordination environment. These aluminum sites possess a chemical shift at ~63 ppm. The NMR spectra differences between high and low textural porosity samples are practically indistinguishable. Al-HMS prepared using Al-alkoxide shows a slightly higher content of octahedral aluminum species (c.f. the shoulders at around 16 ppm) than the one prepared using NaAlOz, possibly because of aluminum oxide clusters generated by fast hydrolysis of aluminum alkoxide. Calcined Al-MCM-41 and clacined Al-HMS exhibited about 30% octahedral (indicated by 18 ppm shoulders) and 70% tetrahedral aluminum sites that possesses a chemical shift position at 63 ppm also. Nevertheless, the completion of the alumination of HMS at ambient temperature suggests that the surfactant molecules inside the framework pores be disregarded by aluminum precursors. As a consequence of a neutral interfacial interaction (H-bonding) between the surfactant and the framework, anionic aluminum precursor species is allowed to migrate easily through the neutral interfacial gaps in the absence of charge repulsion, and to become incorporated into the framework. The amount of acid of each catalyst was measured through TGA analysis of the amount of cyclohexalamine (CHA) desorbed in the temperature range 240 to 420 °C. As shown in Table 4.1, among all five Al-HMS samples regardless of textural porosity and 101 aluminum precursor, the average amount of acidity was 0.34 mmol CHA/g. This accounts up to 100% of the loading of Al, if it is assumed that each Al generates one acid site. This indicates that both tetrahedral and octahedral Al-sites in Al-HMS contribute to the total acidity, which suggest that octahedral Al-sites in Al-HMS be isolated, possibly grafted Al-sites rather than aluminum oxide clusters. In agreement with the results reported by Jones et al, the acidity observed for our 2 mol% Al-HMS is about 3 times higher than that of 2 mol% Al-MCM-41 (i.e. 0.10 CHA mmol/g). The difference in the acidity strongly Table 4.1. Physical and chemical properties for Al-substituted mesoporous molecular sieves Aluminum Saar Pore Volumesa (egg) H-K Pore Acidityb Catalyst Source (m2/ g) V, Vf, Vtcx Size (nm) (mmol lg) 0-tex-Al-HMS Al(OBu)3 1234 0.86 0 0 2.9 0.37 0.5-tex-Al-HMS Al(OBu)3 968 1.08 0.67 0.41 2.8 0.36 1.0-tex-Al-HMS Al(OBu)3 976 1.38 0.73 0.65 3.3 0.38 1.5-tex-Al-HMS Al(OBu)3 1020 1.77 0.80 0.97 3.3 0.30 1.5-tex-Al-HMS NaAlOz 978 1.95 0.80 1.15 4.1 0.30 Al-MCM-4l A1(NO3)3 1070 0.87 0 0 3.1 0.10 3. The total pore volume (Vt), including the textural pore volume (van), was measured on the isotherm at P/Po = 0.99. The framework pore volume (Vfr) was measured at P/Po = 0.5. b. The total acidity for these catalysts were obtained by means of TGA through measurement of cyclohexylamine deorption between 230 to 420 °C. 102 suggest that about 2/3 of Al-sites for Al-MCM-41 be inaccessible by cyclohexylamine. As long as a 27Al-NMR spectrum of Al-MCM-41 does not exhibit a significantly higher number of octahedral Al-sites, the difference in acidity between Al-MCM-4l and Al- HMS with the same Al-loading can not be attributed to large amount of extra-framework aluminum sites. It appears more likely that the lower total acidity in the case of Al-MCM- 41 is due to lower accessibility of Al-sites buried in the ~12 A thick walls. The possibility of Na+ blocking the acid sites was excluded by the fact that a low Na/Al ratio of 0.1 was observed by ICP elementary analysis of this Al-MCM-41. The acidity for all Al-HMS samples are not considerably different from each other. Therefore, the performance differences in acid catalyzed reaction for Al-HMS molecular sieves can not be attributed to the intrinsic difference of acid sites (see below). In summary, Al-HMS with high textural porosity, small fundamental particle size, and wormhole framework porous structure was prepared by carefully controlling the addition of aluminum precursor during the syntheses. In order to minimize the framework deterioration caused by Al incorporation and to maintain the high textural porosity, the aluminum precursor was introduced into the reaction mixture after the formation of the silica framework (i.e., after 20 h of TEOS hydrolysis in the presence of the primary amine structure directors). It worth to pointing out that both aluminate anion and Al(OH)x(OBu)3-,, type of hydrolysis molecules (i.e., possible aluminum species involved in the alumination process) can diffuse into the surfactant-filled framework mesopores and became incorporated into the silica framework at ambient temperature. 103 4.3.3 Catalytic Results As discussed before, four type of Al-HMS molecular sieves with different amount of textural porosity were prepared through control of the reaction conditions. In order to compare the catalytic activity for the Al-substituted mesoporous molecular sieves, the acid catalyzed alkylation of 2,4-di-tert-butylphenol was chosen because bulky reagents are completely inaccessible to the acid centers inside the micropores of H-Y zeolite 13. The OH \ ..... 0 Acid catal st 0 0 * V > O Cin-OH 2,4-DTBP importance of diffusion might be manifested in this large molecule reaction. In the first report of this reaction using mesoporous MCM-41 as the catalyst, the reaction was carried out at 90 °C for 24 h.13 We found that it is not necessary to carry out this reaction for that long, because one of the two reactants, cinnamyl alcohol, is nearly 100% consumed after 6 h at even 60 °C. Most of the by—products from this reaction were transformed from cinnamyl alcohol alone in the presence of catalyst. Three products were identified from the conversion of 2,4-di-tert-bytylphenol by GC-MS analysis. These three products account for up to ~90% of consumed 2,4-di-tert-butylphenol. The major product is flavan. Two by-products were 4-tert-butylphenol and 4,6-di-tert-butyl benzofuran. Since these 104 two by-products only account for less than 5% of consumed 2,4-di-tert—butylphenol, they are not listed in Table 4.1. Interestingly, 4—tert-butylphenol, the product of dealkylation Table 4.2. Alkylation of 2,4-di-tert-butylphenol with cinnamyl alcohol catalyzed by Al-substituted mesoporous molecular sieves3 Aluminum Conv. (%) Selectivity for Yield of Catalyst Source 2,4-DTBP Flavan (%) Flavan (%) 0-tex-Al-HMS Al(OBu)3 50.8 71.3 36.2 0.5-tex-Al-HMS Al(OBu)3 59.6 77.9 46.4 1.0-tex-Al-HMS A1(0Bu)3 60.9 77.1 46.9 1.5-tex-Al-HMS A1(0Bu)3 61.9 83.2 51.5 1.5-tex-Al-HMSb NaAlOz 65.3 86.8 56.6 Al-MCM-4lb A1(N03)3 47.2 71.8 33.9 3. Each catalyst contains 2 mol% Al; the reaction was carried out using 10.0 mmol of each reactant, 500 mg of catalyst , in 50 ml of isooctane at 60 °C for 6 h. b. These catalysts were ion-exchanged with 0.1 M NH4N03 before calcination. of 2,4-DTBP was barely detectable for most Al-HMS catalysts. However, this was a much more apparent by-product for conventional Al-MCM-4l as the catalyst. As shown in Table 4.2, the conversion of 2,4-DTBP and the selectivity as well as the yield of the desired product flavan all increase with the textural porosity of Al-HMS. The selectivity to fiavan was un—precedently high (i.e. ~87%). It is particularly noteworthy that 0-tex-Al-HMS and Al-MCM-41 exhibit very similar catalytic activity, but the activity is much lower than their HMS analogs with high 105 textural porosity. These results signify the importance of textural porosity in facilitating the access of substrate molecules to acid sites. Our MCM-41 sample exhibited almost the same flavan yield as reported by Corma et al for Al-MCM-41 with approximately the same Al-loading. Corma’s Al-MCM-41 catalyst was NH4+-exchanged after removal of the surfactant by calcination whereas our Al-MCM-41 was NH“-exchanged prior to removal of the surfactant. An important conclusion from the similarity in catalytic activity would be that our approach for ammonium exchange prior to removal surfactant is as efficient as normal ammonium exchange for surfactant-free MCM-4l molecular sieves. The physical and chemical properties for all the catalysts, as shown in Table 4.1, are approximately similar to each other except for the BET surface area and pore volume. Even though both of the BET surface area and the framework mesopore volume for 0.5- tex-Al-HMS are smaller than that of 0-tex-Al-HMS, the catalytic conversion and the yield of flavan for the former catalyst are higher. Therefore, the activities of these Al- HMS catalysts are not directly related to BET surface area and mes0pore volume. Indistinguishable differences in Al-siting indicated in 27Al-NMR (Fig. 4.8) agrees very well with the amount of total acidity for all Al-HMS catalysts measurement by thermal gravity analysis (T GA) of cyclohexlamine desorption. As shown in Table 4.1, the average total acidity for all Al-HMS with 2% Al-loading was about 0.34 mmng, which was very close to the theoretical acidity of 0.33 mmng. Therefore, it is impossible to attribute the differences in catalytic activity to intrinsic differences in Al-siting among Al- HMS catalysts. However, Al-HMS generally exhibit three times higher total acidity than analogous MCM-4l, which is in agreement with the results reported. 7 The acidity 106 difference between Al-MCM-41 and Al-HMS with the same Al-loading might be indicative of the fundamental differences in their assembly mechanisms rather than the residual Na“, because the amount of Na+ left in this Al-MCM-41 was only 10 mol% of the total Al. The results in Table 4.2 unambiguously suggest that textural porosity indeed improve the catalytic conversion of bulky substrates in liquid reactions. In addition to textural pores, the Al-HMS samples with high textural porosity also exhibit very thin (i.e < 100 nm, this is the maximum specimen thickness for electron beam accelerated at 120 kV to go through easily to obtain good quality TEM pictures) and small fundamental particles (~ 150 nm). Consequently, this feature of Al-HMS would provide limited pore channel lengths that should be comparable to the fundamental particle size. Thus the diffusion resistance of bulky molecules would be minimized and this enhances the accessibility of catalytic active centers in the framework channels and favors the maximum conversions. 4.4 Conclusions Al-substituted HMS molecular sieves with different amounts of textural porosity relative to the framework mesoporosity were prepared through control of solvent polarity and the time of introduction of the aluminum precursor to the developing wormhole frameworks. Alumination of HMS molecular sieves assembled in the presence of neutral amine surfactant at ambient temperature using NaAlOz or Al(sec-Obu)3 was as effective as literature-reported hydrothermal aluminum incorporation processes for MCM-41 molecular sieves. The purpose for in-situ follow-up alumination is to avoid alternation of 107 structural features (i.e. high textural porosity etc.) desirable for condensed phase catalysis. Ammonium exchange of sodium-form Al-HMS and Al-MCM-4l at ambient temperature in the presence of surfactant in porous channels was also effective enough to achieve sufficient active H-form mesoporous catalysts. The complementary meso-scaled textural pores as well as meso-scaled fundamental particles of mesoporous molecular sieves were very important to improve the catalytic alkylation of bulky 2,4-DTBP with cinnamyl alcohol to flavan under very mild conditions (60 °C, 6h). The conversion of 2,4-DTBP and the yield of flavan both were dramatically improved in comparison with Al-HMS and Al-MCM-41 catalysts without textural pores. 108 4.5 References: 1. 10. ll. 12. l3. 14. 15. 16. 17. Ryoo,R.; Jun, 8.; Kim, J.M.; Kim, MJ. J. Chem. Soc., Chem. Commun. 1997, 2225. Tanev, P.T.; Chibwe, M.; Pinnavaia, T.J. Nature 1994, 368, 321. Zhang, W.; Pauly, T.R.; Pinnavaia, T.J. Chem. Mater. 1997, 9, 2491. Pinnavaia, T.J.; Zhang, W. Stud. Surf. Sci. Catal. 1998, 117, 23 (Elsevier Science B.V.). Zhang, W.; Froba, M.; Wang, J.; Tanev, P.T.; Wong, J.; Pinnavaia, T.J. J. Am. Chem. Soc. 1996, 118, 9164. Yang, R.; Pinnavaia, T.J.; Li, W.; Zhang, W. J. Catal. 1997, 172, 488. Mokaya, R.; Jones, W. J. Chem. Soc., Chem. Commun. 1996, 983 and J. Catal. 1997, 172, 211. Kloetstra, K.R.; van Bekkum, H.; Jansen, J.C. J. Chem. Soc., Chem. Commun. 1997, 2281. Tue], A.; Gontier, S. Chem. Mater. 1996, 8, 114. Corma, A.; Fomes, V.; Navarro, M.T.; Perez-Pariente J. Catal. 1994, 148, 569. Hitz, S.; Prins, R. J. Catal. 1997, 168, 194. Corma, A. Chem. Rev. 1997, 97, 2373; and Sayari, A. Chem. Mater. 1996, 8, 1840. Armengol, E.; Cano, M.L.; Corma, A.; Garcia, H.; Navarro, M. J. Chem. Soc., Chem. Commun. 1995, 519. Armengol, E.; Corma, A.; Garcia, H.; Primo, J. Appl. Catal. A: General 1995, 126, 391. Szostak, R.; Molecular Sieves Principles of Synthesis and Identification; Van Nostrand Reinhold: New York, 1989. Beck, J.S.; Vartuli, J.C.; Roth, W.J.; Leonowicz, M.E.; Kresge, C.T.; Schmitt, K.D.; Chu, C.T.-W.; Olson, D.H.; Sheppard, E.W.; McCullen, S.B.; Higgins, J .B.; Schlenker, J .L. J. Am. Chem. Soc. 1992, 114, 10834. Ryoo, R.; Ko, C.H.; Howe, R.F. Chem. Mater. 1997, 9, 1607. 109 18. 19. 20. 21. 22. Janicke, M.T.; Landry, C.C.; Christiansen, S.C.; Kumar, K.; Stucky, G.D.; Chmelka, B.F. J. Am. Chem. Soc. 1998, 120, 6940. Luan, Z.H.; Cheng, C.F.; He, H-Y.; Klinowski, J. J, Phys. Chem. 1995, 99, 10590. Gunnewegh, E.A.; Gopie, S.S.; van Bekkum, H. J. Mol. Catal. A: Chemicals 1996, 106, 151. Luan, Z.H.; Cheng, C.F.; Zhou, W.Z.; Klinowski, J. J. Phys. Chem. 1995, 99, 1018. Corma, A.; Kan, Q.; Navarro, M.T.; Perez-Pariente, J.; Rey, F. Chem. Mater. 1997, 9, 2123. 110 Chapter 5 Toward MCM-4l with Meso-scaled Channel Lengths and Textural Pores for Catalysis Abstract The concept of altering structures of mesoporous silica molecular sieves through introducing organic promoters to bind to the head group of charged surfactants in the electrostatic S+I' assembly has been successfully illustrated. Mesoporous silica molecular sieves with well ordered 3-dimensional hexagonal, cubic, and uni-dimensional hexagonal phases were selectively assembled using cetyltrimethyl ammonium bromide as the structure director and sodium silicate as the silicon source under a relatively low surfactant to silicon ratio (i.e. 0.11) at 150 °C. Intrinsically, the structural variation was strategically realized through surfactant-organic promoter association that leads to the variation of surfactant micelle packing parameters (g = Vlad) and, ultimately, the inorganic meso-structures. Organic promoters such as salicylic, tartaric, oxalic, and citric acid were found effective to alter meso-structures from hexagonal to either 3- dimensionally hexagonal or cubic. These promoters were also effective at mild hydrothermal condition (i.e. 100 °C) for synthesis of particular MCM-41 with a flaky morphology. These flakes exhibit mesopore channels running orthogonal to the flaky sheets with meso-scaled lengths. Considerably well ordered MCM-41 molecular sieves with meso-scaled fundamental particles and high textural porosity were also prepared by careful control of the addition of organic acids. Consequently, a follow-up in situ alumination of the flaky and textural pororosity MCM-4l in the presence of surfactant micelles inside the pores affords well incorporated 111 Al-sites with over 95% are in tetrahedral coordination even for Al-loading up to 8.8mol%. However, up to a third of these Al-sites are transformed into octahedral sites upon the surfactant removal by calcination. Al-sites judged from 27Al-NMR and the total acidity measured from TPD of cyclohexylamine are indistinguishable from each other for similarly loaded Al-MCM-41 catalysts regardless of particle morphology. Consequently, any differences in catalysis for these Al-MCM-41 catalysts can not be attributed to any differences in the intrinsic acid siting. The number of Al-sites with stronger acidity indicated by the desorption of pyridine at 350 °C was proportionally increased with Al- loading, but the number of Al-sites with weaker acidity indicated by the desorption of pyridine at 217 oC remained the same. The results of the studies of 2,4-DTBP alkylation with cinnamyl alcohol proves that the accessibility of the active Al-sites in flaky and textural porous Al-MCM-41 with short channel lengths is significantly facilitated. At least 30% higher conversion of 2,4- DTBP and 50% higher yield of flavan are obtained relative to Al-MCM-4l catalyst without a flaky or textural porous morphology. This alkylation reaction is likely catalyzed by the mild acid sites since increasing the number of stronger acid sites by increasing the Al-loading does not improve the activity considerably. 112 5.1 Introduction M4lS materials are ordered, mesoporous silicas and aluminosilicate molecular sieves with pore size in the range of 20 — 100 A.1 Among M41S family members, hexagonal MCM-41 and cubic MCM-48 are very promising catalysts and catalyst supports for chemical reactions involving large molecular substrates.2 The pore channels for MCM-48 are bi-continuous, branched and tortuous.3 The pore channels for MCM-4l, however, are uni-dimensional. Uni-dimensional channels may invoke counter-stream diffusion resistance, because the reactants need to diffuse in and the products need to diffuse out of the tubular channels.4 The significance of this particular drawback for MCM-4l has not yet attracted attention. Even though MCM-48 exhibits attractive 3-d pore connectivity, the difficulty to synthesize hetero-atom incorporated cubic meso-structures severely limited the study of this particular meso-structure and far fewer reports have been published.5'8 Actually, the synthesis of Al-MCM-48 is even more problematic, most of the time the obtained materials lack catalytic properties that are distinctive from MCM-4l. Other than the 3-d pore connectivity, MCM-48 has not been found to be fundamentally different from MCM-4l.7 Therefore, for the purpose of catalysis it is desirable to simplify MCM-48 synthesis or to improve the accessibility to the uni-dimensional channels of MCM-41. MCM-4l silica molecular sieves have been synthesized under various conditions. Two distinguishable electrostatic assembly pathways have been reported using positively charged quaternary ammonium surfactants (S+) as the structure director. One is the surfactant—inorganic precursor charge matching mechanism (SIT) working under basic conditions using silicate species (I') and the other is the counterion mediated assembly 113 mechanism (S+X'l+) working under acidic conditions using protonated silicon alkoxide species (1+). 9 The later one encounters significant difficulties for the synthesis of metal- substituted derivatives since the catalytically active metal-centers are barely incorporated into the silica framework under the severe acidic conditions (i.e. pH < 2).10 Although it is possible to synthesize MCM—41 under ambient conditions via the ST assembly pathway, the structure obtained, however, is lacking in thermal and hydrothermal stability.ll Therefore, most MCM-41 samples for catalysis, especially Al-substituted analogs, were synthesized under hydrothermal conditions through the ST pathway."12 To prepare MCM-41 with well ordered hexagonal structure via S‘T pathway, relatively high concentrations of surfactant and high temperatures (> 100 °C) were adopted by Mobil scientists.l Unlike most zeolitic structures that normally crystallize under very basic conditions (for example, NaY forms at pH above 12), mesoporous MCM-41 structures with amorphous silica framework walls usually form under less basic condition (i.e. pH <11) to minimize the dissolution of amorphous silica walls. Therefore, MCM-41 synthesis under basic conditions using sodium silicate as the silicon source has to use some kind of acid to lower the pH. Ryoo and his co-worker reported a step-by-step acid titration approach to obtain well ordered MCM-41 using acetic acid over 4 days at 100 °C.13 White and his co-worker later reported that sulfuric acid was the best acid for adjusting pH during MCM—4l synthesis by following this titration approach.14 Similarly, Mon and his co-workers'5 prepared a hierarchical tubular MCM-41 structure through a day by day addition of sulfuric acid at 100 °C. The function of acid, other than adjusting pH in these preparations, is not clear yet. 114 In general, meso-structures form through surfactant micelle assembly involving inorganic reagents cross-linking and condensing around the micelles to build a framework. As reviewed by Behrens16 the interfacial interactions between surfactant micelle and inorganic precursors are very crucial for assembly of meso-structures. The interactions primarily determine the final phase of the meso-structures.l7 Actually, in the classical surfactant micelle chemistry,""20 the interaction energy is the key factor for the surfactant packing parameter g (g = V/(aol)). If g is around 1/2, surfactant micelles tend to form a hexagonal phase. A cubic phase forms for g values between 1/2 and 2/3. If the g value approaches 1/3, a 3-dimensional (3-d) hexagonal or a spherical cubic phase is favored. Therefore, to tailor the structure and pore conductivity of MCM-41, the easiest way would be to control interfacial interactions. The first objective of the present study was to alter the interfacial interaction to obtain versatile mesostructures using organic acid as the structure promoter. The second objective of our study is to slightly alter the interfacial interaction using organic acid as structure modifier so that meso-scaled fundamental particles and channel lengths could be obtained for MCM-41. Organic acid was chosen, because salicylic acid was known to associate with the head group of cetyltrimethyl ammonium bromide to form helical fibers.21 By using carboxylic acid, our strategy is to minimize the association energy to such a point that it is not strong enough to cause a phase transition but towards a phase transition, allowing the channel lengths and the orientation of particle growth to be tailored. The third objective was to prepare a well-ordered MCM-41 aluminosilicate with textural porosity. Davis and his coworkers reported 22 that the textural porosity for MCM- 115 41 was closely related to the disorder caused by surfactant-rod-micelle-entangling because a significant amount of N2 uptake in the isotherm for textural pores was accomplished without seeing higher order X—ray reflections in the powder patterns. Our results indicate that the textural porosity is not necessarily associated with the structural disorder. However, it is closely associated with meso-scaled fundamental particles. Fundamental particles aggregate to form meso—scaled inter-particle voids as the textural pores. Textural porosity along with meso-scaled fundamental particles indeed facilitates the access to catalytic centers inside mesopore channels.23 In order to shown the importance of textural porosity for catalysis, alkylation of 2,4-di-tert-butylphenol with cinnamyl alcohol in the presence of MCM—41 aluminosilicate molecular sieves as the catalyst was conducted under very mild conditions (i.e. 60 °C, 6 h). Significantly higher conversion (i.e. >60%) of 2,4-DTBP and higher yield of the desired product flavan (~50%) was obtained for Al-MCM-41 with high textural porosity and meso-scaled fundamental particles relative to the analogs without textural porosity. 5.2 Experimental MCM-48 and 3-dimensional hexagonal mesoporous silicas were synthesized at 150 °C through surfactant directed assembly in the presence of organic promoters such as salicylic, citric, tartaric and oxalic acids. A typical synthesis was started by mixing a sodium silicate aqueous solution as the silicon source and cetyltrimethyl ammonium bromide as the structure director at ambient temperature. The structure promoters were added to the mixture at ambient temperature under vigorous stirring for 15 minutes 116 before the mixture was transferred to a Teflon lined autoclave to age at 150 °C for 15 h. The composition for the synthesis was following: 1.0Si : 0.11CT AB : 0.39Na20 : 0.28H+ : 130HzO. In the case of salicylic acid, the promoter to surfactant ratio was controlled to 1, which also contributed part of 0.28 mole H" per mole of silicon and the rest part of H” was compensated by sulfuric acid. In the case of citric, tartaric and oxalic acid, all of the H+ were came from these promoters. MCM-4l sample was prepared using 27wt% sodium silicate solution as the silicon source and cetyltrimethyl ammonium bromide as the structure director. An organic acid such as tartaric, oxalic, citric, gluconic, and salicylic acid was used as the pH adjuster, as well as the structure modifier. Sulfuric acid was included for comparison. Three types of MCM-4l with different textural porosity were prepared by controlling the amount and the time for acid modifier addition. The molar composition for the synthesis of MCM-4l with textural porosity was: 0.02 — 0.10 Al : 1.0 SiOz : 0.25 Surf. : 0.39 NaZO : 0.58 H+ : 83 H20 To control the pH as well as the association energy between the organic acid and the surfactant head group, part of the acid was added to the sodium silicate aqueous solution before the surfactant was added. The remaining part was added later. In a typical experiment, 83% of the total amount of acid shown in the aforementioned composition formula was added to the sodium silicate solution at the beginning, and the remaining 17% was added after 1 day heating at 100 0C. The MCM-4l obtained exhibits textural pores similar to the textural pores formed by aggregates of layered clay particles. A high textural porosity MCM-41 was synthesized by the addition of 100% of the required 117 amount of acid in one-lot after mixing the surfactant and sodium silicate solution, and then heating at 100 0C for 2 days. A typical Mobil MCM-4l without textural porosity was prepared using sulfuric acid at 100 0C for 2 days. The amount of sulfuric acid was 17% less than organic acid used in the aforementioned compositions. The acid was added at the very beginning and the pH of this synthesis was ~10.5. For siliceous MCM-4l, the product was filtered after 2 days aging at 100 °C. For Al-MCM-41, A12(NO3)3 aqueous solution was added at ambient temperature under stirring to the as-made siliceous MCM-41 in its mother liquor, and then it was statically aged at 100 0C for another 2 days. Actually, this was still an in situ alumination process in a follow-up manner. Aluminum nitrate was chosen as the aluminum precursor over other aluminum sources such as aluminum sulfate and sodium aluminate. The purpose for this choice is to avoid the undesired acidic sites generated by residual anions (i.e. 8042') in the case of aluminum sulfate and structural alternation due to pH increase in the case of sodium aluminate. The product was filtered and washed thoroughly with water. The filter cake was ion-exchanged in 0.1 M W 03 solution at ambient temperature for 4 h if needed. Other wise, the filter cake was dried in air and directly transformed into the final catalyst without prior NHf-exchangment by calcination at 540 °C for 6 h. It is possible that the Na+ for neutralization of the negative framework charges generated by Al-incorporation might be replaced by charged surfactant molecules inside the mesopores. In this case, A]- MCM-41 prepared using sodium silicate may not require ammonium exchange to transform it into a catalytically efficient H-form. 118 The powder X-ray diffraction patterns of the samples were recorded on a Rigaku Rotaflex Diffractometer using Cu K0 radiation (71,: 1.542 A). N2 adsorption and desorption isotherms were measured on an ASAP 2000 sorptometer at -196 °C. The sample was evacuated under 10-5 torr at 150 °C over night before the measurement. 27Al NMR spectra were recorded on a Varian VXR-400S spectrometer with a 7-mm zirconia rotor and a spin frequency of 4 kHz. TEM images were taken on a JEOL IOOCX with a CeB6 gun that was operating at acceleration voltage of 120 kV. The specimens were loaded onto a holey carbon film that was supported on a copper grid by dipping the grid into a solid sample suspension in ethanol. Alkylation of 2,4-di-tert-butylphenol (2,4-DTBP) with cinnamyl alcohol was carried out in a 100 ml round bottom flask. A 1.0 mmol quantity of 2,4-DTBP, 1.0 mmol cinnamyl alcohol and 50 ml isooctane as the solvent were used. A 500 mg quantity of catalyst was added when the reaction mixture was heated to 60 °C. The reaction was kept at 60 0C under magnetic stirring for 6 h. The catalyst was filtered and washed with dichloromethane after the reaction. The dichloromethane was separated from the filtrate by distillation under reduced pressure at ~30 °C, so that heat-induced side reactions could be prevented. The reaction products were qualitatively confirmed using a GC-MS. They were quantitatively analyzed by an internal standard method by using a GC with a FID detector and a 15m long SPB-l type capillary column. The conversion was calculated by dividing the amount of 2,4-DTBP consumed by the amount initially present. The selectivity toward flavan was calculated by dividing the amount of falvan formed by the amount of 2,4-DTBP consumed. The yield of flavan was obtained by dividing the amount of flavan formed by the amount of 2,4-DTBP added. The conversion 119 of cinnamyl alcohol was virtually complete in most reactions because cinnamyl alcohol alone could transform into many undesirable products 24 such as cinnamyl aldehyde, 1,5- di-phenyl pentadiene, and polymeric aldehyde. The latter product formed a tar on the catalyst, which required calcination in air to regenerate the catalyst. 5.3 Results and Discussion 5.3.1 Effect of Organic Acid Addition on Silica Mesostructure Formation It has been known for a decade that salicylic acid can associate with head group of cetyltrimethyl ammonium surfactant to form remarkable helical micelle fibers.2| Some other a— and B-hydroxyl carboxylic acids interact with charged surfactants in basic medium to form similar micelle fibers too. It has also been known that the surfactant packing parameter g can be varied upon variation of the surfactant head group size and the alkyl chain length. Actually, variation of g parameter to achieve new meso-structures has been shown quite successful in the preparation of SBA-2. 25 Landry and his co- worker recently reported 26 that the g parameter of cetyltrimethyl ammonium bromide could be varied to allow formation of cubic MCM—48 instead of hexagonal MCM-4l by using 3 mole of EtOH per mole of silicon. Presumably, association of carboxylic acid to the head group of surfactant would affect the g parameter also. We report here that carboxylic acids indeed promote cetyltrimethyl ammonium bromide surfactant to form 3- d hexagonal as well as cubic silica meso-structures under the reaction conditions normally give rise to uni-dimensional MCM-41. As shown in Figure 5.1, a MCM-48 like cubic silica meso-structure was formed in stead of MCM-41 upon addition of 1 mole of salicylic acid per mole of surfactant. The 120 Figure 5.1. 2" S”I' Pathway S" = CTAB l' = Sodium Silicate $71- = 0.12 EtOH, MOM-48 Intensity 420 332 No EtOH, MGM-48 Salicylic acid/CTAB = 1.0 100 110 200 No EtOH, MGM-41 No EtOH, MOM-41 Salicylic acid! CTAB = 3.0 L L .l 1 l J I 1 1 l O 5 1O 15 20 29 (degrees) X-ray diffraction patterns for calcined (540 oC) cubic silica meso- structures forrned at 150°C with a reaction composition of 1.0Si : 0.11CT AB : 0.39Na2O : 0.28H+ : 130H2O. The diffraction patterns for MCM-48 prepared by following Landry’s method as well as the diffraction patterns for a smaple prepared in the absence of both EtOH and salicylic acid are included. 121 S‘T Pathway 3-d Hexagonal Pam" 100 002 fl 8" = CTAB l' = Sodium Silicate $71. = 0.12 > s -I-' .- . . . '5', o o. Cltrlc acid/CTAB =1.0 c 3: 37- d) —— - 4.0 E Oxalic acid/CTAB = 1.5 Tartaric acid/CTAB = 1.5 l l l l l l l l l l l L l 2 4 6 8 1o 26 (degreeS) Figure 5.2. X-ray diffraction patterns for calcined (540 0C) 3—d hexagonal silica meso- structures formed under the same conditions as noted in Figure 5.1 with the promotion of citric, oxalic and tartaric acids. 122 diffraction patterns of MCM-48 prepared through Landry’s procedure and a comparison MCM-4l sample prepared in absence of salicylic acid and ethanol are included in Figure 5.1 too. Interestingly, too much salicylic acid deteriorates the meso-structure while 3 mole of salicylic acid per mole of surfactant was used. This indicated that the association 2' of carboxylic acid with the charged surfactant head group results in diminished long range 8+1~ electrostatic interaction. Consequently, the structure was less ordered judged by X-ray diffractions. Figure 5.2 shows the well resolved XRD patterns for 3-d hexagonal silica formed upon the addition of 1 mole of citric acid per mole of surfactant. Oxalic acid and tartaric acid shown similar promotion effects. These results strongly suggest that carboxylic acids can function as promoters to change the surfactant packing parameter and ultimately cause the silica meso-structure to change from uni—dimensiona hexagonal to 3-d hexagonal and even cubic phases under certain reaction conditions. In comparison to the formation of conventional MCM-41, 3-d hexagonal and cubic meso- structures were formed under relatively lower surfactant to silicon ratio (i.e. 0.11 vs. 0.25) and H+ to NaOH ratio (i.e. 0.5 vs. 0.75) at 150 °C under the promotion of organic acids. 5.3.2 Siliceous MCM-41 Type Mose-structure Formation Mou and his co-workers proposed '5 that a hierarchical tubular MCM-41 formed through a sheet-like intermediate. The tubular MCM-4l formed at 100 °C by ultimately reaching an over all H+ to Si ratio of 0.57 when sulfuric acid was added day-by-day one time a day over 4 days. Our original goal was to synthesize MCM-41 with thin sheet morphology so that the accessibility of framework channels would be greatly enhanced. 123 Figure 5.3. Gluconic Acid 6? 1 : : : : : . it”; ('3 1000, Lactic Acid as I a: M” g ,' Oxalic Acid 800‘ ' Citric Acid ' Sulfuric Acid Mobil MOM-41 N Volume adsorbed a: O 9 Montmorillonite ....‘ ‘:.: . . - o ‘ . ‘ 0 0.2 0.4 0.6 0.8 1 PlPo N2 adsorption and desorption isotherms for calcined (540 °C) siliceous MCM-41 prepared under the promotion of gluconic, lactic, oxalic, and citric acids with a H+/NaOH ratio of 0.75 showing framework mesoporosity as well as clay-like textural porosity similar to that of montrnorillonite. The Mobil MCM-41 sample was prepared using sulfuric acid to adjust the pH (i.e. ~10.5) at the very beginning of the synthesis by control the over-all H‘VNaOH ratio to 0.625. 124 In order to form and isolate the “sheet intermediate” organic acid was purposely added one time a day, and the mixture was aged only for two days instead of 4 days at 100 0C. The over all H+ to NaOH ratio was controlled to 0.75. MCM—41 with characteristic N2 adsorption-desorption hysterisis loop similar to that for textural pores of layered montmorillinite was obtained (c.f. Figure 5.3). The isotherm for a typical MCM-41 prepared using sulfuric acid does not show any textural pores in contrast. TEM and SEM images of this sample showed tubular MCM-41 particles as the dominant form (Figure 5.4A and 5.4B) along with minor potion of flake-like particles (Q0%). The flaky morphology in the MCM-41 samples prepared using organic acids, however, was the major morphology (c.f. Figure 5.4C and 5.4D), no tubular particles were observed. The pores of flaky particles were funning orthogonal to the flaky-surface. Obviously, the pore channel lengths were limited to the thickness of the flakes. That is estimated to be less than 100 nm, because the penetration depth for an electron beam accelerated at 120 kV is about 100 nm.27 These uni-dimensional pores with short channel lengths should be beneficial for diffusion limited catalytic reactions. Actually, MCM-4l molecular sieves with clay-like flaky features indicated by the presence of textural porous hysteresis loop in the isotherms (Figure 5.3) were prepared using many organic acids while keeping the over-all H+lNaOH ratio at 0.75. All of them exhibited a ~3 A width at half of maximum for the Harvoth-Kawazoe pore size distribution. They all exhibit very well resolved hexagonal X-ray patterns with (100), (110), (200), (220) and (300) reflections clearly seen (Figure 5.5). Although White and his co-workers claimed 14 that a well ordered MCM-41 was obtained through tedious sulfuric acid titration during 4 days heating at 100 °C, the sample prepared by following 125 Figure 5.4A and 5.48 126 Figure 5.4 TEM images showing (A and B) interwoven tubular MCM-41 formed upon use of sulfuric acid; (C and D) thin sheet MCM-41 (formed upon use of citric acid) with pore channels running orthogonal to sheet surface (the thickness of the sheet estimated to be ~ 100nm according to the penetration depth of electron beam accelerated at 120 kV); Both samples were prepared with an over all HVNaOH ratio of 0.75. 127 11d 81¢“ IIIIIIIIIIIIIIIIIIIIIIIIIII] _ 5 1O 15 20 25 30 a. 610 . 29 7, Gluconic Acid r: 0 . H 5 3 4 104 M Oxalic Acid 210‘— k Citric Acid Lactic Acid — O IITIIfj’TIIITrTIIIIIIIIIIIIIII 0 5 10 15 20 25 30 26) (degrees) Figure 5.5. X-ray diffraction patterns for MCM-41 molceular sieves as noted in Figure 5.3. 128 their procedures using sulfuric acid does not exhibit significantly superior order over the sample prepared using organic acid. TEM images (c.f. Figure 5.4) for flaky MCM—41 show primarily sub-micrometer sized thin (< 100nm) sheets with pore channels running orthogonal to the sheet surface. However the particle size of MCM-4l reported by White and Ryoo et al are in micrometers. 28 Textural porosity HMS molecular sieves with inter particle voids of aggregated meso-scaled fundamental particles normally form upon fast hydrolysis and nucleation of hydroxylated neutral silicon species in the presence of alkylamine surfactant.29 In order to facilitate the nucleation of silicate species in MCM-41 synthesis, we purposely added an amount of organic acid equivalent to a H+lNaOH ratio of 0.75 in one lot instead of two to allow the pH drop down to ~9 immediately, and then let the reaction mixture to age at 100 0C for 2 days. As illustrated in Figure 5.6, MCM-41 with a significant textural pore volume was produced in this way. The same type of synthesis performed at ambient temperature, however, afforded MCM-41 with lower framework as well as textural pore volume. Accordingly, the X-ray reflection patterns for this particular sample only showed a broad (100) reflection with a hump at higher 20 angles (c.f. Figure 5.7). This signifies the importance of temperature in nucleation and assembly of meso-structures. In agreement with HMS materials,28 the high textural porosity of MCM-4l was correlated with the low bulky density (~0.2 g/cc), which is only half of the value of the MCM-41 without textural pores. This suggests that the high textural porosity MCM-41 samples contain large amount of fluffy tiny particles. In addition, Figure 5.7 shows that MCM-41 with textural porosity still exhibits X-ray patterns with well resolved (100), (110) and (210) reflections. This is in contrast to the report by Davis and his co- 129 Figure 5.6. MGM-41 assembled with Tartaric Acid modification 1600- pH10.5,then 9, ‘ 3 100°C 1400 - E I pH1os.1oo°c E ‘ ' a. pH9,100°C m 120020105413“, 03 - H.5_.___ E i 10 20 so 40 so so 70 80 V 1000': H-Kpore 3’ 800; (‘- 3 ‘ pH=10.5, then 9, r.t. (U 4 lac-ocvm'” or: g 500 pH =1o.5, then 9,100 °c '5 > 400 .-- - e e e e e : 2292;919:91" z“l " o pH=10.5,100 c 200 “:‘."g‘..... u airwmwuwr... 0 0.2 0.4 0.6 0.8 1 PIPo N2 adsorption and desorption isotherms for MCM-41 molecular sieves prepared using tartaric acid as the structure promoter under various conditions showing the controlled formation of significant amount of HMS-like textural pores as well as clay-like textural pores. Horvath- Kawazoe pore size distribution plots are included in the inset. 130 Calcined (540 °C) MGM-41 Assembled Via S‘I' Pathway With Tartaric Acid Modification 61d_ 51dj 41d- n. .................... : i I‘5 10 159 20 215 310 g- ; pH = 10.5, then 9,100 °c 2 31d- 2 - 5 210— pH=10.,5100 °C 4 110 LU pH= 9,100 °C pH= 105 then9 rt. 0 l 'l""l"‘7 0 5 10 15 20 25 30 26 (degrees) Figure 5.7. X-ray powder diffraction patterns for MCM-41 molecular sieves as noted in Figure 5.6. Notably, the MCM-4l with high textural porosity exhibits well resolved (100), (110) and (200) reflections. 131 Figure 5.8. TEM image of MCM-4l showing meso-scaled fundamental particles with inter particle voids as textural pores and regular hexagonal framework mesopores running orthogonal to thin sheet surfaces exemplified by a sample prepared using tartaric acid as the structure promoter. 132 22 workers, which stated that disordered MCM-41 with single X-ray reflection and a N2 hysteresis loop in the isotherm at relative pressure larger than 0.7 was a consequence of rod-like surfactant micelle entanglement. Therefore, X-ray disorder of MCM-4l is not necessarily associated with textural porosity. In accordance with the samples prepared using citric acid, oxalic acid, gluconic acid and acetic acid, the sample synthesized through two step addition of tartaric acid also exhibited a nice X-ray diffraction pattern containing 5 well resolved reflections (Figure 5.7). In addition, they all exhibited a nice flaky-type hysterisis in the isotherm (c.f. Figure 5.6). These results suggest that the “intermediate sheets” to tubular MCM-41 were likely captured to form flaky MCM-4l under the promotion of those carboxylic acid promoters. Similar to HMS, MCM-41 with high textural porosity does exhibit meso-scaled fundamental particles (Figure 5.8), which are aggregated into fractal-like larger particles. Since this sample was not thin-sectioned, but dusted on the copper grid for taking TEM images, the specimen thickness may be considered as the particle thickness. Therefore, the thickness of these fundamental particles would be estimated to be less than 100 nm, also. Moreover, as long as the pore channels are running orthogonal to the particle surface, the channel length of this particular MCM-4l would be estimated as ~100 nm, comparable to the particle thickness. 5.3.3 Catalyst Preparation Normally, Al-MCM-41 is synthesized through an aluminum-silicon co-assembly process. However, Al-MCM-41 prepared in this way loses higher 29 angle X-ray 133 reflections most of the time. The higher the Al-loading, thebroader the (100) reflection. In order to obtain well ordered Al-MCM-41, a follow-up in-situ alumination in the presence of surfactant micelles in the pores was reported by Ryoo and his co-workers.28 However, this particular sample indicated inhomogeneous alumination of the framework with more aluminum incorporated near the pore openings than deep inside the channels. Actually, Ryoo’s MCM-41 was synthesized through an acid titration approach which gave rise to a well ordered MCM-4l with a relatively large particle size (even larger than 2 pm”). Therefore, the migration of aluminum species to deep inside of the channel is very likely hindered in the presence of surfactant molecules. Moreover, the electrostatic interfacial interaction would generate an electric field to prevent negatively charged aluminum species from accessing the negatively charged framework. Hence, the alumination would not be fully completed. To overcome this difficulty and make the alurrrination of the pore walls more homogeneous in the pores, using aluminum nitrate as the alumination agent for our MCM-41 with meso-scaled fundamental particles and meso-scaled channel lengths should have advantages over other MCM-41 samples. The follow-up in-situ alumination of MCM-41 with flaky particles was only successful for Al-loading up to 9%. The flaky particle morphology was lost in attempts to introduce 20% aluminum. At these higher Al loading, the characteristic hysteresis loop responsible for inter particle voids in the isotherms was diminished. Seemingly, the flaky particles were stacked into larger particles. This suggests that flaky sheets are more likely an intermediate of tubular MCM—41. In contrast to intermediate flaky sheets, MCM-41 with textural porosity showing meso-scaled fundamental particles is stable to the follow—up in situ alumination treatment 134 even for an attempt to introduce 20% aluminum. Elemental analysis of the sample with 20% initially Al-content in the synthesis mixture revealed that the actual Al-loading for 1 500 A ‘ A 1 A A j g g . E Textural (D 1 ‘t‘ 651 ooo~ B . 3 ~ . cl IIIIIIIIIII'III'IIIITIT " 1 0 2 4 6 8 10 12 ‘f‘ h A" O 1 H-K Pore Size (nm) ‘ ‘ A .1 A" ,2 LA- .3 AA-A\“‘ " I < . " Al-MCM-41 (Textural) , at . ‘ E 500- ..-‘ 3 o/‘L 3 ' 2'1"“ > ‘ ‘§““ N " A9“ / Z . f . T I I l I I I j I I I I I I I I I I I 0 0.2 0.4 0.6 0.8 1 PIPo Figure 5.9. N2 adsorption and desorption isotherms for 8.8mol%Al-substituted MCM- 41 molecular sieves (calcined at 540 °C) with clay-like and HMS-like textural porosity prepared using tartaric acid as the structure promoter and aluminum nitrate as the follow-up in situ alumination reagent in the presence of surfactant miccelles in the pores. (The inset is the corresponding Horvath-Kawazoe plots showing sharp narrow distribution of framework mesopores). 135 Al-MCM-41 Octfl’etr 8.8% AI (T extural porous, 540 °C) 21179 88%AI (Flaky, 54o °c 8.8%AI 0/1 00 (Tex, as-made) 8.8%AI 0,100 (Flaky, as-made) ' I r ' ' I ' ' ' I ' 'A‘ I ' ' .fl‘, 400 200 0 -200 -400 mm Figure 5.10. 27Al-NMR spectra for as-made and calcined Al-MCM—41 molecular sieves prepared using tartaric acid as the structure promoter with HMS-like and clay-like flaky textural pores. 136 the final product was only 10%. In fact, the 20% aluminum introduced in the form of aluminum nitrate decreased reaction pH too much to allow the incorporation of Al into the pore walls taking place effectively. As shown in Figure 5.9, this particular Al-MCM- 41 still exhibits a good textural pore volume up to 50 % of the total pore volume and well resolved hexagonal X-ray patterns. Not only did tartaric acid afford Al-MCM-4l with textural pores, but also oxalic acid, citric acid, gluconic acid and lactic acid afforded Al- MCM-41 with quite good textural pores. 27Al-NMR spectra show that over 95% of total aluminum sites in the as-made samples are in tetrahedral coordination (Figure 5.10). This result indicated that the follow-up in situ alumination was as effective as in situ silicon-aluminum co-assembly and post-synthesis implantation30 in the absence of surfactant. 2“’Al-NMR spectra of our samples indicates dealumination upon calcination occurred regardless of high or low textural porosity. Approximately a third and a fifth of the total aluminum-sites were transformed from tetrahedral coordination to octahedral coordination upon calcination for 2%Al- and 8.8%Al-MCM-4l, respectively. This verifies the dealumination reported by Corma 3' and Prins 32 et a1. Notably, there are no observable differences in 27Al-NMR spectrum for MCM-4l with and without textural porosity. In conclusion both intra- frarnework and extra-framework aluminum species are approximately identical from sample to sample for all calcined MCM-4l samples with the same Al-loading regardless of the textural porosity. Interestingly, the ratio of tetrahedral to octahedral Al-sites for the calcined samples increased with increasing Al-loading. This was further confirmed by temperature programmed pyridine desorption. As shown in Figure 5.11, the number of 137 Figure 5.11. 1 2 - Pyridine-TPD profile 1 _ 8.8%Al-MCM-41 0.8 F 0.44 mummy/g ’6 \ _ 3 0'6 0.17 mmol Pylg 1? 0.4 " 1“ 1‘ III \I 0.2 _ : “\ I’ K ‘\i' .‘ \ 1",, ‘\ O _ 0.11 mmOI PYIgai"""'\./ 5..., -,-. 2.0%AI-MCM-41 200 250 300 350 400 450 500 550 600 Temperatu retC) Temperature programmed desorption (TPD) of pyridine pre-adsorbed on A1-MCM-41 catalysts and purged at 150 °C in N2 stream for l h before the TPD started. This TPD profile was carried out on a thermal gravity analyzer. 138 weaker acid sites indicated by pyridine desorption peak at 217 °C remained the same, but the number of stronger acid sites indicated by pyridine desorption peak at 350 oC increased by 4 fold when the Al-loading increased 4 times. Assuming the surfactant micelles are homogeneously distributed through the pores, the sample without prior NHX-exchangement but with higher Al-loading would require higher concentration of Na+ cations at the pore opening to balance the negative framework charges generated by tetrahedrally incorporated A1 if most of the Al-sites were at the pore opening as reported by Ryoo et al. 28 Hence, non-linear increase of acidity value with Al-loading would be expected. Notably, our two samples were not subjected to NHf-exchangment. A linear increase of the number of stronger acid sites with increasing Al-loading strongly suggest that aluminum was not likely just incorporated at pore openings, but rather homogeneously incorporated into the framework through the pores. In agreement with the results of Jones et al,33 our previous study also found that 2%Al-MCM-4l exhibited quite low acidity (0.1 mmol CHA/g, compared with 3.4 mmol CHA/g for 2%Al-HMS) possibly due to inaccessible Al-sites buried in MCM—41 framework walls.34 In order to achieve an acidity for Al-MCM-4l more or less comparable to Al-HMS, aluminum-loading for most of the MCM-41 was increased to 8.8%, which gave acidity levels of 0.44 mmol CHA/g, as measured by TGA analysis (c.f. Table 5.1). This value is also comparable to that of the HY from Nth available from Aldrich. As shown in Table 5.1, all Al-MCM-41 samples with similar Al-loading regardless of surface area, porous volume and textural porosity are very similar to each other in total acidity. Aluminum nitrate may provide nearly neutral aluminum hydroxylated species in the alumination medium with a pH at ~9. Therefore, the follow- 139 Table 5.1. Physical and chemical properties of Al-MCM-41 mesoporous molecular Al-MCM-lsillwf:S Si/Al S351 Pore Volume (cc/g)b Pore Size Acidity" Catalyst ratio (m2/ g) Vt Vf, V,” (nm) (mmol lg) H-Y zeolited 2.80 670 0.30 0.30 0 0.72 0.50 Mobil 50.0 1070 0.87 0.87 0 3.3 0.10 (Sulfuric Acid) 11.1 1051 0.92 0.86 0.06 3.2 0.47 Flaky (Citric Acid) 11.1 903 0.91 0.73 0.14 3.6 0.42 Flaky (Oxalic Acid) 11.1 990 0.87 0.71 0.16 3.5 0.43 Flaky (T artaric Acid) 11.1 922 0.87 0.71 0.16 3.5 0.44 Tex-porous (Tartaric) 50.0 927 0.89 0.72 0.17 3.5 0.14 Non-tex (Tartaric)° 1 1.1 967 0.83 0.83 0 3.5 0.44 Tex—porous (Tartaric) 11.1 1045 1.46 0.82 0.64 3.6 0.43 Tex-porous (Citric) 1 1.1 101 1 1.38 0.72 0.66 3.5 0.43 Tex-porous (Oxalic) 11.1 1023 1.13 0.71 0.42 3.5 0.43 3. All Al-MCM-41 samples were not NRC-exchanged except the sample prepared according to Mobil’s method. b. The total pore volume (V,) and the framework pore volume (V2,) were calculated from the isotherm at P/Po 0.99 and 0.5, respectively. The textural pore volume (Vm) is the difference between V. and V2. °. The total acidity for these catalysts were obtained by means of TGA through measurement of cyclohexylamine deorption between 240 to 420 °C. d. This H-Y was obtained through calcination (540 °C) of a NH4-Y with a Si/Al ratio of 2.8 and containing 2.5 wt% Na2O, which was bought from Aldrich. °. This sample has no textural pores prepared using 17% less amount of tartaric acid comparable to Mobile’s MCM-41. 140 up in situ alumination using aluminum nitrate as the reagent in the presence of surfactant molecules in the pores are likely incorporated homogeneously into the framework all the way through the pores. One more reasonable conclusion derived from the data in Table 5.1 and Figure 5.10 and 5.11 is that any differences in catalysis for those Al-MCM-4l with equivalent acidity would not be attributed to any differences in intrinsic acid siting. 5.3.4 Catalytic Alkylation of 2,4-di-tert-butylphenol with Cinnarnyl Alcohol As discussed in last section, a family of Al-MCM-41 with textural pores was prepared using various organic acids through a general method. Like the preparation of Al-HMS with textural pores, the follow-up in situ alumination of MCM-41 with textural pores using aluminum nitrate is efficient enough to incorporate tetrahedral Al into the OH \ CH2OH 0 + m Acid catalyst> I o Cin-OH 2,4-DTBP framework in the presence of surfactant molecules. In order to evaluate the catalytic performance for acid-sites, alkylation of 2,4-di-tert-butylphenol was conducted over Al- MCM-41 catalysts with various textural pores. The importance of accessibility to active sites in diffusion limited catalytic reaction might be manifested by these Al-MCM-41 catalysts because of their different framework pore accessibility. In the first report 35 of using mesoporous MCM-41 as the catalyst, this reaction was carried out at 90 °C for 24 h. As we stated,34 it is not necessary for running this reaction for 24 h because one of the two reactants, cinnamyl alcohol is nearly 100% consumed after 6 h run at even 60 0C. Most of the by-products (about 10% of the total products) from this reaction were from 141 cinnamyl alcohol alone. Three products were identified for conversion of 2,4-di-tert- butylphenol by GC-MS analysis. These three products account up to ~90% of the consumed 2,4-di-tert-butylphenol. The major product is flavan. Two by-products were 4- ten-butylphenol and 4,6—di-tert-butyl benzofuran. Since these two by-products only account for less than 5% of consumed 2,4-di-tert-butylphenol for all Al-MCM-41 catalysts, they were not listed in Table 5.2. Interestingly, 4-tert-butylphenol, the product of dealkylation of 2,4-DTBP was barely detectable for most Al-MCM-41 catalysts, but obviously detected for HY zeolite and textural pores-free Al-MCM-41 prepared using tartaric acid that exhibited quite low flavan selectivity (normally, 11.1% and 46% for HY and textural pores-free Al-MCM-41, respectively). As shown in Table 5.2, the traditional liquid phase alkylation catalyst, sulfuric acid was not at all effective for this reaction. Even though 2 times more protons from sulfuric acid (30 mg) was used in comparison to the acidity of 500 mg 8.8%Al-MCM-41 catalyst (~0.45 mmol/g), the conversion of 2,4-DTBP, selectivity and yield of the desired flavan were still insignificant. This may signify the importance of mild acidity observed for aluminum substituted mesoporous molecular sieve catalysts in use for this type of fine chemical alkylation. It seems that only one of those two types of acid sites on these A]- MCM-41 catalyst detected by temperature programmed pyridine desorption (c.f. Figure 5.11) is effective for this alkylation reaction. Re-use of the not-regenerated catalyst gave rise to a conversion and selectivity only ~15% lower than the values observed for the fresh catalyst. Actually, the number of weaker acid sites with a pyridine desorption peak at 217 °C still maintain the same after one cycle of the reaction. And the stronger acid sites indicated by pyridine desorption at 350 °C was no longer available for pyridine 142 Table 5.2. Alkylation of 2,4-di-tert-butylphenol with cinnamyl alcohol catalyzed by Al-substituted mesoporous molecular sieves Al-MCM41 Conv. (%) Selectivity of Yield of Catalyst 2,4-DTBP Flavan (%) Flavan (%) 30 mg H2804a 25.5 36.9 9.4 HY 13.3 11.1 1.5 Mobilb 47.2 71.8 33.9 (Sulfuric Acid) 50.0 75.2 37.6 Flaky (Citric Acid) 64.4 76.9 49.5 Flaky (Oxalic Acid) 61.9 71.2 44.1 Flaky (Tartaric Acid) 69.2 84.0 58.5 Tex-porous (Tartaric)b 69.7 79.5 55.4 Non-tex (Tartaric)c 46.3 45.9 21.3 Tex-porous (Tartaric) 65.8 83.5 54.9 Tex-porous (Citric) 64.4 79.5 5 1.0 Tex-porous (Oxalic) 63.0 79.0 50.0 *. Each catalyst contains 8.8 mol% Al except indicated samples; the reaction was carried out using 10.0 mmol of each reactant, 500 mg of catalyst , in 50 ml of isooctane at 60 °C for 6 h. a. The reaction was carried out at 90 °C for 24 h instead of 60 °C for 6 h . b. These sample contains 2.0 mol% A]. c. This sample has no textural pores (c.f, Figure 5.6). 143 adsorption because they were totally covered by polymeric aldehyde formed from cinnamyl alcohol condensation. Increasing the Al-loading leads to an increase of the number of stronger acid sites with tetrahedral coordination (c.f. Figure 5.10 and 5.11). However, the conversion and selectivity for alkylation does not improve with the increase of Al-loading (c.f. Table 5.2 for Tex-porous-tartaric-entries). This indicates that the Al- sites with mild acidity is crucial for this alkylation reaction in liquid phase. The accessibility of active sites is another crucial factor for this bulky alkylation reaction. The poor performance of HY zeolite as alkylation catalyst is simply due to the inaccessibility of the active Al-sites in the micropores even though the amount of acidity for HY is approximately equivalent to that of 8.8%Al-MCM-41. In contrast to the conventional Mobil Al-MCM-4l catalyst with particle sizes of a few micrometers, the catalyst with a thin flaky particle morphology and short pore channels exhibited much higher 2,4-DTBP conversion and flavan yield (i.e. at least by 30%). Also, Al-MCM-41 prepared using sulfuric acid in the same way as the organic acid promoters to adjust the pH two times over 2 days, exhibits primarily a micrometer sized tubular particle morphology and an only slightly improved catalytic activity. However, the conversion of 2,4-DTBP, and the selectivity and yield to the flavan all increase with increasing textural porosity. The conversion of 2,4-DTBP (i.e. 66%) for texturally porous Al-MCM—4l molecular sieves was almost as good as texturally porous Al-HMS.34 The selectivity to flavan was quite remarkable also (i.e. ~83%). 144 5.4 Conclusions Well ordered MCM-41 molecular sieves with flaky morphology and short uni- dimensional pores orthogonal to flaky-sheet surface, as well as meso-scaled fundamental particles with high textural porosity were prepared through controlled addition of organic acids such as tartaric, citric, oxalic, gluconic, lactic and salicylic acid. Structural promotion effects of the organic acids were explicitly demonstrated by the formation of 3-d hexagonal and cubic structure at 150 0C (instead of 100 °C for MCM-41) and surfactant to silicon ratio of 0.11 (instead of 0.25 for MCM-41) and H+ to NaOH ratio of 0.49 (instead of 0.75 for MCM—4l). Intrinsically, this is most likely a consequence of the variation of surfactant packing parameter g. The g value is likely varied through the surfactant-organic acid association to cause the phase change in formation of 3-d hexagonal and cubic meso—structures. On the other hand, slightly variation of surfactant packing parameters under conditions for MCM-4l preparations upon organic acid promotion leads to the formation of flaky as well as texturally porous silica molecular sieves. Follow-up in situ alumination of the flaky and textural porous MCM-41 in the presence of surfactant micelles in the pores afforded well incorporated Al-sites with over 95% are in tetrahedral coordination even for Al-loading up to 8.8mol%. However, up to a third of these Al-sites are transformed into octahedral sites upon removal of the surfactant by calcination. The Al-sites judged from 27Al-NMR and the total acidity measured from temperature programmed desorption of cyclohexylamine are in-distinguishable for similarly loaded Al-MCM-41 regardless of particle morphology. Consequently, any differences in catalysis for these Al-MCM-41 catalysts with equivalent acidity can not be attributed to any differences in the intrinsic acid siting. The number of Al-sites with 145 stronger acidity indicated by desorption of pyridine at 350 °C was proportionally increased with Al-loading, but the number of Al-sites with weaker acidity indicated by desorption of pyridine at 217 OC remained the same. The results of the studies of 2,4-DTBP alkylation with cinnamyl alcohol proves that the accessibility of the active Al-sites in flaky and textural porous Al-MCM-41 with short channel lengths is significantly facilitated. At least 30% higher conversion of 2,4- DTBP and 50% higher yield of flavan are obtained relative to Al-MCM-41 without a flaky or textural porous morphology. This alkylation reaction seems to be catalyzed by the mild acid sites since increasing the number of stronger acid sites by increasing the Al- loading does not improve the activity considerably. 146 5.5 References: l. 199°.‘19‘ 10. 11. 12. 13. 14. 15. 16. 17. Beck, J.S.; Vartuli, J.C.; Roth, W.J.; Leonowicz, M.E.; Kresge, C.T.; Schmitt, K.D.; Chu, C.T.-W.; Olson, D.H.; Sheppard, E.W.; McCullen, S.B.; Higgins, J .B.; Schlenker, J .L. J. Am. Chem. Soc. 1992, 114, 10834. Corma, A. Chem. Rev. 1997, 97, 2373; and Sayari, A. Chem. Mater. 1996, 8, 1840. Alfredsson, V.; Anderson, M.W. Chem. Mater. 1996, 8, 1141. Yang, R.; Pinnavaia, T.J.; Li, W.; Zhang, W. J. Catal. 1997, 172, 488. Monnier, A.; Schuth, F.; Huo, Q.; Kumar, D.; Margolese, D.I.; Maxwell, R.S.; Stucky, G.D.; Krishnamurty, M.; Petroff, P.; Firouzi, A.; Janicke, M.; Chmelka, B.F. Science 1993, 261, 1299. Anderson, M.W. Zeolite 1997, I9, 220. Zhangw, W.; Pinnavaia, T.J. Catal. Lett. 1996, 38, 261. Romero, A.A.; Alba, M.D.; Klinowski, J. J. Phys. Chem B 1998, 102, 123. Huo, Q.; Margolese, D. I.; Ciesla, U.; Feng, P.; Gier, T. E.; Sieger, P.; Leon, R.; Petroff, P. M.; Schuth, F.; Stucky, G. D., Nature 1994, 368, 317. Zhang, W.; Froba, M.; Wang, J.; Tanev, P.T.; Wong, J.; Pinnavaia, T.J. J. Am. Chem. Soc. 1996, 118, 9164. a. Stucky, G.D.; Monnier, A.; Schuth, F.; Huo, Q.; Margolese, D.; Kumar, D.; Krishnamurty, M.; Petroff, P.; Firouzi, A.; Janicke, M.; Chmelka, B.F. Mol. Cryst. Liq. Cryst. 1994, 240, 187; b. Tanev, P.T.; Pinnavaia, T.J. Chem. Mater. 1996, 8, 2068. a. Kloetstra, K.R.; van Bekkum, H.; Jansen, J.C. Chem. Commun. 1997, 2281; b. Luan, Z.H.; Cheng, C.F.; He, H-Y.; Klinowski, J. J, Phys. Chem. 1995, 99, 10590. Ryoo, R.; Kim, J .M. Chem. Commun. 1995, 711. Edler, K.J.; White, J .W. Chem. Mater. 1997, 9, 1226. a. Lin, H.-P.; Mou, C.-Y. Science 1996, 273, 765; b. Lin, H.-P.; Cheng, S.; Mou, C.-Y. Chem. Mater. 1998, 10, 581. Behrens, P. Angew. Chem. Int. Ed. Engl. 1996, 35, 515. Huo, Q.; Margolese, D.I.; Stucky, G.D. Chem. Mater. 1996, 8, 1147. 147 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. Israelachvili, J .N .; Mitchell, D.J.; Ninharn, B.W. J. Chem. Soc., Faraday Trans. 2 1976, 72, 1525. Hyde, S.T. Pure Appl. Chem. 1992, 64, 1617. Henriksson, U.; Blackmore, E.S.; Tiddy, G.J.T.; Soderman, O. J. Phys. Chem. 1992, 96, 3894. Fuhhop, J-H.; Helfrich, W. Chem. Rev. 1993, 93, 1565. Davis, M.E.; Chen, C.-Y.; Burkett, S.L.; Lobo, R.F. Mater. Res. Soc. Symp. Proc. 1994, 346, 831. Pinnavaia, T.J.; Zhang, W. Stud. Surf. Sci. Catal. 1998, 117, 23 (Elsevier Science B.V.). Armengol, E.; Corma, A.; Garcia, H.; Primo, J. Appl. Catal. A: General 1995, 126, 391. Huo, Q.; Leon, R.; Petroff, P.M.; Stucky, G.D. Science 1995, 268, 1324. Gallis, K.W.; Landry, C.C. Chem. Mater. 1997, 9, 2035. Flegler, S.L.; Heckman, Jr., J .W.; Klomparens, K.L. “Scanning and Transmission Electron Microscopy- An Introduction ” 1993, (W .H. Freeman and Company), New York. Ryoo, R.; Ko, C.H.; Howe, R.F. Chem. Mater. 1997, 9, 1607. Zhang, W.; Pauly, T.R.; Pinnavaia, T.J. Chem. Mater. 1997, 9, 2491. Ryoo,R.; Jun, S.; Kim, J .M.; Kim, M.J. Chem. Commun. 1997, 2225. Corma, A.; Fomes, V.; Navarro, M.T.; Perez—Pariente J. Catal. 1994, 148, 569. Hitz, S.; Prins, R. J. Catal. 1997, 168, 194. Mokaya, R.; Jones, W. Chem. Commun. 1996, 983 and J. Catal. 1997, 172, 211. Zhang, W.; Liu, Y.; Pinnavaia, T.J. J. Phys. Chem. in preparation. Armengol, E.; Cano, M.L.; Corina, A.; Garcia, H.; Navarro, M. Chem. Commun. 1995, 519. Janicke, M.T.; Landry, C.C.; Christiansen, S.C.; Kumar, K.; Stucky, G.D.; Chmelka, B.F. J. Am. Chem. Soc. 1998, 120, 6940. 148 Chapter 6 Incorporation of Structural Order into Mesoporous Silica Molecular Sieves Assembled Via a Metal Salt Mediated N°(M*X')I° Pathway Abstract The concept of incorporating structural order into mesoporous silica molecular sieves through electrostatic interactions between electrically neutral surfactant micelle and neutral inorganic precursors has been successfully illustrated. Mesoporous silica molecular sieves with well ordered domains consisted of a hybrid 3-dimensional hexagonal and spherical cubic phases (3-d-hex-cubic), as well as an uni-dimensional hexagonal structure and a wormhole structural motif were assembled using alkyl polyethylene oxide surfactant as the structure director and tetraethyl orthosilicate as the silicon source through a new counterion mediated N°(M”X')I° pathway controlled by judicial choice of a surfactant to metal ratio, and also by choice of a metal salt from the group of €002, MnCl2, NiCl2, CuCl2, ZnCl2, FeCl2, FeCl3, RhCl3, Cer and LiCl. Structural variation was strategically mediated through surfactant-metal complex formation that leads to variation of surfactant packing parameters (g = V/aol) and ultimately the inorganic meso-structures. The 3-d-hex-cubic structure is favored by “2+” metal cations such as Coz+, Ni2+ and Mn2+. Hexagonal structure is favored by both “1+” and “2+” metal cations such as Li+, C02": Cu2+, Zn“, and Fe“. No structural order is faVored by “3+” metal cations such as Cr”, Fe3+ and Rh“. It seems that 3-d-hex-cubic structure formation prefers a lower interfacial charge density with a surfactant to metal ratio in the range of 4 to 12. Hexagonal structure formation is favored by a relatively higher interfacial charge density with a lower surfactant to metal ratio in the range of l to 149 3. Wormhole structure are obtained near zero interfacial charge density with a quite high surfactant to metal ratio above 17. 6.1 Introduction Hexagonal MCM-41 and cubic MCM-48 molecular sieves have been primarily assembled via an electrostatic ST pathway relying on the lyotropic properties of the 3 Analogous hexagonal and cubic silica meso- quatemary ammonium surfactants." structures have been also prepared through liquid crystal templating using alkyl polyethylene oxide surfactants in the presence of significant amounts of mineral acid. 4 Consistently, hexagonal mesoporous silica with pore size up to 300 A was recently obtained using big triblock copolymer nonionic surfactants via an electrostatic S”X'I+ assembly mechanism in the presence of large amount of mineral acid.5 Mesoporous silica and alumina molecular sieves with wormhole channel motif, however, were cooperatively assembled via an electrically neutral N"Io pathway with H-bonding interactions between nonionic surfactant (N°) and neutral inorganic reagent (1").6'8 Although the neutral assembly pathways possess environmentally benign advantages over the electrostatic assembly pathways, our previous studies only obtained disordered wormhole structures. Actually, as a consequence of weak H-bonding assembly, both S°I° "0 and N°I° assembly lead to silica6 and alumina7 mesoporous molecular sieves with 9 wormhole channels. Nonetheless, chemical control of the surfactant assembly process for formation of desired mesoporous molecular sieves is still a challenge. Structurally well designed mesoporous molecular sieves are desirable for use as special sorbents, catalyst supports and catalysts, functional ceramics, sensors and separation-membrane precursors. 150 Therefore, it would be chemically significant and scientifically interesting to produce ordered meso-structures from neutral assembly by a facile method. It has been known that the interfacial interactions between surfactant micelles and inorganic precursors play an important role in directing the formation of meso- structures.ll Longer ranged electrostatic interaction usually affords relatively higher order of mesostructures in comparison with the shorter ranged H-bonding interaction that leads to X-ray disordered but uniformly porous meso-structures.12 Our strategy to incorporate higher structural order into mesoporous silica molecular sieves assembled using nonionic surfactant and neutral inorganic precursors is based on partial coulomb interactions at the interface between surfactant micelles and inorganic precursors without altering the fundamental feature of H-bonding assembly. Polyethylene oxide has been known to form helical complexes with Li+, Cu2+ and Zn2+.l3'15 In this work we examine the capability of alkyl polyethylene oxide surfactants to complex metal cations as an effective approach to incorporate coulomb forces at the interfaces between surfactant and hydroxylated neutral silicon species to form relatively well ordered mesoporous silica molecular sieves. Here we report a versatile novel method to prepare mesoporous silica molecular sieves. Mesostructures such as a hybrid structure consisted of 3-d hexagonal and spherically cubic (3-d-hex-cubic), an uni-dimensional hexagonal, and a wormhole-like meso-structures were synthesized through judicious choice of metal cations and control of the surfactant to metal ratio. We also purposely allow the assembly to take place in a neutral medium to maintain the H-bonding feature through a metal salt mediated neutral assembly pathway (N°(M’X')I°). The strategy involves variation of surfactant molecular 151 packing parameter g (i.e. g = V/aol) '6 through formation of metal complex which varies the effective head group area of the surfactant substantially to cause a phase change of the assembled products. 6.2 Experimental Structurally ordered mesoporous silica molecular sieves were synthesized using tetraethylorthosilicate (TEOS) as the silicon source and nonionic surfactant including Brij 56 (C16H33(EO)10H)2 Tergitol (Crz-I6H25-33(EO)8-IOH) and Triton X-100 (C3H17Ph(EO)ioOH) as the structure directors in the presence of metal salts such as CoCl2, NiCl2, MnCl2 and LiCl. The reactions were carried out at a neutral pH and a temperature between 35 to 60 °C in a reciprocal thermal bath under very gently shaking for 40 h. The product was filtered and dried in air. The obtained 3-d-hex-cubic and uni- dimensional hexagonal mesoporous silicas were denoted as MSU-C and MSU-H, respectively. The molar composition for MSU-C synthesis was 1008i : 2-7Surf : 0.2-0.8MX : 8500H2O. MSU-C was only formed in the presence of “2+” metal salts especially in favor of €002, NiCl2 and MnCl2. The molar composition for MSU-H preparation was 100Si : 2-7 Surf : l-7MX : 8500H2O. MSU-H was assembled using the same reagents, under the same reaction conditions for MSU-C, except for the surfactant to metal ratio. MSU-H was also formed using LiCl, CuCl2 and ZnCl2 in place of CoCl2. As-made material was either refluxed in ethanol for 2 h or calcinded at 600 °C for 6 h to remove the surfactant from their mesopores. 152 The powder X-ray diffraction pattern of the samples was recorded on a Rigaku Rotaflex Diffractometer using Cu K0 radiation (2.: 1.542 A). N2 adsorption and desorption isotherms were measured on an ASAP 2000 sorptometer at -l96 °C. The sample was evacuated under 10'5 torr at 150 °C over night before the measurement. 29Si NMR spectra were recorded on a Varian VXR-400S spectrometer with a 7-mm zirconia rotor, at a spin frequency of 4.2 kHz. TEM images were taken on a JEOL IOOCX with a CeB6 gun that was Operating at an acceleration voltage of 120 kV. The specimen was loaded onto a holey carbon film that was supported on a copper grid by dipping the grid into a suspension of the solid sample in ethanol. 6.3 Results and Discussion 6.3.1 Materials Characterization Figure 6.1 shows TEM images of MSU-C and MSU-H. MSU-C (top image) exhibits cavity type pores with an entrance pore diameter of ~5.5 rim and inter-pore distance of ~13.7 nm. MSU-H exhibits MCM-41-like uni-dimensional channels with a diameter of about 4.0 nm. Selected area electron diffraction patterns (insets) confirms uni-dimensional hexagonal structure for MSU-H. However, the diffraction patterns for MSU-C is complicated. No explicit structure assignment could be achieved for this diffraction pattern. To assign this diffraction pattern more accurately, a series of patterns were generated using an image-processing program, NIH Image. For the sake of comparison, the TEM image of MSU-C and its electron diffraction pattern are shown in Figure 6.2a and 6.2b, respectively. The optic diffraction pattern 2c obtained through Fourier-transformation of the image 28 shows two interdigitated 153 Figure 6.1. TEM images of calcined (600 °C) mesoporous silica molecular sieves prepared in water using the alkyl polyethylene oxide type surfactant Brij 56 (C16H33(EO)10H) as the structure director and TEOS as the silicon source in the presence of €002 salt as the structure promoter under the same reaction conditions (45 °C, 40 h reaction time) except that the surfactant to Co2+ ratios are different. The surfactant to silicon ratio was 1 to 24 for both samples, but the surfactant to metal ratio was 8.5 for the sample denoted as MSU-C (top photo) and 2.0 for the sample denoted as MSU-H (bottom photo). These images are provided as negatives for easier viewing of the pores as black spots. Pore to pore distance is constantly ~13.7 nm through the specimen for MSU- C. The insets are the electron diffraction patterns for areas imaged. 154 Figure 6.2. fihfiiifi§Ot *‘%*Q§ ‘3"Qfifidt ##‘rfiitti 3i§ittaer iUQ‘QQth 8Q.i"fitta .‘Q‘Qfifi‘ (a). Selected area TEM image of MSU-C; (b). Electron diffraction pattern of MSU-C; (c). Optic diffraction pattern obtained from the Fourier- transformation of image (a); (d). Image obtained by reverse-Fourier- transforrnation of the center tetragonal pattern in the optic diffraction pattern (c); (e). Image obtained from the total reverse-Fourier- transformation of optic diffraction pattern (c); (1). Image obtained from the reverse-Fourier-transformation of 7 diffraction spots selected from pattern (c) in such a way that these 7 spots resembling 7 spots in the real electron diffraction pattern (b). 155 Figure 6.3. A TEM segment for MSU-C showing a different orientation of the structure. The optic diffraction pattern (obtained by Fourier-transformation of arrow indicated area in the image) is attributable to both a (01—10)-3-d- hexagonal orientation and a (210) cubic orientation. These patterns are very similar to the diffraction pattern for meso—structured silica films with pores running orthogonal to the film surface, as reported by Brinker et al.” Notably, the pore to pore distance is ~13.7 nm. 156 hexagonal pattems and a central tetragonal pattern. This is indicative of a faulted hybrid structure that combines both 3-dimentional hexagonal and cubic structural symmetries. Reverse Fourier-transformation of the center tetragonal pattern gave rise to image 2d, which is obviously different from original image 2a, but it shows the same inter-pore- distance of 13.7nm as for cavity pores. Reverse Fourier-transformation of the whole optic pattern 2c generated a nice cavity image 2e with cylindrical pores around the cavity- shaped pores. This image is more like the observed image as shown in 2a. Surprisingly, if we just reverse-fourier-transform 7 reflection spots chosen from the pattern in 2c in such a way that they are arranged similar to the observed electron diffraction pattern 2b, the resulted image (21) clearly show 6 cylindrical pores in a distorted hexagon around the cavity-shaped pores. This is very much like the observed image 2a. Since only ideal 3-d hexagonal and cubic hybrids can generate diffraction pattern as shown in 2c, image 2f suggests that the obtained MSU-C is more like a distorted structural hybrid than a pure phase. These TEM images are similar to those of a meso-structured silica film reported by Brinker et al.'7 It seems that the hybrid meso—structure is dynamically evolving under certain assembly conditions. Actually, we also observed a segment of our MSU—C sample by TEM (Figure 6.3), which shows cavity pores arranged in distorted cubic fashion. The optic diffraction pattern (from Fourier-transformation of the image section pointed by an arrow) is consistent with either a (210)-cubic orientation or a (01-10)-3-d-hexagonal orientation. In addition to the hybrid ordering, there are disordered wormhole structural domains that co-exist with the well ordered structural domains within the same particle for both MSU-H and MSU-C. The disordered domains presumably resemble the ordered 157 counterparts in porous connectivity. Three-dimensional porous connectivity for the 3-d- hex-cubic structure and uni-dimensional porous conductivity for the hexagonal structure should be retained through the whole particle for MSU-C and MSU-H, respectively. The X-ray powder diffraction patterns of MSU-H further confirms the hexagonal feature by showing (100), (110) and (200) reflections (Figure 6.4). The XRD pattern for MSU-C only exhibits a single reflection peak. The intense peak at low 29 angle exhibits a d-spacing of 6.1 nm. By assuming this peak to be a cubic (210) reflection, the calculated unit cell parameter is 13.7 nm which is consistent with the inter-pore-distance observed by TEM for this sample (c.f. Figure 6.1). Lacking higher order reflections in the X-ray, the diffraction pattern also is in consistent with the presence of disordered 3-d wormhole pores. N2 adsorption and desorption isotherms of MSU-C (Figure 6.5) indeed exhibit a characteristic hysterisis loop for necked cavity pores. In reasonable agreement with the TEM measurement, the effective pore size calculated using the Horvath-Kawazoe model is 6.0 and 4.3 nm for MSU-C and MSU-H, respectively. The pore size for MSU-C is about 2 times larger than the pore diameter of a MCM-48 (i.e. 2.5 nm)18 assembled using cetyltrimethyl ammonium bromide which has the same alkyl chain length as Brij 56. A high surface area of 1103 m2/g and a pore volume of 1.06 cc/g for MSU-C strongly suggest that the over-all quality of this silica molecular sieve is comparable to a bicontinuous cubic MCM-48.18 Similar to that of MCM-41, the isotherm for MSU-H showing no hysterisis loop is consistent with uni-dimensional hexagonal pores. 158 Figure 6.4. Intensity NIM‘X‘) I ° N0 = 016113350) 10” l° = TEOS MIX. = (30612 MSU-C INF/Co” = 8.5 MSU-H 7 Wee” = 2 (hkl) duplnm dale/um 100 6.14 6.14 110 3.53 3.54 L: 3.02 3.07 | | 7 I I I T f l I I I I . . . , . . 0 2 4 6 8 10 29 (degrees) X-ray diffraction patterns for the calcined MSU-C and MSU-H samples that were used to obtain the TEM images shown in Figure 6.1. Both d- spacings, observed and calculated according to an uni-dimensional hexagonal phase for MSU-H, are provided for comparison. 159 E 800i I°=TEOS MSU-c N°Ic6*=as 17. 65 B 3 HK = 4.3 nm I~I°Ic62+ = 2 N Volume Adsorbed .5 O 9 200 0 0.2 0.4 0.6 0.8 1 Figure 6.5. N2 adsorption and desorption isotherms for calcined MSU-C and MSU-H. Pore necking hysterisis loop for MSU-C is indicated by the step at desorption branch, which is ~0.2 unit of relative pressure behind the adsorption branch. MSU-H exhibits almost identical adsorption and desorption branches signify the uniformity of uni-dimensional pores. 160 Figure 6.6. TEM images showing the wormhole (top) and 3-d-hex-cubic (lower) meso-structures assembled at C02+ : Brij56 molar ratios of 1:13 and 1 : 8.5 respectively. TEOS was the silicon source and the silicon to surfactant ratio was 17 to 1. 161 In accordance with previously reported MSU-X silicas, only a wormhole structural motif was formed when little or no metal salt was used to assembly a mesostructure at a quite high surfactant to metal ratio of 17 (Figure 6.6, top). However, when this ratio was reduced to 8.5, a nice 3—d-hex—cubic structure was reproduced (Figure 6.6, bottom). The details of this marvelous image reveals a very high resemblance to a spherical cubic surfactant meso-phase in space group of Im3m showing cavity type linkages which has been observed previously.19 Figure 6.7 shows X-ray diffraction patterns of mesostructures formed at various surfactant to Li+ ratios. Li+ was chosen because there are many reports on Li+- polyethylene oxide complexes.”21 Relatively well ordered hexagonal silica with higher 29 angle reflections was formed at a Li+ISurf ratio of 1 to 2. Disordered silicas with single reflection were formed at Li+/Surf ratio of 0 or greater than 3. Interestingly, no cubic or hybrid meso-structure was obtained under any circumstances for Li“, K+ and Na". The N2 adsorption and desorption isotherms (Figure 6.8) without hysterisis loops confirm the presence of uni-dimensional hexagonal pores for the silica with hexagonal reflections. In contrast, the N2 isotherms for the disordered silicas that exhibit a single X- ray reflection show hysterisis loops indicating significant pore necking (cavity-shaped worrnholes), and probably 3-dimensional pore connectivity. As expected, the effect of polyethylene oxide complex formation with “3+” metal cations are different from those observed for complexes with “1+” and “2+” metal ions. The experiment performed for Fe”, Cr3+ and Rh3+ indeed showed very different results from those with Fe“ and Co”. All of the silicas prepared in the presence of “3+” metal cations exhibit wormhole structure regardless of the surfactant to metal ratio. Also, their 162 810‘, 710 T 610 ‘j Li’lBri]56=5 {a 510 i a .I 3 . g 410 i Ll’lBri156=2 : - o . l:- 3104; (th) duplnm dalclnm * (100) 5.52 5.52 1 (110) 3.15 3.18 . (200) 2.77 2.75 210 : LIVBI'I]56=1 1 10 i Ll’lBri]56=0.5 j Ll’lBrl]56=0 OT'I'H'I'H'I'H'I'H'I 1 2 3 4 5 6 7 8 29 (degrees) Figure 6.7. X-ray diffraction pattern of the calcined (600 °C) meso-structured silica molecular sieves assembled using Brij 56 as the structure director and TEOS as the silicon source at a silicon to surfactant ratio of 15 to 1 under the same reaction conditions (45 °C, 40 h reaction time), except that the surfactant to Li" ratio was varied as shown. 163 s = 836 m’lg — BET 1000 4 HK = 4.8 nm * Lr/Bri] 56 = 1 Ethanol Extracted (cc/g, STP) aoOj _ 2 Sag—892mlg HK=4.3nm N Volume Adsorbed at O ‘1 HK = 5.5 nm Li’lBri] 56: 0 Ethanol Extracted II I IlfiIflfi I II 0 0.2 0.4 0.6 0.8 1 PIPo Figure 6.8. N2 adsorption and desorption isotherms of as-made meso-structured silica molecular sieves as described in Figure 6.7. The surfactant in the mesopores was removed by ethanol extraction (i.e. refluxing in ethanol for 2 h) prior to the adsorption measurement. Again, the hexagonal silica exhibits no hysterisis loop, as expected for uniform uni-dimensional pores, whereas the wormhole silicas show pore necking hysterisis loops. 164 51d MSU-X Calcined at 600 °C N?M*x1l ° n° = r ItoI 15-s-12 M‘X‘ = Metal Chloride is 310‘ g «5 II Intensity zid‘ 11047 Figure 6.9. 5d 0‘ 0 5 10 15 20 29 (degrees) X-ray diffraction patterns for calcined (600 °C) mesoporous silica molecular sieves prepared using the mixed alkyl polyethylene oxide surfactant Tergitol 15-S-12 (C12-16H25-33(EO)3-10H) as the structure director and TEOS as the silicon source under the same reaction conditions (45 °C, 40 h reaction time and a surfactant to metal cation ratio of approximately 24 to 1), but different cation valences for the promotional metal salts. The X-ray patterns show a decrease of d-spacing from 6.4 to 4.65 nm upon replacing Co2+ and Mn“ with Cr“. 165 1 000 0 N1M*X‘)l o MSU X Calcined at 600 C N° = Tergitol 15-S-12 .o = reos N°IMn2’ = as 800 .. MT = Metal Chloride E (0 :5 E 600" ‘6 ID 2 a, 400‘ o, E 2 O > z 200' Figure 6.10. N2 adsorption and desorption isotherms for mesoporous molecular sieves described in Figure 6.9. The Cr3+ mediated structure shows a pore size 2.0 nm in comparison with a pore size of 3.3 nm for the structures mediated by Co2+ and Mn“. 166 Figure 6.11. 2.5 urn SEM (top) and TEM (bottom) images of mesoporous silica prepared using Tergitol 15-S-12 in the presence of Cr3+ as a structure promoter, showing micrometer sized intergrown spherical particles with flatten surfaces on the spheres. 167 pore size is at least 7 A smaller than the silicas prepared in the presence of “1+” or “2+” metal cations. For example, the mesoporous silica formed in the presence of Cr3+ exhibits a much smaller d-spacing (i.e. 4.6 nm), in addition to smaller mesopores (~2.0 nm), in comparison with the materials promoted by Mn“ and C021 (Figure 6.9 and 6.10). Actually, all of the mesoporous molecular sieves formed with Rh“, Fe”, and 0'“ as promoters exhibit smaller pores (~2.0 nm). These same metal cations promote the formation of intergrown spherical particles, many with flatten surfaces (Figure 6.11). Notably, these mesoporous molecular sieves were synthesized under the same reaction conditions using Tergitol 15-S-12 (a mixture of C12 to C16 alkyl polyethylene oxide surfactant) as the structure director. The different pore structure and morphologies obtained for "3+" and "2+" cations strongly suggest that metal complex formation play an essential role in determining the phase and pore connectivity in the final meso-structure. 6.3.2 Discussion The possibility that metal cations may be incorporated into the silica framework is excluded, because the transition metal along with the surfactant can be easily removed by ethanol extraction. Chemical analysis confirms that no residual metal is detectable for ethanol extracted samples. Therefore, first, the metal cations are very likely complexed by surfactant head groups. Second, the assembly process in the presence of metal salt maintains neutral H-bonding interfacial interactions that are weak enough to allow removal of surfactant by solvent extraction. Consequently, MSU-C and MSU—H are very likely assembled through a new metal complex mediated N°(M*X')I° mechanism. 168 Figure 6.12 shows 29Si NMR spectrum for as-made MSU-H, a representative silica mesostructure prepared by a metal-salt mediated pathway. Typical 29 Si NMR spectra for HMS obtained by S°I° assembly and two MCM-4l samples obtained by ST and S+XT’ pathways are also included for comparison. The spectra for both MSU-H and HMS (S°I°) MSU-H (N‘lM’X) 1°) MGM-41 (560*) MGM-41 ($1) I I I I I I I I I I I I l I I I I I I I l I I I l -40 -60 -80 -100 -120 -140 -160ppm Figure 6.12. 29Si NMR spectra for as-made HMS(S°I°), MSU-H, MCM-41(S"X‘1") and MCM-41(S+l'). MSU-H was synthesized using Brij 56 as the structure director in the presence of CoCl2. HMS and MCM-41s were prepared as reported.”4 169 Scheme 1 170 MCM-41 (S+X'F) exhibit three well resolved resonance bands at -92, -100, and -110 ppm representing 2Q, 3Q and 4Q silicon cross-linking connectivity. Apparently, the spectrum for MSU-H is very much like the spectrum for MCM-4l formed though the counterion mediated S+X'1" pathway, but different from either HMS (S°I°) or MCM-41(S’T). This resemblance strongly supports that MSU-H was assembled through a new counterion- mediated N°(M*X')I° pathway. By forming complexes with metal cations, we incorporate longer ranged coulomb interactions into the surfactant micelles that cooperatively assemble with silica species to form either ordered uni-dimensional hexagonal or 3—d-hex-cubic structures. As illustrated in Scheme 1, small Li+ likely form intra-molecular complex with polyethylene oxide head group in a helical conformation.l3 The rod-like micelles formed from this surfactant- metal complex should be more rigid and easily aligned into a long-range hexagonal inorganic-organic mesophase with the help of interfacial coulomb interactions between the micelles and inorganic precursors. However, for larger Co”, which is well-known for adopting versatile coordination numbers, intra— or inter-surfactant molecular complexes are likely to form upon changing the surfactant to metal ratio (SIM). The formation of helical intra-molecular surfactant-metal complex may be more favorable as a result of positive charge repulsion at a lower SIM ratio (i.e. l/l). As a consequence of competitive complexation, inter-molecular surfactant-metal complex may be more favorable at a higher S/M ratios (i.e. 8.5/1). In the case of inter-molecular complex formation, a larger effective head group area (a0) would result a significant decrease in surfactant micelle packing parameter g to about 1/3 (g = V/aol). Therefore, in forming a thermodynamically 171 favored product, spherical micelles (at low surfactant concentration, i.e. 0.027 M) could pack into a body centered Im3m cubic meso-structure.22'23 Noticeably, an ideal Im3m cubic structure is fundamentally different (c.f. Scheme 1) from bicontinuous Ia3d cubic MCM-48 and Pm3n cubic SBA—l assembled using cetyltriethyl ammonium bromide via electrostatic 8”)!“+ pathway.23 In the case of intra-molecular complex formation, surfactant micelles with a relatively smaller head group area and larger micelle packing parameter g at about 1/2, rod-like surfactant-complex micelles may pack cooperatively into hexagonal meso—structures. Meso-structure formation is known to be driven by non-polar and polar interactions in a medium with suitable polarity. Previous work has revealed that meso- structure phases depended on the lyotropic properties of the surfactant.3 Under the cooperative assembly conditions used in the present work, the initial surfactant concentration was far less than needed to form liquid crystal phases.24 However, the surfactant molecules were eventually occluded in the framework pores as if they were in liquid crystal phases. The structure of the assembled product was primarily determined by the surfactant structure and reaction conditionsf"23 For example, cubic MCM-48 was synthesized at higher surfactant concentrations than hexagonal MCM-4l.3 3-d hexagonal SBA-2 was synthesized using double headed quaternary ammonium gemini surfactant rather than single headed surfactant.” Similarly, hexagonal, 3-d hexagonal and cubic mesoporous silicas were made using different polyethylene and polypropylene oxides block copolymer surfactants in the presence of concentrated mineral acids.26 And worm- 68 hole like mesoporous silicas were made using nonionic surfactant in neutral medium. Obviously, these strategies focused on changes of surfactant structure or surfactant 172 concentration or co-surfactant ”'28 to change the packing parameter (g) and this effectively varied the phase of inorganic meso-structure. Our approach by changing the surfactant packing parameter through chemical complex formation between surfactant head group and metal cations that ultimately change the assembled meso-structure is new. Although the possibility of forming complexes between polyethylene oxide surfactant and transition metal salt has been demonstrated for the formation of meso- structured transition metal oxides,29 the metal salt has not been found to be influential in affecting the structure of the final product. In our synthesis approach, however, the metal cations were used just as structure promoters rather than a framework precursor. In addition, none of the reported meso—structures prepared using nonionic surfactants have exhibited a TEM image similar to ours for the hybrid 3-d-hex-cubic structure. 6.4 Conclusion Modification of polyethylene oxide surfactant head groups through complexation with transition and alkaline metal cations was carried out for silica mesoprous molecular sieve synthesis. The variation in meso-structure upon variation of the valence of metal cations was quite dramatic. Mesoporous silica molecular sieves with either ordered 3-d- hex-cubic or uni-dimensional hexagonal domains were synthesized by judicious choice of metal cations. The 3-d-hex-cubic structure was favored by “2+” metal cations such as Co“, Ni2+ and Mn”. The uni-dimensional hexagonal structure were favored by both “1+” and “2+” metal cations such as Li", C021, Cu2+, Zn“, and Fe“. Wormhole structure were favored by “3+” metal cations such as Cr“, Fe3+ and Rh“. It seems that the 3-d- hex-cubic structure prefers a lower interfacial charge density with a surfactant to metal ratio in the range of 4 to 12. The uni-dimensional hexagonal structure favors a relatively 173 higher interfacial charge density with a lower surfactant to metal ratio in the range of 1 to 3. The wormhole structure favors a zero or low interfacial charge density with a quite high surfactant to metal ratio, above 17. 6.5 Reference 1. Kresge, C.T.; Leonowicz, M.E.; Roth, W.J.; Vartuli, J.C.; Beck, J .8. Nature 1992, 359, 710. 2. Beck, J.S.; Vartuli, J.C.; Roth, W.J.; Leonowicz, M.E.; Kresge, C.T.; Schmitt, K.D.; Chu, C. T.-W.; Olson, D.H.; Sheppard, B.W. J. Am. Chem. Soc. 1992, 114, 10834. 3. Vartuli, J.C.; Schitt, K.D.; Kresge, C.T.; Roth, W.J.; Leonowicz, M.E.; McCullen, S.B.; Hellring, S.D.; Beck, J.S.; Schlenker, J.L; Olson, D.H.; Sheppard, B.W. Chem. Mater. 1994, 6, 2317. 4. Attard, G.S.; Glyde, J .C.; Goltner, C.G. Nature 1995, 378, 366. Zhao, D.; Feng, J.; Huo, Q.; Melosh, N.; Fredrickson, G.H.; Chmelka, B.F.; Stucky, G.D. Science 1998, 279, 548. 6. Bagshaw, S.A.; Prouzet, E.; Pinnavaia, T.J. Science 1995, 269, 1242. 7. a. Bagshaw, S.A.; Pinnavaia, T.J. Angew. Chem. 1996, 108, 1108; Angew. Chem. Int. Ed. Eng. 1996, 35, 1106. b. Zhang, W.; Pinnavaia, T.J. Chem. Commun. 1998, 1185. 8. Prouzet, E.; Pinnavaia, T.J. Angew. Chem. 1997, 109, 533; Angew. Chem. Int. Ed. Eng. 1997, 36, 516. 9. Tanev, P.T.; Pinnavaia, T.J. Science 1995, 267, 865. 10. Zhang, W.; Pauly, T.R.; Pinnavaia, T.J. Chem. Mater. 1997, 9, 2491. 11. Behrens, P. Angew. Chem. Int. Ed. Engl. 1996, 35, 515. 12. Zhang, W. “Ph.D. Dissertation in Michigan State University”, Chapter 1, 1999. 13. Armand, MB. in Polymer Electrolyte Review - I (Eds.: J .R. MacCallum, C.A. Vincent), Elsevier Applied Science, London and New York, 1987, p l. 174 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. Jr. Bailey, F.E.; Koleske, J.V. in Alkylene Oxides and Their Polymers, Marcel Dekker, New York, 1990. McBreen, J.M.; Yang, X.-Q.; Lee, H.—S. J. Electrochem. Soc. 1996, 143, 3198. Israelachvili, J .N .; Mitchell, D.J.; Ninham, B.W. J. Chem. Soc., Faraday Trans. 2 1976, 72, 1525. Lu, Y.; Ganguli, R.; Drewien, C.A.; Anderson, M.T.; Brinker, C.J.; Gong, W.; Guo, Y.; Soyez, H.; Dunn, B.; Huang, M.H.; Zink, J .1. Nature 1997, 389, 364. Zhang, w.; Pinnavaia, T.J. Catal. Lett. 1996, 38, 261. Ober, C.K.; Wegner, G. Adv. Mater. 1997, 9, l7. Striebel, K.A.; Wen, S.J.; Ghantous, DJ. J. Electrochem. Soc. 1997, 144, 1680. Tajitsu, Y. J. Mater. Sci. 1996, 31, 2081. Hyde, S.T. Pure & Appl. Chem. 1992, 64, 1617. Huo, Q.; Margolese, D.I.; Stucky, G.D. Chem. Mater. 1996, 8, 1147. Huo, Q.; Margolese, D. 1.; Ciesla, U.; Demuth, D. G.; Feng, P.; Gier, T. E.; Sieger, P.; Firouzi, A.; Chmelka, B. F.; Schiith, F.; Stucky, G. D. Chem. Mater. 1994, 6 , 1176. Huo, Q.; Leon, R.; Petroff, P.M.; Stucky, G.D. Science 1995, 268, 1324. Zhao, D.; Huo, Q.; Feng, J.; Chmelka, B.F.; Stucky, G.D. J. Am. Chem. Soc. 1998, 120, 6024. Gallis, K.W.; Landry, C.C. Chem. Mater. 1997, 9, 2035. Chen, H.; Huang, L.; Li, Q. Chem. Mater. 1997, 9, 2685. Yang, P.; Zhao, D.; Margolese, D.I.; Chmelka, B.F.; Stucky, G.D. Nature, 1998, 396, 152. 175