.23.? >1 .sfifi .mvfimm flufim 3,. J. v. .a . ,.J3a .. . .. . gm .wmm‘... .1 [3.51.11 .11 firming ; 3% 4A ; x ..‘..... .. ‘ ....x:‘v.1~sll . 2.»:- This is to certify that the dissertation entitled Surfactant-Mediated Assembly of Crystalline Mesoporous Aluminas: Synthesis, Characterization, and Application in Hydrodesulfurization presented by Randall Wayne Hicks has been accepted towards fulfillment of the requirements for Ph.D. . Chemistry degree in fl] P J l 2 2002 Date uy ’ MSU is an Affirmative Action/Equal Opportunity Institution 0-1277l LIBRARY Michigan 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 6101 c-JCIRC/DateDuo.p65-p.15 SURFACTANT-MEDIATED ASSEMBLY OF CRYSTALLINE MESOPOROUS ALUMINAS: SYNTHESIS, CHARACTERIZATION, AND APPLICATION IN HYDRODESULFURIZATION By Randall Wayne Hicks A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 2002 ABSTRACT SURFACTANT-MEDIATED ASSEMBLY OF CRYSTALLINE MESOPOROUS ALUMINAS: SYNTHESIS, CHARACTERIZATION, AND APPLICATION IN HYDRODESULFURIZATION By Randall Wayne Hicks Aluminas are utilized in many industrial applications, including as adsorbents, abrasives, ceramics, catalysts, and catalyst supports‘. Many different phases of aluminas exist, but due to its favorable combination of surface and textural properties, the most important phase is y-alumina. Surface areas and pore volumes of conventional aluminas are typically less than 250 m2/g and 0.5 cc/g, respectively. Performance in catalytic applications is limited in part by these properties. As has been shown in Silica chemistry, the use of surfactants to aid in the assembly of a mesostructure leads to improvement in textural properties? Similar advances are anticipated in alumina chemistry, but to date, little progress has been made in this area. Since the disclosure of mesostructured aluminas in 19963, all but a couple of subsequent reports have described compositions of mesostructured aluminas with amorphous walls. This limits their thermal and hydrothermal stability, and thus their potential use in catalytic applications. The present work describes the synthesis of the first members of a new family of crystalline mesostructured aluminas prepared from the hydrolysis of aluminum sec-butoxide in the presence of either polyethylene oxide (PEO) or amine surfactants as porogens. Members of this family exhibiting the y-AI203 phase, denoted MSU-y aluminas, have surface areas, pore Sizes, and pore volumes in excess of 400 mz/g, 15 nm, and 1.5 cc/g, respectively‘. In addition, forms of these aluminas have expressed thermal and hydrothermal stability. Mesostructured boehmiteS, MSU-B aluminas, have also been obtained using Similar synthetic methods. In the synthesis of either alumina, the key step is the formation of an MSU-S/B surfactant-boehmite precursor, which is converted to the desired final product through calcination. To demonstrate their usefulness, MSU-y aluminas have been utilized as catalyst supports in dibenzothiophene hydrodesulfurization reactions. Conversions and selectivities superior to that of a commercial catalyst have been achieved. Their success in this capacity can be traced to the stability of the y- AI203 phase, the ability to effectively disperse the active catalytic component, and increased accessibility to the active sites. References (1) Misra, C. American Chemical Society Monograph, Vol. 184: Industrial Alumina Chemicals, 1986. (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, E. W.; et al. J. Am. Chem. Soc. 1992, 114, 10834-10843. (3) BagShaw, S. A.; Pinnavaia, T. J. Angew. Chem, Int. Ed. Engl. 1996, 35, 1102-1105. (4) Zhang, Z.; Hicks, R. W.; Pauly, T. R.; Pinnavaia, T. J. J. Am. Chem. Soc. 2002, 124, 1592-1593. To my wife, Jennifer, for without her sacrifice this work would not be possible. ACKNOWLEDGEMENTS There are many people to thank for their contributions toward the completion of my doctoral degree during my time at MSU. First, I express gratitude to my research advisor, Dr. Thomas J. Pinnavaia, for his support, both intellectual and financial, over the last six years. There are two specific items that need mention. First, thanks for encouraging me to stay in the Ph. D. program when I was prepared to leave with a master’s degree. I’m glad that I stayed. Second, thanks for not objecting to my desire to teach. I could have exclusively conducted research, but I feel that I’m better for having done both, even if it meant a longer tenure at MSU. So thank you for letting me do it my way. Thanks to the department, especially DrS. Paul Hunter and Kathy Hunt, for allowing me to teach over the last year. I hope that the department found it as good an experience as I did. My success in the job market this year was largely a consequence of that opportunity. Thanks to Chris Marshall at Argonne National Laboratory for the generous donation of time and facilities. The data obtained from the catalytic testing performed in his labs was crucial to the completion of this work. Thanks to former Pinnavaia group members Peter Tanev for introducing me to aluminas and Tom Pauly for valuable discussions concerning my project. Thanks to Zhaorong Zhang for our collaborations over the last 2+ years. Having also worked on it, you’re the only other person that truly knows how frustrating this project has been. To the rest of the current group members, thanks for making lab life more enjoyable. You can now all feel free to steal my lab equipment and mess up my bench as much as you’d like. Thanks to three friends that I met at MSU, Pete LeBaron, Paul Szalay, and Bill Scanlon. Graduate school would have been nearly intolerable without the good times that we had over the years. Thanks to my family for their support, even if they weren’t sure exactly what I was doing here. To all my parents, Charles (Linda) Hicks and Neva (Ron) Penix, thanks for preparing me for this experience. Thanks too, to my mother-in-law, Leone Beck, who has always provided words of encouragement. Thanks to my brother, Ryan, and my other friends back home, Rick & Jess, Mark & Shelley, Vern, Frog, and John for support and much needed stress relief. I only wish that we had more time to spend together Since I’ve been here. Most importantly, thanks to my wife and my buddy, Jennifer. I am eternally grateful for the love and support that you have given me. I can’t express here in words what that has meant. We’ve both had to make sacrifices so that I could finish what I started, more so for you, particularly over this last year. But it will all be worthwhile in the end. And soon, it‘wiIl be our time again. vi TABLE OF CONTENTS LIST OF TABLES .......................................................................................... x LIST OF FIGURES ........................................................................................ xi LIST OF ABBREVIATIONS ........................................................................... xix Chapter 1: Introduction .............................................................................. 1 1.1. Classification of Aluminas ........................................................... 1 1.2. Transition Aluminas ..................................................................... 5 1.3. Hydrodesulfurization .................................................................... 9 1 .4. Mesostructures ............................................................................ 18 1.5. Mesostructured Aluminas ............................................................ 21 1.6. Stabilization of Aluminas ............................................................. 23 1.7. Objectives and Rationale ............................................................ 26 1.8. References .................................................................................. 28 Chapter 2: Synthesis and Characterization of Mesostructured Boehmite ............................................................................... 31 2.1 . Introduction .................................................................................. 31 2.2. Experimental ............................................................................... 34 2.2.1. Synthesis ....................................................................... 34 2.2.2. Characterization ............................................................ 35 2.3. Results and Discussion ............................................................... 36 vii 2.3.1. Formation of MSU-S/B .................................................. 36 2.3.2. General Properties of MSU-B Aluminas ........................ 39 2.3.3. Peptization ..................................................................... 54 2.4. Mechanistic Aspects of MSU-B Alumina Synthesis .................... 55 2.5. Conclusions ................................................................................. 58 2.6. Future Directions ......................................................................... 59 2.7. References .................................................................................. 61 Chapter 3: Synthesis and Characterization of Mesostructured y- AI203 ........................................................................................ 63 3.1. Introduction .................................................................................. 63 3.2. Experimental ............................................................................... 65 3.2.1 . Synthesis ....................................................................... 65 3.2.2. Characterization of MSU-ry Aluminas ............................. 67 3.3 Results and Discussion ................................................................ 68 3.3.1. MSUq Aluminas from PEO Surfactants ........................ 68 3.3.2. MSU-y Prepared from Amine Surfactants ...................... 85 3.3.3 Stability of MSUJY Aluminas ............................................ 93 3.4. Conclusions ................................................................................. 103 3.5. Future Directions ......................................................................... 105 3.6. References .................................................................................. 106 viii Chapter 4: Hydrodesulfurization Over MSU-y Supported Catalysts ....... 108 4.1. Introduction .................................................................................. 108 4.2. Experimental ............................................................................... 1 10 4.2.1. Sample Preparation ....................................................... 110 4.2.2. Reactor Design and Experimental Conditions ............... 113 4.2.3. Characterization ............................................................ 1 15 4.3. Results and Discussion ............................................................... 116 4.3.1. Catalytic Reactions ........................................................ 116 4.3.2. Structural and Textural Properties of MSUJY Catalysts. 119 4.3.3. Relationships Between Catalyst Structure and Performance .......................................................... 128 4.4. Conclusions ................................................................................. 139 4.5. Future Directions ......................................................................... 139 4.6. References .................................................................................. 142 Chapter 5: Concluding Remarks ............................................................... 144 Table 2.1. Table 3.1 . Table 3.2. Table 3.3. Table 3.4. Table 3.5. Table 4.1 . Table 4.2. LIST OF TABLES Synthetic parameters and textural properties of MSU-B Aluminas .................................................................................. Synthetic parameters and textural properties of MSU-y aluminas prepared from aluminum sec-butoxide and PEO surfactants ............................................................................... Comparison of textural properties of calcined MSUJY formed at 500 °C and calcined MSU-6 alumina formed at 800 °C ............................................................................................. Textural properties of MSU-y aluminas prepared through the calcination of MSU-S/B surfactant-boehmite compositions and MSU-B boehmites at 500 °C ............................................ Textural properties of doped MSU-y aluminas prepared from aluminum sec-butoxide and P84. In each synthesis, the dopant was administered as a metal salt which was dissolved in the surfactant solution prior to addition of the aluminum sec-butoxide ............................................................ Textural properties of MSU-y aluminas prior to and after treatment at 600 °C in a 20% steam atmosphere for 5 hr ....... Results of HDS reactions over MSUJy catalysts exhibiting distinctly different morphologies .............................................. Textural properties of MSU-y alumina HDS catalysts .............. 44 72 87 92 94 104 117 127 Figure 1.1. Figure 1.2. Figure 1.3. Figure 2.1 . Figure 2.2. Figure 2.3. Figure 2.4. Figure 2.5. LIST OF FIGURES Family of alumina compounds ................................................. Transformation sequence observed upon the thermal dehydration of aluminas (adapted from Wefers and Misra) ..... Activity of transition metal sulfides in DBT HDS plotted as a function of periodic position of the metal (adapted from Pecoraro and Chianelli) ........................................................... Cross-sectional (upper) and three-dimensional (lower) depiction of the boehmite double-layer. The thickness of the double layer is ~ 12 A .............................................................. Schematic of interactions leading to formation of MSU-S/B. Each slab represents one boehmite double layer. Lines on the boehmite layers represent hydroxyls forming hydrogen bonds to the surfactant micelles. Areas of overlapping boehmite layers indicate cross-linking through condensation of adjacent hydroxyls ............................................................... Representative low angle XRD patterns for the as- synthesized surfactant-boehmite composite, denoted MSU- S/B, prepared from aluminum sec-butoxide and dodecylamine in an alcoholic/aqueous medium as described by line 3 in Table 2.1 and the surfactant-free analog, MSU-B, obtained by calcination of the MSU-S/B composite at 325 °C for 10 hr ................................................................................... Representative wide angle XRD patterns for MSU-S/B and MSU-B aluminas Shown in Figure 2.3. Crystallographic planes contributing to the peaks are noted .............................. Nitrogen adsorption — desorption isotherms and BJH adsorption pore size distributions (inset) for MSU-B aluminas differing in surfactant concentration. The percentage given corresponds to the mass percent of dodecylamine in the surfactant solution. The three isotherms, offset by 200 cc/g for clarity, correspond from top to bottom to samples described in lines 3, 8, and 19 in Table 2.1. PSD curves are displayed in the same order as the isotherms ......................... xi 4 6 12 33 38 40 41 43 Figure 2.6. Figure 2.7. Figure 2.8. Figure 2.9. Figure 2.10. Nitrogen adsorption - desorption isotherms and BJH adsorption pore Size distributions (inset) for MSU-B aluminas prepared from aluminum sec-butoxide and dodecylamine at ambient temperature in alcoholic/aqueous solvents of varying ethanol content. The three isotherms, offset by 200 00/9 for clarity, correspond to MSU-B aluminas described in lines 1-3 of Table 2.1. PSD curves are displayed in the same order as the isotherms ............................................................. Nitrogen adsorption — desorption isotherms and BJH adsorption pore Size distributions (inset) for MSU-B aluminas prepared from aluminum sec-butoxide and dodecylamine at different temperatures. The four isotherms, offset by 200 cc/g for clarity, correspond to MSU-B aluminas described in lines 3-6 of Table 2.1. PSD curves are displayed in the same order as the isotherms ............................................................. Nitrogen adsorption - desorption isotherms and BJH adsorption pore size distributions (inset) for MSU-B aluminas prepared from the hydrolysis of aluminum sec-butoxide in the presence of different amine surfactants as porogens. The Six isotherms, offset by 200 cc/g for clarity, correspond to MSU-B aluminas described in lines 7-12 of Table 2.1. PSD curves are displayed in the same order as the isotherms ................................................................................. Bright field TEM images (A and C) and electron diffraction patterns (B and D) of MSU-B materials prepared from aluminum sec-butoxide and dodecylamine as described in Table 2.1, (A) line 8 and (B) line 3. The open framework results from the random orientations of fundamental particles. Electron diffraction patterns, (B) and (D), corresponding to bright field images (A) and (C), respectively Show diffuse rings, Signifying the polycrystalline nature of the material. The reflections contributing to the pattern are labeled in (D) ........................................................................... 27Al NMR Spectrum of MSU-B prepared from aluminum sec- butoxide and dodecylamine prepared as described in Table 2.1, line 3. The resonance at 7 ppm is consistent with the structure of boehmite, in which all aluminum ions are in octahedral environments ......................................................... xii 46 47 49 52 53 Figure 2.11. Figure 3.1. Figure 3.2. Figure 3.3. Figure 3.4. Wide angle XRD patterns of MSU-S/B prepared from aluminum sec-butoxide and dodecylamine at 20 °C Showing evolution of the boehmite phase over time. The presence of weak reflections assignable to boehmite after only 1 minute Show that the formation of MSU-B does not proceed through an MSU-X type of intermediate with amorphous framework walls ......................................................................................... Representative wide angle powder XRD patterns of an MSU- S/B surfactant-boehmite composition made from aluminum sec-butoxide and the PEO surfactant P84 under reaction conditions described in Table 3.1, line 3 and the corresponding MSUJy alumina prepared from calcination of the as-synthesized MSU-S/B at 500 °C for 4 hr ...................... Nitrogen adsorption-desorption isotherm and BJH adsorption pore Size distribution (inset) of an MSU-Iy alumina prepared from aluminum sec-butoxide and the PEO surfactant P84 as described in Table 3.1, line 3. Included for comparison is the same data for an alumina prepared without surfactant as described in Table 3.1, line 20. lsotherms are offset 200 cc/g for clarity and PSD curves are displayed in the same order as the isotherms ...................................................................... 27Al NMR of calcined (500 °C) MSU-y alumina prepared from aluminum sec-butoxide and P84 as described in Table 3.1, line 3. The two resonances at 6.7 ppm and 72 ppm are assigned to aluminum in octahedral and tetrahedral coordination, respectively. The peaks are in the approximate 75:25 ratio consistent with y-AI203 ........................................... Bright field TEM image of (A) calcined (500 °C) MSU-y alumina prepared from aluminum sec-butoxide and P84 as described in Table 3.1, line 3. (B) Lower magnification image of the same product and the corresponding electron diffraction pattern, with diffuse rings assignable to the [311], [400], and [440] crystallographic planes. (C) Dark field image of the same particles as in (B), but formed from an ~79 arc of the [440] diffraction ring. The crystal orientations contributing to the [440] ring appear aS bright sports in the image ............. xiii 57 69 71 74 76 Figure 3.5. Figure 3.6. Figure 3.7. Figure 3.8. Figure 3.9. Nitrogen adsorption-desorption isotherms of MSU-y aluminas prepared from aluminum sec-butoxide and P84 as described in Table 3.1, lines 3 and 6-10 showing the effect of solvent polarity on the textural properties. The volume percent of 2- butanol in the solvent is labeled for each isotherm. Isotherms are offset 300 cc/g for clarity. The BJH adsorption pore size distributions (inset) are in the same order as the isotherms... Nitrogen adsorption-desorption isotherms and BJH adsorption pore size distributions (inset) for MSUJY aluminas prepared from aluminum sec-butoxide and P84 as described in Table 3.1, lines 3 and 11-14 showing the effect of surfactant concentration (mass P84 / mass solvent) on the textural properties. The mass percent of P84 is labeled for each isotherm. Isotherms are offset 300 cc/g for clarity. The PSD curves are displayed in the same order as the corresponding isotherms ......................................................... Nitrogen adsorption-desorption isotherms and BJH adsorption pore Size distributions (inset) for MSU-ry aluminas prepared from aluminum sec-butoxide and differing PEO surfactants as described in Table 3.1, lines 3 and 11-14. Isotherms are offset 200 cc/g for clarity. The PSD curves are displayed in the same order as the corresponding isotherms. Nitrogen adsorption-desorption isotherm and wide angle powder XRD pattern (inset) of a calcined (800 °C) MSU-6. The initial MSU-S/B surfactant-boehmite intermediate was prepared from aluminum sec-butoxide and the PEO surfactant P84 under conditions described in Table 3.1, line 13 ............................................................................................. TEM image of MSU-y prepared through the calcination of MSU-B boehmite at 500 °C for 4 hr. The initial MSU-S/B surfactant-boehmite composition was prepared from aluminum sec-butoxide and dodecylamine (DDA) at a reaction stoichiometry of AlzDDAzEtOHzH20 = 5:1:68.0:73.4 and a reaction temperature of 20 °C. The surfactant free MSU-B boehmite phase was obtained by calcining the MSU- S/B precursor at 325 °C for 10 hr. The electron diffraction pattern (inset) demonstrates the crystallinity of the framework of the mesostructure, as indicated by the diffuse rings attributable to the [311], [400], and [440] crystallographic planes of Y-AIan ............................................ xiv 80 82 83 86 9O Figure 3.10. Figure 3.11. Figure 3.12. Figure 3.13. Figure 3.1 4. Nitrogen adsorption-desorption isotherms and BJH adsorption pore size distributions (inset) for MSU-y aluminas prepared through the calcination (500 °C) of amine-mediated MSU-S/B surfactant-boehmite precursors (samples A and C) and surfactant free MSU-B precursors (samples B and D). The AlzDDAzEtOHzH20 reaction stoichiometry was 5:1 :68.0:73.4 for all samples. Reaction temperature was 20 °C for samples A and B and 45 °C for samples C and D. Isotherms are offset 300 cc/g for clarity. The PSD curves are displayed in the same order as the corresponding isotherms. Wide angle powder XRD patterns of MSU-y aluminas prepared from aluminum sec-butoxide and P84: pristine and doped with either lanthanum or cerium. Peaks marked with an asterisk are attributable to a cerium aluminate phase, CeAl03. The doped MSU-y aluminas were prepared through the calcination (500 °C) of doped MSU-S/B precursors prepared as described in Table 3.1, line 13 ............................ Nitrogen adsorption-desorption isotherms and BJH adsorption pore Size distributions (inset) of MSUJY aluminas: pristine and doped with either lanthanum or cerium. The doped MSUJY aluminas were prepared through the calcination (500 °C) of doped MSU-S/B precursors prepared as described in Table 3.1, line 13. lsotherrns are offset 400 cc/g for clarity. The PSD curves are in the same order as the corresponding isotherms ......................................................... TEM images of MSU-y aluminas prior to (A,C,E) and after steaming (B,D,F) at 600 °C for 5 hr in a 20% steam atmosphere. Images A and B are for pristine MSUq. Images C, D and E, F are for MSUq doped with 5% La3+ and 5% Ce‘“, respectively. The MSUJY aluminas were prepared as described in the caption to Figure 3.11. The scale bar represents 100 nm in all figures .............................................. Nitrogen adsorption-desorption isotherms and BJH adsorption pore size distributions (inset) of pristine and doped MSU-y aluminas after steaming for 5 hr at 600 °C in a 20% steam atmosphere. Samples were prepared as described in the caption to Figure 3.11. Isotherms are offset 300 cc/g for clarity. The PSD curves are in the same order as the corresponding isotherm ................................................ . 91 96 97 100 101 Figure 3.15. Figure 4.1. Figure 4.2. Figure 4.3. Figure 4.4. Figure 4.5. Wide angle powder XRD patterns of pristine and doped MSU-y aluminas after steaming for 5 hr at 600 °C in a 20% steam atmosphere. Peaks marked with an asterisk are attributed to CeAlOa, cerium aluminate. Samples were prepared as described in the caption to Figure 3.11 ............... Schematic of the HDS reactor employed in the study. The liquid feed of dibenzothiophene in hexadecane was introduced to the gas feed, a mixture of hydrogen and nitrogen. The two leads are passed through the preheater (set to 350 °C), where all feed is vaporized. Then the feed reactants are passed over the catalyst bed at 400 °C. Liquid products are condensed and collected while gas products are scrubbed in 2M NaOH prior to venting .............................. Reaction pathways of DBT in HDS reactions. The upper pathway involves direct desulfurization leading to the formation of biphenyl (BP). The lower pathway involves a hydrogenation step prior to desulfurization, leading to the formation of cyclohexylbenzene (CHB) ................................... Wide angle XRD patterns for lath-12 MSUJY alumina taken at various stages of the catalytic testing process. The y-Al203 phase is evident throughout the process and the lack of peaks assignable to M082 indicate that the active catalyst is present in small domains and is well dispersed on the suppon ..................................................................................... Wide angle XRD patterns for scaffold-12 MSU-y aluminas taken at various stages of the catalytic testing process. The y—AI203 phase is evident throughout the process and the lack of peaks assignable to M082 indicate that the active catalyst is present in small domains and is well dispersed on the suppon ..................................................................................... Wide angle XRD patterns for the commercial Crosfield catalyst, taken at various stages of the catalytic testing process. The decrease in intensity of the peaks assignable to y-AI203 in the pattern of the sulfided sample may be due to the coating of the alumina surface by small domains of M082. Peaks from y—Al203 and M032 in the pattern of the Spent sample indicate that M082 particles may aggregate sufficiently to cause scattering and expose the surface of the suppon ..................................................................................... 102 114 118 120 121 122 Figure 4.6. Figure 4.7. Figure 4.8. Figure 4.9. Figure 4.10. Nitrogen adsorption - desorption isotherms and BJH adsorption pore size distributions (inset) for lath-12 MSUay catalysts taken at various stages of the catalytic testing process. The pore volume steadily decreases as the catalyst iS loaded, sulfided, and spent. Isothems are offset 100 cc/g for clarity and PSD curves are displayed in the same order as the corresponding isotherms .............................................. Nitrogen adsorption - desorption isotherms and BJH adsorption pore size distributions (inset) for scaffold-12 MSUJy catalysts taken at various stages of the catalytic testing process. The pore volume decreases slightly as the catalyst is loaded, sulfided, and spent while the pore size is increased and the distribution broadened over the same conditions. Isothems are offset 0, 100, 100, and 200 cc/g from bottom to top for clarity and PSD curves are displayed in the same order as the corresponding isotherms ................. Nitrogen adsorption — desorption isotherms and BJH adsorption pore size distributions (inset) for the commercial Crosfield catalyst taken at various stages of the catalytic testing process. The pore volume decreases slightly as the catalyst is sulfided and spent. However, the pore Size and distribution remain relatively unaffected over the same conditions. Isothems are offset 100 cc/g for clarity and PSD curves are displayed in the same order as the corresponding isotherms ................................................................................. Bright field TEM images of the scaffold-12 MSU-ry catalyst (A) after loading and (B) after using in an HDS reaction for 3 hr at 400 °C and 400 psi. The scale bar represents 100 nm. The inset iS an approximately 30 nm x 30 nm area of the catalyst. That the framework of the alumina is largely unaffected by the conditions of the HDS reaction provides evidence of its stability ............................................................. Bright field TEM images of the lath-12 MSU-y catalyst (A) after loading and (B) after using in an HDS reaction for 3 hr at 400 °C and 400 psi. The scale bar represents 100 nm. The inset is an approximately 30 nm x 30 nm area of the catalyst. The local pore structure (inset) is compromised to some degree, as is Shown by the decreased void space in (B) as compared with (A). The long range order of the alumina laths is lost after use in the HDS reaction .................. 124 125 126 130 132 Figure 4.11. Figure 4.12. Figure 4.13. Bright field TEM images of the commercial Crosfield catalyst (A) after loading and (B) after using in an HDS reaction for 3 hr at 400 °C and 400 psi. The scale bar represents 100 nm. The inset is an approximately 30 nm x 30 nm area of the catalyst. Although the local pore structure (inset) appears to withstand the conditions encountered in the HDS reaction, it appears that there is some aggregation of the fundamental particles after use .................................................................... The Daage-Chianelli rim-edge model of the M082 catalyst particle. Sulfur removal takes place on rim and edge Sites while hydrogenation of the aromatic ring iS restricted to rim Sites only .................................................................................. Influence of morphology on the ratio of rim sites to total sites available on the edges of M082 particles. While increasing the stack height leads to a Slight increase in the overall number of edge Sites, the more pronounced effect is the decrease in the relative number of rim sites ............................ xviii 134 136 138 LIST OF ABBREVIATIONS 4H-DBT Ring hydrogenated analog of DBT, dibenzothiophene BET Brunauer Emmett Teller BJH Barrett Joyner Halender BP Biphenyl S-BuOH sec-butanol, 2-butanol CHB Cyclohexylbenzene DBT Dibenzothiophene DDA Dodecylamine EO Ethylene oxide EPA Environmental Protection Agency EtOH Ethanol GC-MS Gas chromatography - mass spectrometry h-bond Hydrogen bond HDA Hexadecylamine HDS Hydrodesulfurization HMS Hexagonal mesoporous Silica HOMO Highest occupied molecular orbital M418 Mobil family of mesoporous SilicaS MOM-41 Mobil composition of matter 41 MSU-tS Mesoporous alumina comprised of a 0-AI203 framework MSU-y Mesoporous alumina comprised of a y-AI203 framework xix MSU-B MSU-SIB MSU-X nm NMR OA P/Po PEO PO ppm PSD s-BuOH surf TalA Tal TetA Tal TriA TEM TMS v/v XRD Mesoporous alumina comprised of a boehmite framework Surfactant-boehmite composition, precursor to MSU aluminas Wormhole mesoporous aluminas (or Silicas) nanometer Nuclear magnetic resonance Octylamine Relative pressure, P = pressure, P0 = saturation pressure Polyethylene oxide Propylene oxide parts per million Pore Size distribution sec-butanol, 2-butanol Surfactant Tallow amine Tallow tetraamine Tallow triamine Transmission electron microscopy Transition metal sulfide volume ratio X-ray diffraction Chapter 1 Introduction 1 .1. Classification of Aluminas Aluminas, a class of compounds comprised of aluminum, oxygen, and often hydrogen, are wonderfully interesting and diverse. The name alumina iS broadly applied to all members of this group that includes the aluminum hydroxides, aluminum oxide hydroxides, and aluminum oxides. The properties of these compounds vary greatly, from powders that are relatively soft to larger crystals of extreme hardness, inert to readily reactive, free flowing to viscous, catalytically active to inactive. Regardless of nature, it is interesting that all of these aluminas, when heated to temperatures high enough and for periods of time long enough, are eventually converted to d-alumina or corundum, the only stable form of alumina. Aluminas are commercially important chemicals. The world’s primary source of aluminum is from the mineral bauxite, an impure hydrous ore of aluminum and other metals. Bauxite is refined to useful alumina chemicals through the Bayer process. The ore is first treated with hot concentrated sodium hydroxide to extract the aluminum hydroxides and remove the impurities. Crystalline aluminum hydroxide is then precipitated, washed, and thermally dehydrated to produce aluminum oxide‘. While most of the alumina produced is used in the production of metallurgical grade aluminum, a large amount is devoted strictly to alumina chemicals. According to the International Aluminum Institute, over 40 million tonnes of alumina was produced in 20012. Not surprisingly, aluminas find uses in a variety of applications including as fillers in plastics and paper, fire retardants, adsorbents, abrasives, ceramics, and perhaps most importantly, as catalysts and catalyst supports. As mentioned above, a range of composition is observed within the family of aluminas. They may be crystalline or amorphous. The crystalline aluminas are aluminum hydroxides, AI(OH)3, aluminum oxide hydroxides, AlOOH, and aluminum oxides, A|203. There are three forms of the trihydroxides: gibbsite, bayerite, and nordstrandite. Of these, gibbsite is the most common and is the product of the Bayer process. In its structure, aluminum ions occupy 2/3 of the octahedral vacancies in the close packed lattice of hydroxide ions. Hydroxide ions are aligned so that the sequence of layer stacking is AB...BA...AB...BA. Bayerite is not found in nature, but has been prepared synthetically. Its structure is similar to that of gibbsite, but the layer sequence is AB...AB...AB. The third trihydroxide, nordstrandite, has a structure that is a composite of the previous ones. Alternating double layers of gibbsite and bayerite are stacked to form the structure‘. The best method to distinguish between these different crystalline phases is x-ray diffraction (XRD). In addition to the crystalline trihydroxides, amorphous AI(OH)3 also exists. Two other members of the alumina family are aluminum oxide hydroxides, AlOOH, boehmite and diaspora. Boehmite occurs as a component of bauxite and can also be synthesized through neutralization of aluminum salts or hydrolysis of aluminum alkoxides around the boiling point of water". The boehmite structure is comprised of extended chains of AlOOH dimers. The chains form a double layer in which hydroxyls in one layer are positioned over the depressions between adjacent hydroxyls in the other layer. The oxygen ions are arranged in a cubic close packed lattice. Diaspore is found in nature associated with geologically older bauxite deposits and metamorphic rocks and it is therefore believed that high pressures and temperatures are required for its formation. Its structure is Similar to that of boehmite in that the basic unit is the chain of AlOOH dimers, however, the oxygen atoms are arranged in a hexagonal close packed lattice. In denoting aluminas, the common crystallographic designation assigns a prefix of a- to the hexagonal phases and a prefix of y— to the cubic phases. Therefore, boehmite is often represented as y—AIOOH and diaspore as a-AIOOH. The only thermodynamically stable oxide of aluminum is corundum, a-Alzos. It can be prepared by several methods, the most common being thermal degradation of other aluminas at temperatures over 1000 °C. Corundum iS also found in nature in igneous and metamorphic rocks. The 0- designation is consistent with its hexagonally close packed oxygen lattice in which aluminum ions occupy 2/3 of the octahedral holes. ln gem quality crystals of corundum, trace impurities of chromium lead to the formation of rubies while iron and titanium impurities lead to sapphires. The family of these alumina compounds is summarized in Figure 1.1. Aluminas Hydroxides Oxides Gibbsite, y-AI(OH)3 Aluminum Oxide, o-AI203 Bayerite, d-AI(OH)3 (Corundum) Nordstrandite, A|(OH)3 V Oxide Hydroxides Boehmite, v-AIOOH Diaspore, a-AIOOH Figure 1 .1. Family of alumina compounds. 1.2. Transition Aluminas AS mentioned previously, all aluminas eventually are converted to corundum upon heating at high temperatures. The appropriately named transition aluminas are disordered crystalline phases formed at temperatures of 250 — 800 °C during the transformation to a-Al203 at approximately 1100 °C. At least seven different transition aluminas have been reported: p (rho), X (chi), n (eta), y (gamma), K (kappa), B (theta), and 0 (delta). The particular phase observed depends upon both the precursor and the temperature to which that precursor is heated. Gibbsite is converted first to X at a temperature of 350 °C, then to K around 750 °C while the other trihydroxides proceed though an n phase at lower temperatures, then to 6 at temperatures over 800 °C. All of them are transformed to 0 beyond 1100 °C. The oxide hydroxide phases exhibit different transformation sequences than those of the trihydroxides. Boehmite is converted to v around 500 °C, to 0 above 750 °C, and then to 6 above 900 °C before reaching a above 1000 °C. The behavior of diaspore is interesting in that it converts directly to a at a temperature of 500 °C. The similarity in the hexagonal packing of oxygen atoms in the dense structures of these phases, requiring therefore only minimal rearrangement upon transformation, is thought to be responsible for this observation. Within the group of transition aluminas, the nearly anhydrous 6, K, and 0 phases are more ordered than the other forms, as their formation requires higher temperatures‘. The transformation sequences are summarized in Figure 1.2. GibbsiteI—pl Chi |——+ | Kappa |Alpha | |Boehmite| —> | Gamma | Delta |Theta |Alpha] | Diaspore | —'—P | Alpha | 100 200 300 400 500 600 700 800 900 1000 1100 Temperature °C Figure 1.2. Transformation sequence observed upon the thermal dehydration of aluminas (adapted from Waters and Misraa). Despite the number of different phases, the structures of the transition aluminas are fairly Similar with respect to each other. This in part explains the observed transformation sequence. X-Al203 exhibits a highly disordered hexagonal packing of oxygen anions while the high temperature form of k- Al203 is also hexagonal but more ordered. n- Al203 has a tetragonally distorted spinel structure in which 21 1/3 of the 24 cation vacancies are occupied by aluminum ions. The 6- AI203 structure is similar to that of the lower temperature n form. The structure of y- Al203 is also of the spinel type, but it can be distinguished from that of the n form by its Significantly larger tetragonal distortion. 6- Al203 is Similar to v, but it exhibits slightly greater long-range orders. In general, the identification of transition aluminas is best performed through XRD and 27Al nuclear magnetic resonance (NMR). Each transition alumina has a slightly different XRD pattern, the higher temperature forms exhibiting sharper peaks than the others. 27Al NMR can be used to distinguish the distribution of aluminum ions in the octahedral and tetrahedral vacancies of the oxygen lattice. As the transition aluminas are disordered phases, dehydration upon heating and subsequent rearrangement of the lattices lead to the formation of porosity and internal surface area. While there iS a loss of mass upon heating, the particle sizes remain the same, resulting in the development of porosity. In the dehydration of the trihydroxides, surface areas can reach as high as 350 m2/g at lower temperatures before rapidly and continually declining toward the conversion to a- Al203. For the oxide hydroxides, there iS also an initial increase in porosity leading to surface areas of 90-100 m2/g at a temperature of 500 °C that decreases to 15-20 m2/g upon heating at about 550 °C‘. Upon the thermal dehydration of aluminas, the surfaces of the transition aluminas are altered extensively. The condensation of surface hydroxyls and their subsequent removal as water create a rough surface, exposing coordinatively unsaturated aluminum ionS and leaving the remaining hydroxyl ions in various environments. This leads to acidic and basic Sites on the surface of the alumina. While they are also used as adsorbents in liquid drying, water purification, and chromatographic columns, the combination of textural properties (pore volume and surface area) and surface properties leads to activity in catalytic applications. Consequently, transition aluminas are also referred to as activated aluminas. Of all of the different phases, y-Alea is perhaps the most important and most widely utilized transition alumina. It is formed upon thermal dehydration of boehmite at temperatures ranging from 450 to 750 °C as described by the following equation: 2 AlOOH -> Al203 + H20 y- AI203 has been used as a catalyst in alcohol dehydration reactions to form alkeneS and ethers and in isomerization reactions. y- A|203 has also proven to be extremely useful as a catalyst support in numerous reactions due to its combination of heat resistance, chemical inertness, and favorable textural and surface properties. y- AI203 can be produced with surface areas ranging from 185 — 250 m2/g with pore volumes typically around 0.20 cc/g"3. Two primary methods are employed in the preparation of supported catalysts, impregnation and coprecipitation. In the incipient wetness impregnation method, a salt solution of the active species iS deposited on the alumina and then thermally degraded to form the active catalyst. In the coprecipitation method, salt solutions of both the aluminum and active component are precipitated Simultaneously, again followed by thermal degradation to the produce the active catalyst. High surface area is a desirable property because a larger amount of active catalyst can be loaded onto the support. Many of the reactions in which alumina supported catalysts are used are important in the petroleum industry. 1.3. Hydrodesulfurization One such reaction is the hydrodesulfurization (HDS) reaction, a crucial component of petroleum refining. Heteroaromatic sulfur organic molecules, common components of crude oil deposits, must be removed before fuels can be cleanly consumed. Fuels high in sulfur, when burned, produce acidic sulfur oxide species that lead to the production of smog and acid rain. Federal regulations put forth by the EPA are calling for a 97% reduction in sulfur content in fuels, from 500 to 15 ppm sulfur, by the year 2010‘. At the same time that these constraints are put into place, lower sulfur content crudes are being depleted, making it necessary to rely upon feedstocks that contain larger amounts of sulfur. In addition, the presence of sulfur has a detrimental effect on other catalysts involved in downstream catalytic processing of feeds. Therefore, development of improved methods for sulfur removal is pertinent both environmentally and economically. In the HDS reaction, sulfur is removed from organic molecules as H28. Many studies of HDS processes are performed using thiophene or dibenzothiophene (DBT) as substrates. These molecules, as well as substituted and polyaromatic analogs, are representative of the sulfur containing compounds found in crude oil stocks. Sulfur removal in these larger molecules is especially challenging as steric effects come into play. Catalytic studies involving dibenzothiophene are therefore better indicators of potential performance using actual oil feedstocks. Reactions are usually performed in a catalytic reactor at temperatures ranging from 350 — 400 °C and pressures of 400 psi. The catalysts employed in HDS reactions are transition metal sufides (TMS). Many TMS Show activity in HDS reactions, but because of their considerably lower cost, the HDS catalysts typically employed in industrial processes are either molybdenum or tungsten sulfides, combined with either nickel or cobalt sulfides, loaded on a y— Al203 support. Some classic works by Chianelli and others""14 over the last 20 years explain why these catalysts are effective, making them the preferred choice for use in the petroleum industry. Much work has been done toward Characterizing transition metal sulfide catalysts. Several structure/function relationships have been derived as models for catalytic activity and selectivity have been proposed to explain the nature of these TMS materials. The primary factor in determining the utility of a TMS catalyst is in the electronic effect that is observed in these catalysts. That is, 10 correlations to and trends in activity in HDS can be made based upon the position of the metal in the periodic table. Pecoraro and Chianelli measured the activity of TMS catalysts in DBT HD85. When the activity is plotted against the periodic position, a “volcano” plot is observed (Figure 1.3), where a maximum in the activity is seen for Group VIII metalSS. This trend was explained based upon the metal-sulfur bond strengths. As first described by Sabatier in 1911, in order to be an effective catalyst in a given reaction, a Species must form bonds of intermediate strength with the substrate7. In the case of HDS, the metal-sulfur bond should be strong enough that the substrate bind to the metal, but weak enough that the catalyst allows H28 to desorb from the active Site. If the heat of formation (M-S bond strength) is too strong, then the active site is in essence poisoned by the substrate. If too weak, then the substrate does not bind tightly enough to lead to C-8 bond cleavage. The transition metals in the center of the d-block meet this requirement and are Show the greatest activity in HDS. The activity for the 4d and 5d metals is considerably higher than the 3d metals which are relatively inactive in HDS. Additionally, the 3d TMS do not exhibit a volcano plot behavior, indicating that factors other than periodic position have an influence on catalytic activity as well. Using transition metals in anionic octahedrally coordinated M86” clusters, Harris and Chianelli carried out calculations of the electronic structure of TMS in an attempt to better explain the observed HDS activitiesa. Several factors were identified to be responsible for these observations. Included were the degree of covalency of the TMS, a measure of MS d-p orbital interaction and covalent ll I: O 1*"t Row Ru ' : Ib A 2Ind Row I 3rd ROW OS 10" . I 1‘ b d : :1 U I 1 I mmol"s'1 d °.. ‘1 Mo I Inn] Molecules of DBT Converted Q I 1 2]; ._ . . . ....I Co 10“ , Ni - E W ‘2', I 1 I- -I F . Periodic Position Figure 1.3. Activity of transition metal sulfides in DBT HDS plotted as a function of periodic position of the metal (adapted from Pecoraro and Chianelli)? 12 bond strength of the M-8 bond. The bond strength decreases on going across a period as more electrons are placed in antibonding orbitals. At the same time though, the degree of covalency increases as M-S orbital mixing increases. These factors are in line with the Sabatier principle. A third factor was identified, the symmetry and occupation of the highest occupied molecular orbital (HOMO). This effect is seen in bonding aspects of TMS. The activity is best when the HOMO in these clusters is a metal-based t2g set of antibonding orbitals. The bonding to a thiophene molecule can be through the 8 atom in a o-bonding mode or through the ring in a TT-bonding mode. For 3d metals, the o-bonding interactions are generally weak as the lone pair of electrons on the sulfur that is donated to the metal in bond formation is tied up in aromaticity. For 4d metals, there is a o and Tr component to the bonding, thought to be a requirement for effective bonding of thiophene to transition metals. This is further supported by the fact that thiophene can act as a Tr-acceptor ligand and backbonding from the metal to an antibonding TT' orbital on thiophene is observed, weakening the C-8 bond and thereby increasing activity. These three factors were grouped into a Single activity parameter that correlated well with activity. As mentioned previously, although other TMS exhibit greater activity in HDS catalysis, it is the M082 that is the industry workhorse. When promoted by a second metal such as Co, there is an increase in the electron density, and hence activity, of the primary metal. This is the primary function of the promoter. A model Similar to the previous one designed for unpromoted catalysts took into account a second 3d metal, using MoM’Sg'“ clusters to study the electronic basis 13 for promotiong. AS the metal was varied across the period, the t2g set of orbitals decreases in energy relative to the higher 4d t2g set. Upon reaching cobalt in the series, the 4d t2g set is at an energy between the t2g and eg sets of the 3d metal. Cobalt has seven d electrons, Six of which entirely fill the t2g set while the seventh is now placed in the t2g set of Mo, formally donating to M0 and increasing its electron density. The same effect is observed for nickel. All of the copper orbitals lie below the Mo orbitals and electrons are withdrawn from the molybdenum in this case. Other 3d metals had no effect. While this model does not predict the exact location of the promoter atoms, it is assumed that they are linked to the molybdenum atoms through a bridging sulfur atom. Furthermore, as electrons are transferred from antibonding Co or Ni orbitals to antibonding Mo orbitals, the MOS bond iS weakened as the 00-8 bond is strengthened. A sulfur shared between the two metals will have an intermediate M-S strength and, in accord with Sabatier’s principle, Should exhibit maximum activity. There are several possible locations available to the promoter atom in the M082 catalyst. If the catalyst is supported on, for example, alumina, the cobalt may be embedded in the upper layer of the support, a phase denoted as CozAl203. Cube-like domains of 00983 have also been observed as a component of catalysts. Topsoe and Clausen identified a third phase as well, the so-called Co-Mo-S structure, a mixed Co-Mo sulfide phase in which the three elements are in close proximity to one another”. Through the use of Mossbauer spectroscopy, the type and quantity of cobalt structures were determined. It was found that the Co-Mo«8 structure was the catalytically active phase as trends in activity 14 correlated favorably with the amount of this phase present in the catalyst. It was determined by use of analytical electron microscopy that the most common position of the cobalt is on the edge of an M082 layer. This position for Co iS now generally accepted and has led to the view that M082 itself can be treated as a support for the promoter, allowing for a high content of cobalt and improved activity. In addition to electronic considerations in catalyst performance, a geometric effect is also observed. This has to do with the nature of the crystal structure of the catalyst. For example, RuS2 has an isotropic structure whereas M082 has an anisotropic structure. The ruthenium catalyst’s structure is cubic and there exists a correlation between HDS activity and the surface area of the catalyst. On the other hand, the molybdenum structure is a layered one in which the sheets are stacked only 2 - 10 nm thick, but extend 100 - 300 nm in a direction perpendicular to the layer stack. Accordingly, activity does not correlate to surface area of the M082 catalyst. While the basal planes of this material contribute greatly to the surface area, they do not have an effect on activity in HDS. However, oxygen chemisorption correlates well with HDS activity for both. In the case of the M082, oxygen was Shown to accumulate on edges of the particles, indicating that the edge planes contain the catalytically active sites". While the electronic and geometric effects have an impact on the activity of an HDS catalyst, the crystal structure also plays a role in selectivity. Daage and Chianelli considered the two-site nature of the edges of M082 particles in developing the rim-edge model‘z'“. In HDS of DBT, predominantly two products 15 are formed: biphenyl (BP) and cyclohexylbenzene (CHB). BP is formed from direct desulfurization of DBT while CHB is formed when DBT iS first hydrogenated before undergoing desulfurization. M082 was prepared under various thermal conditions, leading to samples with varying degrees of crystallinity. The layer stack height was determined by measuring the width of the 002 peak in the powder XRD pattern. It was found that as the particles became increasingly crystalline and the stack height increased, there was a severe decrease in the rate constant for the hydrogenation reaction. It was proposed then that hydrogenation occurs only on rim Sites, those which are at the edge of the basal planes. Desulfurization occurs on both rim and edge Sites, the latter located at the edge of internal layers. A convenient feature of this model is that the number of rim sites is‘ dependent only upon the stack height, not the shape of the particles. Thus, selectivity is governed by height within the particle, activity by the diameter of the particle. The active sites of the HDS catalyst are sulfur vacancies at the edges of the M082 particles. In promoted catalysts, it is likely that the active Sites are sulfur vacancies near cobalt that is in the Co-Mo-S phase. A clear mechanism to explain HDS catalysis has yet to be developed. Information used in developing a mechanism has come from solid-state chemistry, organometallic chemistry, and surface science studies. Many model systems use thiophene as the substrate. Prins et al. suggest the following as a possibility“. Adsorption of the thiophene sulfur atom onto the active site occurs first, followed by a four-electron reduction process to yield butadiene and hydrogen sulfide. The electrons for the reduction 16 are supplied either by four one-electron or two two-electron oxidations of molybdenum. In any model, thiophene (or DBT) can bind to the active site in two manners: one-point end-on adsorption through sulfur or Side-on adsorption through a carbon-carbon double bond. The proposed Side-on adsorption mechanism involves the formation of persulfide ions. One support for this mode is that persulfide ions are found in the structures of RuS2, 0382, two of the highest activity TMS catalysts, and in the sulfides of cobalt and nickel, the common promoters of activity. For the larger DBT molecule and its substituted analogs, steric hindrance may become a factor in deciding which mode of adsorption is most likely. Daage and Chianelli address this in their report of the rim-edge model”. The HDS reaction proceeds through a vertical (8 n‘) adsorption mode that can occur on all edge sites. The hydrogenation reaction progresses through a flat adsorption mode that requires a larger surface area of the catalyst. This is unlikely to occur on the edge in general, but at the rim, where part of the molecule will dangle in space. For thiophene, end- on adsorption is common, but for larger molecules such as DBT, a planar adsorption is preferred. While the theoretical studies described above are based on unsupported catalysts, it is a common practice to use a catalyst support in actual HDS reactions. Alumina is the support of choice in this reaction for several reasons. The material can withstand the temperatures and pressures involved in the HDS reaction. The support itself is inert and is not involved in the reaction. This is important in terms of selectivity and efficiency, where acidity can lead to cracking l7 reactions and increased consumption of costly hydrogen. Additionally, the surface area of conventional y- AI203 is high enough that the catalyst can be dispersed effectively, leading to a larger number of catalytically active sites. The use of an alumina support is also cost effective in two ways. First, alumina itself is inexpensive. Second, the use of the support allows for the dilution of the metal, allowing for a less expensive catalyst. One concern with supported catalysts is that the active Sites of the catalyst that are in contact with the support may be inaccessible, thus lowering the activity of the catalyst. If the generally accepted rim-edge model is envisioned and the M082 interacts with the support through the basal planes, then the Sites that are blocked would be rim Sites, on which hydrogenation occurs. Therefore, selectivity would be improved. However, if the M082 particles attach to the support in other manners, then edge sites could become buried, lowering the activity. 1 .4. Mesostructures Since the disclosure of mesostructured Silica molecular sieves by Mobil researchers in 199215”, there has been much work done in the development of these materials. The original work employs cationic surfactant micelles to template silicate oligomers into mesostructures. Removal of the surfactant creates the final products, the M418 family of materials. These materials are characterized by their uniform and tunable pore sizes from ~15 - 100 A, large pore volumes, and high surface areas in excess of 1000 m2/g. The principal 18 member of this family, MGM-41, possesses hexagonal pores with long-range order. Therefore, the mesoporous molecular sieves can be viewed as large-scale analogs to microporous zeolites. Noting the impact that mesostructured Silicas have had in catalysis, Similar improvements may be anticipated with the use of mesostructured aluminas. Particularly, these aluminas could be utilized as catalyst supports in HDS. The larger pore volumes and surface areas of these materials when compared to conventional aluminas would allow for better dispersal of catalyst and increased access to active Sites. Additionally, larger amounts of catalyst may be loaded onto the support without the blockage of pores. With this additional loading of catalyst and increased number of accessible active Sites, the loss of active some Sites due to contact with the support can be tolerated. Following Mobil’s work, Stucky and coworkers identified the important steps in the formation of mesostructured Silicas through electrostatic pathways". First, there is multidentate binding of Silicate oligmers to the surfactant. Second, the oligomers polymerize in the interface region. Third, charge matching between inorganic and organic Species dictates the extent to which the reaction proceeds. These factors were then used to extend MobiI’S original synthetic pathway into four general electrostatic pathways leading to mesoporous materials”. In addition to Silica, mesostructures of antimony, tungsten, iron, lead, zinc, and aluminum oxides could also be prepared from these pathways, although not all materials were thermally stable. l9 Work from the Pinnavaia group established two additional pathways based upon hydrogen bonding interactions between neutral inorganic precursors and neutral primary amine surfactants”. The materials afforded differed from their predecessors. Both materials exhibited uniform pores, but, owing to the relatively weaker hydrogen bonding forces that govern the assembly of the materials in latter work, long range order was not observed. These disordered Silicas, named HMS, were characterized by the wormhole structural motif that arose. Additionally, the wall thickness of materials prepared in this manner was greater than that of materials synthesized from electrostatic pathways, where wall thickness is determined by charge matching. The neutral assembly pathway imposes no such restriction. The textural properties, however, were comparable to those of the previously described materials. Additionally, surfactant molecules could be removed (and recycled) from the aS-synthesized HMS materials through solvent extraction, as opposed to calcination which is required to remove surfactant from electrostatically prepared materials. A study by Tanev20 compared the properties of the two types of materials and reported that HMS materials, due to the thicker framework walls, exhibit improved thermal stability. The smaller particle Size of HMS leads to the formation of complimentary textural porosity, facilitating access to framework-confined mesopores. Another neutral assembly pathway, similar to the previously mentioned one, was also discovered”. The composition of materials formed from these neutral pathways could also be specified by selecting the appropriate inorganic 20 precursor. Dative bonding pathways to tantalum and niobium oxide mesoporous materials were proposed by Antonelli and Yingzz'za. In the ten years that have passed since MobiI’S initial report, the development and proliferation of mesostructured Silica materials has been widespread. To quantify this explosion of research in this field to some extent, consider that a literature search for reports of mesoporous Silica Since 1992 will result in over 1,800 references while a similar search for mesoporous aluminas will yield less than 8% of that amount. The 8% is exaggerated to some degree, as many of the references describe aluminas prepared by conventional, non-templating, methods or alumina in mixed metal oxide materials. The results of the search are especially discouraging, noting the extensive use of aluminas in catalysis. It is also a testament to the elusiveness of the ability to synthesize these materials. While the increased utility of mesoporous Silicas in catalytic applications has already been realized“, a Similar increase in the use of mesoporous aluminas is anticipated in the near future. 1.5. Mesostructured Aluminas Many aluminas are porous by nature. The transition aluminas formed through thermal dehydration of aluminum hydroxides are examples. However, the pore size in these materials is not uniform; a broad distribution of sizes is observed. This leads to necking that can limit access to the pores. Mesoporous aluminas assembled through templating pathways offer narrower pore Size distributions 21 and improved textural properties. Bagshaw and Pinnavaia first reported a mesoporous alumina synthesized in this manner”. A series of aluminas, titled MSU-X, were synthesized from hydrolysis of aluminum sec-butoxide in the presence of nonionic surfactants. After calcination, these materials exhibited surface areas up to 535 m2/g and pore volumes up to 0.68 cc/g with uniform pore sizes. The wormhole frameworks were thermally stable to temperatures of 500 °C. The 5-coordinate aluminum centers detected by 27Al NMR spectrum, a property of amorphous aluminas, suggests that Lewis acid sites may be present in these materials. Other reports of mesoporous alumina quickly surfaced following Bagshaw’s initial communication. Yada and coworkers synthesized alumina mesophases from aluminum nitrate and sodium dodecyl sulfate26'27. A lamellar phase can be converted into a disordered hexagonal phase during hydrolysis of the aluminum salt by urea. Although the hexagonal phase is not evident by XRD after calcination at 600 °C, a low angle peak is observed, indicating that a disordered mesoporous phase is present. A report from Davis’ group described the formation of mesoporous aluminas prepared from the hydrolysis of aluminum alkoxides in the presence of carboxylic acids as structure directors”. Although the materials were calcined at relatively low temperatures for Short durations, surface areas up to 710 mz/g were reported. The pore size exhibited by these materials was near 20 A regardless of surfactant Chain length. 22 Reporting a generalized synthesis strategy leading to large pore mesoporous metal oxides, Stucky and coworkers provided an example of a mesoporous alumina”. Nonionic surfactants were used to template the oxide from an aluminum Chloride precursor. The homogenous sol formed was allowed to gel in an open Petri dish at 40 — 60 °C. The resulting alumina, once calcined, showed an average pore Size of 140 A with a pore volume of approximately 0.35 cc/g. The framework walls of the alumina were amorphous. A contribution from Gabelica and coworkers indicated a novel approach to mesoporous aluminas”. Mixing of anionic and cationic surfactants was employed to achieve better charge matching between the inorganic and organic phases. Aluminum salts and Keggin ions were precipitated in the presence of these surfactant mixtures to yield aluminas with surface areas up to 810 m2/g, an unprecedented value. Remarkably, there has been no further work reported stemming from this group’s initial report. The pore Size and pore volume reported for the high surface area alumina were 27 A and 0.61 cc/g, respectively. The material was comprised of a regular arrangement of pores with amorphous walls, as judged by XRD and the 27AI NMR. 1 .6. Stabilization of Aluminas All of the previously described aluminas are comprised of amorphous frameworks. While these materials were generally thermally stable up to 500 °C, they may not remain stable under certain catalytic conditions where higher 23 temperatures and pressures and hydrothermal conditions are encountered. Over time and with extended use under these conditions, alumina particles can grow in Size and Sinter, suffering structural and textural modification. Transition alumina phases are only metastable and will eventually be converted to d-Al203 under extreme thermal conditions. A concomitant loss of textural properties is observed with this structural transformation. To combat these effects, aluminas have been doped with small percentages of lanthanides, typically 1 — 5% lanthanum and cerium. This has been carried out primarily by two methods. Coprecipitation of the two oxide precursors or incipient wetness impregnation of lanthanide containing salts can afford the stabilized materials. There are also two explanations for the resulting stability of the doped materials. Shaper31 and Beguin32 concluded that a lanthanum aluminate compound covered the surface of the alumina. Oudet33 proposed that a LaAI03 perovskite type structure behaves similarly. In each case, the stabilization is due to strong interactions between the lanthanide structure and the alumina, inhibiting the conversion to o- AI203. Alternatively, it was proposed by Burtin34 that stability was due to the occupation of lanthanide ions in cationic vacancies of the spinal type lattice of alumina. Braun et al. also concluded Similarly in their doping work”. In this manner, the lanthanide ions fortify the alumina structure and limit the amount of diffusion of aluminum ions on conversion to corundum. Other metal ions have been used in stabilization, however, lanthanum and cerium have consistently shown the best results. Cerium exhibits two common oxidation states, Ce3+ and Ce“. Oxidation to the higher state can lead to the formation of CeO2 and its 24 leaching from the alumina. Lanthanum exhibits only the La3+ oxidation state and is therefore a better stabilizer. The stabilization of MSU-X aluminas has been achieved through lanthanide doping”. Indeed, textural properties were improved in doped aluminas as compared to the pure aluminas. In addition to the stabilization of amorphous aluminas, it is of interest to synthesize mesoporous crystalline aluminas in an effort to improve catalytic properties. Owing to the crystallinity, these materials exhibit inherent stability when compared to their amorphous analogs. The first report of such an alumina was recently made by Pérez-Pariente and coworkers”. In a synthesis Similar to that of MSU-X alumina, aluminum alkoxides were hydrolyzed with nonionic surfactants and dipropylamine to yield a thermally stable, crystalline mesoporous alumina. The synthesis was altered to include thermal aging of the freshly made gel. Upon calcination, mesoporosity was retained within the crystalline product. The XRD pattern of the calcined product Showed peaks assignable to a transition alumina phase, confirming a crystalline phase. Additionally, unlike many of the amorphous aluminas, there was no peak in the 27Al NMR corresponding to 5- coordinate aluminum centers. Surface areas in excess of 500 mzlg and pore volumes of 0.91 cc/g were attained. The degree of crystallinity and the textural properties were compromised after calcination at 700 °C. 25 1.7. Objectives and Rationale AS can be inferred from the reports summarized above, the synthesis and application of mesoporous aluminas are areas that are ripe for exploration and development. One objective of this study is to synthesize novel mesoporous aluminas. Alumina is perhaps the world’s most widely used catalyst support. As such, new materials with improved structural and textural properties would be a welcome addition to the alumina family. Noting the unstable nature of the amorphous aluminas, the focus here is on crystalline aluminas. To date, only one example of a crystalline mesostructured alumina has been reported. Although these materials promise to be more stable than their amorphous relatives, corundum, o-Al203, is the only phase of alumina that can be considered truly stable. Therefore, doping of lanthanides and transitional metals into the alumina frameworks in an effort to improve thermal and hydrothermal stability, with retention of textural properties, will also be explored. The doping of ions into the frameworks of these aluminas can also stabilize a particular phase. This is important in many catalytic applications where the elevated temperatures required for catalysis initiate phase transitions of the alumina, accompanied by degradation of textural properties. Additionally, the incorporation of transition metal ions may introduce catalytically active centers into the alumina. Whatever strategy is employed for stabilization, the goal is to produce high performance mesoporous aluminas. 26 Finally, another goal is to demonstrate the value and utility of these newly designed aluminas. Especially of interest is the ability of these materials to function as catalyst supports in hydrodesulfurization reactions. Crystalline aluminas offer the necessary stability under the conditions used in HDS reactors. Aluminas with improved properties give hope that more effective and efficient routes to sulfur removal will be found, enabling the petroleum and automotive industries to meet the demands of government and society. This iS an increasingly crucial need as high quality oil feedstocks are consumed, leaving heavier and dirtier stocks as the primary fuel source. 27 1.8. References (1) (2) (3) (5) (6) (7) (8) (9) (1 0) (11) (12) (13) (14) (15) (16) Misra, C. American Chemical Society Monograph, Vol. 184: Industrial Alumina Chemicals, 1986. International Aluminum Institute, 2002. Wefers, K.; Misra, C. "Oxides and Hydroxides of Aluminum," Alcoa Laboratories, 1987. In New York Times: New York, NY, 2000. Pecoraro, T. A.; Chianelli, R. R. J. Catal. 1981, 67, 430-445. Chianelli, R. R. NATO AS! Sen, Ser. C1983, 105, 361-378. Sabatier, P. Ber. Deutsch Chem. 693. 191 1 , 44, 2001 . Harris, 8.; Chianelli, R. R. J. Catal. 1984, 86, 400-412. Harris, 8.; Chianelli, R. R. J. Catal. 1986, 98, 17-31. Topsoe, H.; Clausen, B. S. Catal. Rev. - Sci. Eng. 1984, 26, 395-420. Tauster, S. J.; Pecoraro, T. A.; Chianelli, R. R. J. Catal. 1980, 63, 515- 519. Daage, M.; Chianelli, R. R. J. Catal. 1994, 149, 414-427. Daage, M.; Chianelli, R. R.; Ruppert, A. F. Stud. Surf. Sci. Catal. 1993, 75, 571-584. Prins, R.; De Beer, V. H. J.; Somorjai, G. A. Catal. Rev. - Sci. Eng. 1989, 31, 1-41. Beck, J. 8.; 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.; et al. J. Am. Chem. Soc. 1992, 114, 10834-10843. Kresge, C. T.; Leonowicz, M. E.; Roth, W. J.; Vartuli, J. 0.; Beck, J. 8. Nature (London) 1992, 359, 710-712. 28 (17) (18) (19) (20) (21) (22) (23) (24) (25) (26) (27) (28) (29) (30) (31) (32) Monnier, A.; Schuth, F.; Huo, O.; Kumar, 0.; Margolese, D.; Maxwell, R. 8.; Stucky, G. D.; Krishnamurty, M.; Petroff, P.; et al. Science (Washington, D. C., 1883-) 1993, 261, 1299-1303. Huo, 0.; Margolese, D. l.; Ciesla, U.; Feng, P.; Gier, T. E.; Sieger, P.; Leon, R.; Petroff, P. M.; Schueth, F.; Stucky, G. D. Nature (London) 1994, 368, 317-321 . Tanev, P. T.; Pinnavaia, T. J. Science (Washington, D. C.) 1995, 267, 865-867. Tanev, P. T.; Pinnavaia, T. J. Chem. Mater. 1996, 8, 2068-2079. Bagshaw, 8. A.; Prouzet, E.; Pinnavaia, T. J. Science (Washington, D. C.) 1995, 269, 1242-1244. Antonelli, D. M.; Ying, J. Y. Angew. Chem, Int. Ed. Engl. 1996, 35, 426- 430. Antonelli, D. M.; Ying, J. Y. Chem. Mater. 1996, 8, 874-881. Corma, A. Chem. Rev. (Washington, D. C.) 1997, 97, 2373-2419. Bagshaw, S. A.; Pinnavaia, T. J. Angew. Chem, Int. Ed. Engl. 1996, 35, 1 102-1 105. Yada, M.; Machida, M.; Kijima, T. Chem. Commun. (Cambridge) 1996, 769-770. Yada, M.; Hiyoshi, H.; Ohe, K.; Machida, M.; Kijima, T. Inorg. Chem. 1997, 36, 5565-5569. Vaudry, F.; Khodabandeh, 8.; Davis, M. E. Chem. Mater. 1996, 8, 1451- 1464. Yang, P.; Zhao, D.; Margolese, D. |.; Chmelka, B. F.; Stucky, G. D. Chem. Mater. 1999, 11 , 2813-2826. Valange, 8.; Guth, J. L.; Kolenda, F.; Lacombe, 8.; Gabelica, Z. Microporous Mesoporous Mater. 2000, 35-36, 597-607. Schaper, H.; Doesburg, E. B. M.; Van Reijen, L. L. Appl. Catal. 1983, 7, 211-220. Beguin, B.; Garbowski, E.; Primet, M. Appl. Catal. 1991, 75, 119-132. 29 (33) (34) (35) (35) (37) Oudet, F.; Courtine, P.; Vejux, A. J. Catal. 1988, 114, 112-120. Burtin, P.; Brunelle, J. P.; Pijolat, M.; Soustelle, M. Appl. Catal. 1987, 34, 239-254. Braun, 8.; Appel, L. G.; Zinner, L. B.; Schmal, M. Br. Ceram. Trans. 1999, 98, 77-80. Zhang, W.; Pinnavaia, T. J. Chem. Commun. (Cambridge) 1998, 1185- 1186. Gonzalez-Pena, V.; Diaz, l.; Marquez-Alvarez, C.; Sastre, E.; Perez- Pariente, J. Microporous Mesoporous Mater. 2001, 44-45, 203-210. 30 Chapter 2 Synthesis and Characterization of Mesostructured Boehmite 2.1. Introduction Much effort has been invested in the development of alumina phases for industrial use. Heavy interest lies in the production of transition aluminas with favorable surface and textural properties. The primary target phase is y-Al203. This phase is achieved by thermal dehydration of boehmite‘. Noting the importance of y-AI203 as a catalyst and catalyst support, one can appreciate that boehmite is also a commercially important chemical. However, comparatively few reports on the synthesis of boehmite can be found in the literature. Aside from its most important use as the precursor to y-Al203, boehmite has other uses as well. Boehmite has served as a component In the synthesis of aluminum containing anionic clays or layered double hydroxides2 and carboxylate alumoxanes3. It has been used infrequently as a catalyst support and as a catalyst itself. The lack of applications of boehmite as a catalyst stems from the widespread use of pseudoboehmite in this capacity. The differences in these two materials are discussed below. Boehmite is an aluminum oxide hydroxide (AlOOH). The structure of boehmite is comprised of close packed oxide and hydroxide ions in which double chains of AlOOH molecules extend along the a-axis. These units are stacked in the direction of the b-axis where hydroxyls of one layer are located over the 31 depressions between adjacent hydroxyls in the other layer (Figure 2.1)“. Hydrogen bonding holds the double layers together. Pseudoboehmite, or gelatinous boehmite, can be classified as alumina structurally Similar to that of boehmite, but having a greater water content and smaller particle Size than boehmite. Aluminum alkoxides are generated as byproducts in the formation of linear alcohols by the Ziegler process. Hydrolysis of these alkoxides results in the production of pseudoboehmite, available commercially under the name of Catapal‘. This material has a high surface area and is easily converted to a useable sol form by peptizing with dilute (typically nitric) acids. In this respect, it is discernable from boehmite. When deciding which material to use in applications where a soluble aluminum source is desired, this property makes pseudoboehmite the material of choice. Not surprisingly, the terms boehmite and pseudoboehmite are often incorrectly used in the literature. In addition, both have also been referred to as aluminum hydroxides, causing even more confusion. There has been some debate as to the actual differences in the two materials. Boehmite contains 15% H2O by mass where pseudoboehmite has been found to contain up to 30% H2O. The location of the excess water in the structure was originally thought to be contained in the region between the double layers. However, d-Spacings found for the 020 peak in the XRD patterns of the materials were not large enough to accommodate such an excess volume of water“. This leads to the currently accepted idea that pseudoboehmite is structurally identical to boehmite, differing only in its smaller crystallite size and higher water content. In this model 32 OH OH OH OIH CiH (1H I I —Al— -—A1— —Al— Ti I‘i‘i—i“? I _i_.I_I._I_ —:I—O—:l— O—TI—O—Al—O _.T ”on, “on l‘l_,0H,\ III/OH‘R “(1H \f'oIH (1H OIH 'OlH' (1H — —AI— —Al— —Al— —?1— (r—Tl— (I—Tl— ‘I i I I I I _ _ _ — A1—(|)— Tl—O— ITI—O— Tl—O— Ald—O [I] o OIH OH OH OH OH OH 12A Figure 2.1. Cross-sectional (upper) and three-dimensional (lower) depiction of the boehmite double-layer. The thickness of the double layer is ~ 12 A. 33 of pseudoboehmite, the increased number of terminal hydroxyl groups accounts for the excess water. The smaller the crystals, the greater the number of hydroxyls present. Traditionally, boehmite is synthesized by the neutralization of aluminum salts under hydrothermal conditions“. Unlike the y-Al203 and the other transition aluminas, boehmite iS relatively non-porous. Mesoporous aluminas prepared in the presence of surfactants have been reported in the literature, however, all but a few describe materials that are amorphouss’”. Similarly, the use of surfactants to direct the formation of mesostructured boehmite is reported here. The novelty of this approach lies in the choice of the inorganic precursor and the surfactants employed in the synthesis and the method of surfactant removal. The use of aluminum sec-butoxide allows for formation of a boehmitic alumina phase during synthesis. Primary and tallow amine surfactants decompose at a temperature that allows for their removal via calcination without the concomitant conversion of boehmite to y-Al203. 2.2. Experimental 2.2.1. Synthesis The synthesis of mesostructured boehmite was executed as follows: The surfactant, a primary or tallow amine (see footnote to Table 2.1), was dissolved in a solution of ethanol and water (75/25 v/v) at ambient temperature. To this solution, aluminum sec-butoxide was added under vigorous stirring by the use of 34 a magnetic stirrer. The mixture was either allowed to stir at ambient temperature or was transferred to a temperature controlled shaker bath for 20 hours in a sealed vessel. Precipitates formed upon hydrolysis of the alkoxide were recovered by vacuum filtration, washed with water and ethanol, and allowed to dry at ambient temperature. The final surfactant-free products were obtained by calcining the aS-synthesized products in air at 325 °C for 10 hours. Alternatively, surfactant could be removed by ethanol extraction at 80 °C. While the amounts of the starting materials were varied, a typical synthesis had the following molar ratios: 5.0 AI(8-BUO)3 : 1.0 Surfactant : 28.7 Ethanol : 31.0 Water 2.2.2. Characterization Powder diffraction patterns were collected on a Rigaku Roteflex Diffractometer using Cu K, radiation (A = 0.154 nm). Intensity was calculated in counts per second every 0.02 degrees 26. Sweep rates were varied from 2 — 8 degrees 26/min. Nitrogen adsorption — desorption isotherms were collected on a Micromeritics TriStar 3000 sorptometer at 77K. Samples were degassed at least 12 hours at 150 °C under a vacuum of 1043 torr prior to analysis. Surface areas and pore sizes were determined by the BET15 and BJHl6 methods, respectively. TEM images were collected on a JEOL 100CX microscope with a CeBs filament at an accelerating voltage of 120 kV. Samples were prepared by sonicating a suspension of the powder in ethanol for 20 min. One to two drops of 35 the suspension were then placed on a carbon coated holey film that was supported on a 3mm, 300 mesh copper grid. 27Al NMR spectra were collected on a Varian 400 VRX solid — state NMR spectrometer. Samples were placed in a zirconia rotor and Spun at 4 kHz with a pulse width of 0.5 us and a pulse delay of 0.5 3. Chemical Shifts were referenced to AI(H2O)63+ which is assigned a value of 0 ppm. 2.3. Results and Discussion 2.3.1 . Formation of MSU-SIB Addition of the aluminum sec-butoxide to the surfactant solution results in rapid hydrolysis of the alkoxide, instantly forming a precipitate of a surfactant — boehmite composition, denoted MSU-SIB. The conversion of the alumina species is as follows: Al(O-‘°’Bu)3 + 2H2O —-> AlOOH + 3 s-BuOH It has been reported that the formation of boehmite requires hydrothermal conditions‘. Although these reactions are performed at ambient temperature, two reasons may help to explain the presence of boehmite as a product of the reactions. First, the hydrolysis of the alkoxide is exothermic and a considerable amount of heat is released. The volume of the solution in the reaction mixture iS not great enough to disperse this heat and initially the temperature in the reaction vessel is much greater than ambient temperature. Second, and more importantly, by providing multiple Sites for the formation of hydrogen bonds between amine 36 head groups and boehmite hydroxyls, the use of the surfactant aids in the growth of particles. Boehmite particles will terminate with an exposed layer of hydroxide groups at the surface. These groups readily form h-bonds with the surfactant head groups (Figure 2.2). In contrast, hydrolysis of aluminum alkoxides in the absence of surfactant leads to the formation of pseudoboehmite. Indeed, this is one method of commercial production of pseudoboehmite. Upon calcination, the surfactant is removed and the as-synthesized MSU-SIB iS converted to a mesostructured boehmite, denoted MSU-B. Noting the transformation sequence upon dehydration of aluminas, the conversion of boehmite to y-Al203 occurs at temperatures of 400 °C‘. Consequently, if calcination is performed above 400 °C, the structural integrity of the boehmite phase is compromised. The use of primary and tallow amines is key in this regard because they will degrade at temperatures below 400 °C and can therefore be removed at lower calcination temperatures. Also upon calcination, particles may Sinter leading to growth of larger crystals at the expense of smaller ones. This helps to ensure that the oxide framework is comprised of boehmite and not pseudoboehmite. . Calcination could also be performed at higher temperatures. In fact, MSU-B materials were stable to calcination at 350 °C for short durations. At longer periods or at 400 °C, the onset of conversion from boehmite to y—Al203 was observed as judged by the appearance of peaks assignable to y-Al203 in the wide angle XRD patterns. Therefore, milder conditions were preferred and 37 F 3 L J [ Irfir r l IIIITETIIII «at m //‘\\ /\"". . . [\\,- P 3 4 1" ‘ ‘ VV "rat-5?.» 9.5:, r W °” 5“” Figure 2.2. Schematic of interactions leading to formation of MSU-SIB. Each Slab represents one boehmite double layer. Lines on the boehmite layers represent hydroxyls forming hydrogen bonds to the surfactant micelles. Areas of overlapping boehmite layers indicate cross-linking through condensation of adjacent hydroxyls. 38 calcination was performed at the lowest temperature possible to preserve the boehmite phase while still removing the surfactant. Calcination is a convenient method of removing the surfactant, although the surfactant is destroyed in the process. Surfactant removal and recovery can be achieved through ethanol extraction if desired. However, the small amounts of primary and tallow amines used and their costs, compared to other more expensive surfactants such as polyalkylene oxides, do not warrant their recovery. Calcination also ensures that hydroxyls on the boehmite surface are condensed, eliminating water and fortifying the structure of MSU-B. 2.3.2. General Properties of MSU-B Aluminas MSU-B aluminas were characterized by powder XRD, nitrogen adsorption, TEM, and 27Al NMR. Figures 2.3 and 2.4 Show the low angle and wide angle powder XRD patterns, respectively of aS-synthesized MSU-SIB and MSU-B materials, prepared under conditions as described in Table 2.1, line 3. From peaks in the wide angle patterns it is evident that the alumina is present in the boehmite phase. This occurs invariably in all samples prepared in the method described above. However, while a low angle peak is observed for this sample, this is not always the case. As discussed in greater detail below, MSU-B materials derived from alkoxide are not formed through a true templating mechanism. Therefore, low angle peaks that are typically taken as evidence of mesophase formation due to pore-to-pore correlation distances in the synthesis of mesoporous materials cannot be relied upon here. 39 6.2 nm 4,-- MSU-B Intensity 8 10 Degrees (2(9) Figure 2.3. Representative low angle XRD patterns for the as-synthesized surfactant-boehmite composite, denoted MSU-SIB, prepared from aluminum sec- butoxide and dodecylamine in an alcoholic/aqueous medium as described by line 3 in Table 2.1 and the surfactant-free analog, MSU-B, obtained by calcination of the MSU-SIB composite at 325 °C for 10 hr. 40 020 120 MSU-B . 140 031 051 200 231 Intensity Degrees (20) Figure 2.4. Representative wide angle XRD patterns for MSU-SIB and MSU-B aluminas Shown in Figure 2.3. Crystallographic planes contributing to the peaks are noted. 41 In the particle assembly of these materials, the morphology of the oxide formed does not necessarily reflect the shape of the surfactant micelle. Rather, it is believed that a low angle peak, when observed, iS due to the thickness of fundamental particles. While the observance of the peak itself is not always reproduced from one reaction product to another under identical reaction conditions, the d-spacings extracted from the peak position are not. The height of the boehmite double layer is approximately 12.2 A. The recurring values of the observed d-Spacings of calcined samples are approximately integer multiples of 12 A. It appears that 5 — 7 double layers per particle is most common as suggested by the position of the low angle reflection that occurs most frequently at 26 values corresponding to d-Spacings of 60, 72, and 84 A. There is more variance in the d-Spacings of the MSU-SIB samples, but there may be some water or alcohol intercalated between the layers that causes the repeat distance to differ from that of the theoretical value of 12 A. The nitrogen adsorption — desorption isotherms of MSU-B materials are Shown in Figure 2.5. These isotherms correspond to aluminas prepared with varying surfactant concentrations under conditions described in lines 3, 8, and 19 of Table 2.1. The isotherms for MSU-B most Closely resemble the Type IV ”"8 and associated with mesostructured materials. The isotherm defined by Sing lack of a well-defined step in the adsorption branch indicates that the pore Size distributions are somewhat broad. This is in line with the particle assembly pathway leading to formation of the material. Although not well defined, the isotherms indicate that MSU-B materials have more uniform porosity as 42 1000 2 a 2 8004 1’: A ‘ 4% DDA O'l . B . oungflfit'oflflisflfizo 3 Pore Diameter (nm) '0 6004 d) n 4 L O . g 9% DDA < - q, 4004 E _ 2 o _. > - 0% DDA 200* (Blank) o"'l"‘l"'l"'lr" 0 0.2 0.4 06 08 1 PIP 0 Figure 2.5. Nitrogen adsorption — desorption isotherms and BJH adsorption pore size distributions (inset) for MSU-B aluminas differing in surfactant concentration. The percentage given corresponds to the mass percent of dodecylamine in the surfactant solution. The three isotherms, offset by 200 cc/g for clarity, correspond from top to bottom to samples described in lines 3, 8, and 19 in Table 2.1. PSD curves are displayed in the same order as the isotherms. 43 Table 2.1. Synthetic parameters and textural properties of MSU-B aluminas. b Pore Pore Al/SurI/EtOH/H2O Temp 833 , Surfactant“ . 2 Sue" Volume Molar RatIo (°C) (m lg) (nm) (cc/0) 1 DDA 5 2 1 220.1 : 196 20 380 5 0.59 2 DDA 5 : 1 :42.6 2 138 20 428 3 0.72 3 “DDA” 5 :“’i”:‘68.0”:“73.4“””20“””4‘55”" ’7 4.5 '7 ‘ 0195“” '4 "”‘bD‘A 5?'1‘"‘“:”58I0':75.4"“ 45 “ ”"4977W3.5‘ 084 5 DDA 5 2 1 268.0 2 73.4 65 460 5 0.82 6 DDA 5 2 1 2 68.0 273.4 95 451 4 0.69 " 7 DA 5: 1 : 20.0 : 21.5 20 325 5 0.47 8 DDA 5 : 1 :28.7 2 31.0 20 378 3 0.59 9 HDA 5 : 1 237.4 2 40.4 20 391 3.5 0.59 10 Tal A 5 : 1 :46.2 249.9 20 410 2.5 0.60 11 Tal TriA 5 : 1 264.0 2 69.0 20 429 2 0.60 12 Tal TetA 5 : 1 272.8 2 78.7 20 436 2.5 0.61 “1'37 OA 2.2: 1 503221.577 ~'20 #4527 5" “Ciel—T” 14 DDA 3.1 2 1 :20.02 21.6 20 418 5 0.63 15 HDA 4.0 : 1 220.0: 21.6 20 434 3 0.64 16 Tal A 4 0: 1 220.0: 21.6 20 410 3 0.60 17 Tal TriA 5.9 : 1 220.0: 21.6 20 462 3 0.63 18 Tal TetA 6.9 : 1 :20.0: 21.6 20 459 3 0.61 "1‘9“"Nofiemw 570728.77513—7720 322 "4" l ”0.71””7 a DDA = dodecylamine, OA = octylamine, HDA = hexadecylamine, Tal A = tallow amine, Tal TriA = tallow triamine, Tal TetA = tallow tetraamine The tallow amine structure is generalized by the following formula: R-HN-(CH2CH2CH2NH)x-H Where R = a mixture of alkane groups ranging from C12-C13 and x = 0, 1, 2, or 3 for tallow amine, diamine, triamine, and tetraamine, respectively. b Ambient temperature taken as 20 °C ° BJH pore diameter as measured from the adsorption branch compared to the sample prepared without surfactant. The shape of the isotherm of the blank sample does not proceed through an inflection point at any partial pressure above 0.2. In this manner, its behavior most closely resembles a Type II isotherm, exhibited by non-porous materials. A summary of textural properties of MSU-B aluminas iS provided in Table 2.1. Included also are the textural properties for one sample prepared in absence of surfactant. From the data, it can be seen that surface areas of MSU-B aluminas in the approximate range of 380 — 460 m2/g, pore Sizes 2.0 — 5.0 nm, and pore volumes 0.60 - 0.95 CC/g. Entries in lines 1-3 of the table correspond to samples in which the surfactant solution becomes higher in ethanol content. Based upon XRD evidence, more surfactant is incorporated into the mesostructure at higher ethanol content. Free surfactant peaks appear around 20 degrees 26 in the wide angle powder XRD patterns (not Shown) of precipitates borne out of low ethanol content solvents. Judging by the textural properties of these samples, surface area, pore size, and pore volume all increase at higher ethanol content. The more surfactant incorporated into the mesostructure, the greater the void left upon its removal. The isotherms are provided in Figure 2.6. The third through sixth entries in Table 2.1 are for samples of identical stoichiometry aged at different temperatures. Figure 2.7 provides the isotherms for these aluminas. While the surface areas seem random, both pore size and volume steadily decrease with an increase in aging temperature. At higher temperatures, IeSS surfactant iS incorporated into the mesostructure. 45 (NM!) 1000 1 K 75% EtOH V V I 1' ‘T‘T YYYYYYYYYYY 1 00 O O I o 5 10 15 Pore Diameter (nm) Volume Adsorbed (cc/g) a: O O _ 50% EtOH 400— ‘ 25% EtOH 2006 O I I I I I I I I r I I I I r I I . . . 0 0.2 0.4 0.6 0.8 1 PIP 0 Figure 2.6. Nitrogen adsorption - desorption isotherms and BJH adsorption pore Size distributions (inset) for MSU-B aluminas prepared from aluminum sec- butoxide and dodecylamine at ambient temperature in alcoholic/aqueous solvents of varying ethanol content. The three isotherms, offset by 200 cc/g for clarity, correspond to MSU-B aluminas described in lines 1-3 of Table 2.1. PSD curves are displayed in the same order as the isotherms. d N O O 1 l 1 1 dV/dD IIIIIIIIIIIIIIII Pore Diameter (nm) Volume Adsorbed (cc/g) Figure 2.7. Nitrogen adsorption — desorption isotherms and BJH adsorption pore size distributions (inset) for MSU-B aluminas prepared from aluminum sec- butoxide and dodecylamine at different temperatures. The four isotherms, offset by 200 cc/g for clarity, correspond to MSU-B aluminas described in lines 36 of Table 2.1. PSD curves are displayed in the same order as the isotherms. 47 Filling of pores also occurs at lower partial pressures for products of higher temperature reactions, signifying larger particle growth. Entries in lines 7-18 of Table 2.1 are for aluminas prepared from different amine surfactants under otherwise identical processing conditions. Nitrogen adsorption - desorption isotherms for these MSU-B samples are provided in Figure 2.8. The aluminum to surfactant molar ratio was held constant at 5:1 and the surfactant solution in each case was 9% by mass surfactant (lines 7-12). Then this series of experiments was repeated, maintaining a constant aluminum to surfactant mass ratio (lines 13-18). It is interesting that the pore size and pore volume are approximately constant regardless of the size, mass, or molar mass of the surfactant. The surface areas do vary, but not predictably. There is a steady increase with increasing surfactant size in the constant molar ratio products, but they do not correlate well with surfactant size in the latter case. in traditional templated syntheses, pore sizes and volumes increase with increasing size of the surfactant molecule. That the pore size remains nearly constant under a range of synthetic conditions gives support to a particle assembly mechanism. The porosity of MSU-B boehmites arises from the voids between individual boehmite particles. As noted above, the pores of the oxide framework do not mimic the size or shape of the surfactant. So, although regular micelles may be formed, the boehmite particles may be ordered randomly about the micelles, leading to irregularly shaped pores. Again, noting the difference in these materials and silica materials prepared through supramolecular templating, there may be instances when the observed d-spacing is actually smaller than the 48 1.00:, k ‘3 “'0 j k TaiTetA A i ................ m _ ll 3 1'0 1T5 2'0 3 1200 _ Pore Diameter (nm) Tal TriA 3 q ,_. 'o i .2 _ TalA h .. O 1 '8 J HDA < 800—: d) d/, g _: DDA '5 */ f > 2 // 400: / 0A OIIIIII'FT“l‘r"'Tr 0 02 0.4 06 08 1 PIP 0 Figure 2.8. Nitrogen adsorption - desorption isotherms and BJH adsorption pore size distributions (inset) for MSU-B aluminas prepared from the hydrolysis of aluminum sec-butoxide in the presence of different amine surfactants as porogens. The six isotherms, offset by 200 cc/g for clarity, correspond to MSU-B aluminas described in lines 7-12 of Table 2.1. PSD curves are displayed in the same order as the isotherms. average pore size. in materials prepared from templating synthesis, pore sizes are smaller than the d-spacings because the d-spacing is indicative of the repeat distance of both pore and framework wall. However, in MSU-B, the XRD peak is most likely due to the particle thickness and the pore size is due to void spaces, between particles. Therefore, the fact that the small angle XRD d-spacings do not correlate with the pore size is understandable. The presence of a scaffold morphology was identified through TEM. Bright field images of representative MSU-B aluminas are shown in Figure 2.9. The open framework resulting from the random orientation of fundamental boehmite particles can be seen. it is estimated that average particle thickness in these samples is 3 — 7 nm. Electron diffraction patterns are also provided for these materials. The presence of diffuse rings confirms the polycrystalline nature of the materials. Many small crystals in all orientations give rise to the diffuse rings. This is in contrast to a single crystal that would show distinct points in the pattern. The patterns contain four rings containing the strongest reflections of boehmite. 27Al NMR spectra were obtained to verify the coordination environment of the aluminum in MSU-B) (Figure 2.10). 2"Al NMR chemical shifts are referenced to [Al(H20)6]3", which is assigned a value of 0 ppm. Octahedral (6-coordinate), tetrahedral (4-coordinate), and 5-coordinate aluminum will have shifts of less than 10 ppm, 65 — 75 ppm, and approximately 35 ppm, respectively. In boehmite, all aluminum has octahedral coordination. The shift observed for MSU-B at 7 ppm is in agreement with a boehmite structure. The small peak coincident with 50 Figure 2.9. Bright field TEM images (A and C) and electron diffraction patterns (B and D) of MSU-B materials prepared from aluminum sec-butoxide and dodecylamine as described in Table 2.1, (A) line 8 and (B) line 3. The open framework results from the random orientations of fundamental particles. Electron diffraction patterns, (B) and (D), corresponding to bright field images (A) and (C), respectively show diffuse rings, signifying the polycrystalline nature of the material. The reflections contributing to the pattern are labeled in (D). 51 52 7.00 [lTTIllTTTIrrFTrUIIllllllrllllliiiTlTlrl] 200 150 100 50 0 -50 -100 -150 -200 ppm Figure 2.10. 27Al NMR spectrum of MSU-B prepared from aluminum sec- butoxide and dodecylamine prepared as described in Table 2.1, line 3. The resonance at 7 ppm is consistent with the structure of boehmite, in which all aluminum ions are in octahedral environments. 53 the spinning sideband may signify a small amount of Al in a tetrahedral environment, possibly due to the onset of the boehmite to gamma alumina phase transition. 2.3.3. Peptization Boehmite and pseudoboehmite can be differentiated from each other based upon their ability to be dispersed in dilute acidic solutions. The small particle size and high water content of pseudoboehmite will cause this form of boehmite to dissolve and form a sol in such solutions. This form of boehmite is said to be peptizable. Boehmite, on the other hand, does not behave in this manner. MSU- B materials were tested for peptizability to determine the identity of the oxidic framework in these materials. Small amounts of as-synthesized MSU-SIB, calcined MSU-B, and ethanol extracted MSU-B materials were added to 0.05M HNOa. in no instance did the materials peptize, as judged by the settling of the materials. It is possible that in the case of the as-synthesized samples, the presence of surfactant restricts access to the boehmite framework. If this is the case, then it cannot be concluded that the framework in the MSU-SIB materials is boehmite. However, after surfactant removal, neither calcined nor ethanol extracted MSU-B materials peptize, indicating that these surfactant free materials truly are boehmites. The sample prepared without surfactant was also tested for peptizability. it was determined that the as-synthesized material is peptizable. That small particle pseudoboehmite is formed in this synthesis is not an unreasonable result when 54 the hypothesis that the presence of surfactant allows for particle growth is employed. After calcination, however, this product also does not peptize. Therefore it is believed that calcination leads to sintering and growth of the particles that governs the formation of boehmite as opposed to pseudoboehmite. Calcination may play a dual role in the synthesis of MSU-B. Not only does this lead to surfactant removal, but it also helps to ensure that boehmite is the phase that is formed. 2.4. Mechanistic Aspects of MSU-B Alumina Synthesis Bagshaw and Pinnavaia first reported the synthesis of mesoporous alumina, named MSU-X aluminas. It was prepared using aluminum alkoxide and PEG surfactants, but under different synthetic parameters and calcination procedures than those used in the present work. In the MSU-X synthesis, the alkoxide to surfactant ratio is greater than the ratio used here and the two reagents are mixed homogenously prior to the hydrolysis step. The amount of water used in the synthesis is considerably less as well. The addition of the water is done over a 10-minute period, after which the reaction mixture forms a gel. in the MSU-B synthesis, the alkoxide is added to an aqueous solution containing the surfactant. The rate of hydrolysis and precipitation in this procedure is faster and a precipitate rather than a gel is formed. Accordingly, the morphologies of the two aluminas afforded are different. Amorphous MSU-X exhibits a wormhole motif where the crystalline MSU-B has a scaffold morphology. It is believed that the faster hydrolysis and precipitation in the MSU-B reactions are responsible for the 55 difference in morphology. In the MSU-B synthesis, boehmite particles are formed first. These particles are directed to form a mesostructure and hydroxyls between individual boehmite particles are condensed upon calcination, leading to the scaffold structure. In this manner, the surfactant is not used as a rigid template, but as a generic porogen used to impart increased porosity and surface area to the alumina. Under the controlled hydrolysis procedure used in the MSU-X synthesis, alkoxide molecules have time to interact with the surfactant molecules prior to hydrolysis, allowing for the formation of the amorphous wormhole structure. In this manner, the formation of MSU-X can be considered a templating mechanism whereas the formation of MSU-B is governed by a particle assembly mechanism. Particle assembly syntheses leading to other mesostructured oxides have been reported, including Sn02'9, ZrOzzo, and mixed metal oxides”. In the synthesis of mesostructured y-Al203 from aluminum salt precursors, it is believed that the alumina phase first formed has amorphous framework walls like MSU-X, which are converted to the crystalline boehmite walls of MSU-SIB upon thermal aging”. However, in the syntheses of MSU-S/B from aluminum alkoxide as reported here, no evidence exists to support such a claim. Inspection of the wide angle powder XRD patterns of the MSU-SIB shows that at least in a disordered state, formation of a boehmite phase occurs in the first minute after addition of the alkoxide to the surfactant solution. Within 4 hours, the boehmite phase is well organized and within 12 — 24 hours, the reaction and condensation is complete, as shown in Figure 2.11. 56 lntensi 1m rillllllllllfilTIITIIIIITIIITTYITIjT] 1o 20 30 4o 50 60 7o 80 Degrees (20) Figure 2.11. Wide angle XRD patterns of MSU-SIB prepared from aluminum sec-butoxide and dodecylamine at 20 °C showing evolution of the boehmite phase over time. The presence of weak reflections assignable to boehmite after only 1 minute show that the formation of MSU-B does not proceed through an MSU-X type of intermediate with amorphous framework walls. 57 2.5. Conclusions Reported here is what is believed to be the first mesostructured boehmite material, MSU-B, prepared from the hydrolysis of aluminum sec-butoxide in the presence of amine surfactants as porogens. The key in the synthesis strategy is the formation of the MSU-SIB surfactant/boehmite precursor using surfactants that are easily vaporized at low temperature calcination. it has been shown that these materials are synthesized through a particle assembly pathway as opposed to the templating mechanisms often used in the preparation of mesostructured materials. Additionally, an MSU-X intermediate is not formed, as boehmite is precipitated immediately. This affords products having a scaffold morphology with an open framework comprised of boehmite walls. This open framework allows for the realization of materials having surface areas of up to 497 m2/g, pore sizes of up to 6.8 nm, and pore volumes of up to 0.95 cc/g. The materials are stable to calcination at 350 °C without conversion of boehmite to transition alumina phases. Although surfactant can be removed by ethanol extraction, calcination is the preferred method as it ensures that the inorganic framework is boehmitic in nature. Further evidence that the MSU-B materials consist of walls of the boehmite phase was provided from the lack of the ability to peptize these materials. The textural properties of MSU-B materials could be controlled only slightly by altering the synthesis parameters. Increasing the mass of surfactant used in the surfactant solution leads to increased surface areas and pore volumes. Surface areas are relatively independent of the reaction temperature, but pore volumes 58 decrease with increasing temperature, signifying that less surfactant is incorporated into the framework at higher temperatures. At a constant aluminum to surfactant molar or mass ratio, however, increasing the surfactant size does not correlate to an increase in pore size or volume. In fact, pore volumes and pore sizes remain nearly constant over this series of experiments. The textural properties of MSU-S/B materials are not controlled by surfactant size. Also, the mass of surfactant used in any given synthesis does not match the observed pore volume of the surfactant free product. Additionally, the peak in the low angle region of the powder XRD patterns is not reproducible. From these observations, it is concluded that MSU-S/B surfactant-boehmite compositions are formed through a particle assembly pathway rather than a supramolecular templating pathway. 2.6. Future Directions Pseudoboehmite is actually the industrially preferred form of the aluminum oxide hydroxide because it is conveniently dispersed into a sol and easily introduced into other formulations or used in other applications. While the present work has focused on the synthesis of boehmite and avoidance of pseudoboehmite, a method to produce pseudoboehmite from a less expensive precursor would be of considerable interest. Similarly, it would be interesting to synthesize MSU-B materials from other aluminum sources as well. Although pseudoboehmite can be formed without the aid of a surfactant, it would interesting to know if a mesostructured (pseudo)boehmite sol could be 59 prepared. Another application of pseudoboehmite is that of a vaccine adjuvant23. A mesostructured sol may provide for increased loading of vaccine, effectively lowering the amount of aluminum administered per dose. Boehmites are more often used as precursors to catalysts than as catalysts themselves. Finding suitable catalytic applications for MSU-B would be a logical direction in which to proceed. Since these materials are stable only under relatively lower temperatures, they may be of utility as catalyst supports for organic conversions under mild conditions. There also exists another aluminum oxide hydroxide, a-AIOOH, diaspore. This polymorph is formed at high temperatures and under high pressures. If it is feasible to form a mesostructured diaspore, then this would also be of great interest. Because diaspore is structurally similar to a-Al203, corundum, it is converted directly to corundum at 550 °C. Prepared from other aluminas, temperatures in excess of 1000 °C are required for this transformation. Thus a mesostructured diaspore may lead to a mesostructured G-AI203 phase with significantly improved textural properties. Finally, while the target of the syntheses described in this work was mesostructured boehmite, it is possible to calcine MSU-SIB (and MSU-B) materials at elevated temperatures, converting them to mesostructured y—Al203. This is one of the synthetic approaches discussed in the following chapter. 60 2.7. References (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11) (12) (13) (14) Misra, C. American Chemical Society Monograph, Vol. 184: Industrial Alumina Chemicals, 1986. Stamires, D.; Brady, M. F.; Jones, W.; Kooli, F. In US; (Akzo Nobel N.V., Neth.). Us, 2002, pp 12 pp., Cont.-in-part of US. Ser. No. 21,839, abandoned. Callender, R. L.; Harlan, C. J.; Shapiro, N. M.; Jones, C. D.; Callahan, D. L.; Weisner, M. Fl.; MacQueen, D. B.; Cook, Fl.; Barron, A. Fl. Chemistry of Materials 1997, 9, 241 8-2433. Wefers, K.; Misra, C. "Oxides and Hydroxides of Aluminum,” Alcoa Laboratories, 1987. Yoldas, B. E. Am. Ceram. Soc., Bull. 1975, 54, 289-290. Baker, B. R.; Pearson, R. M. J. Catal. 1974, 33, 265-278. Guzman-Castillo, M. L.; Bokhimi, X.; Toledo-Antonio, A.; Saimones- Blasquez, J.; Hernandez-Beltran, F. Journal of Physical Chemistry B 2001, 105, 2099-2106. Bagshaw, S. A.; Pinnavaia, T. J. Angew. Chem, Int. Ed. Engl. 1996, 35, 1102-1105. Vaudry, F.; Khodabandeh, S.; Davis, M. E. Chem. Mater. 1996, 8, 1451- 1464. Yada, M.; Kijima, T. Zeoraito1997, 14, 104-111. Valange, S.; Guth, J. L.; Gabelica, 2. Stud. Surf. Sci. Catal. 1998, 117, 517-518. Cabrera, 8.; El Haskouri, J.; Alamo, J.; Beltran, A.; Beltran, D.; Mendioroz, S.; Marcos, M. D.; Amoros, P. Adv. Mater. (Weinheim, Ger.) 1999, 11, 379-381. Gonzalez-Pena, V.; Diaz, l.; Marquez-Alvarez, C.; Sastre, E.; Perez- Pariente, J. Microporous Mesoporous Mater. 2001, 44-45, 203-210. Zhang, Z.; Hicks, R. W.; Pauly, T. R; Pinnavaia, T. J. Journal of the American Chemical Society 2002, 124, 1592-1593. 61 (15) (16) (17) (18) (19) (20) (21) (22) (23) Brunauer, S.; Emmett, P. H.; Teller, E. J. Am. Chem. Soc. 1938, 60, 309- 319. Barrett, E. P.; Joyner, L. G.; Halender, P. P. Joumal of the American Chemical Society 1951 , 73, 373. Sing, K. S. W.; Everett, D. H.; Haul, R. A. W.; Moscou, L.; Pierotti, R. A.; Rouquerol, J.; Siemieniewska, T. Pure Appl. Chem. 1985, 57, 603-619. Rouquerol, F.; Rouquerol, J.; Sing, K. Adsorption by Powders and Porous Solids; Academic Press: London, 1999. Severin, K. G.; Abdel-Fattah, T. M.; Pinnavaia, T. J. Chemical Communications (Cambridge) 1998, 1471 -1472. Pacheco, G.; Zhao, E.; Garcia, A.; Sklyarov, A.; Fripiat, J. J. Journal of Materials Chemistry 1 998, 8, 219-226. Wong, M. S.; Jeng, E. S.; Ying, J. Y. Nano Letters 2001, 1, 637-642. Zhang, Z.; Pinnavaia, T. J. Journal of the American Chemical Society 2002, submitted. Hem, S. L.; White, J. L. PHARMACEUTICAL BIOTECHNOLOGY FIELD 1995, 6, 249-276. 62 Chapter 3 Synthesis and Characterization of Mesostructured y-A1203 3.1. Introduction Transition aluminas are disordered crystalline phases formed through the thermal dehydration of aluminum hydroxides and oxyhydroxides‘. These oxides are used as adsorbents and catalysts or catalyst supports in many chemical processes, including the cracking, hydrocracking, and hydrodesulfurization of petroleum feed stockse. At least seven transition aluminas have been reported. Of these, y-Al203 is perhaps the most important tor catalytic applications. The utility of y-Al203, as well as other transition aluminas, can be traced to a favorable combination of textural properties (i.e., surface area, pore volume and pore size) and acid - base characteristics. The structure of y—Al203 is based on that of a tetragonally distorted defect spinel (MgAl204) structure. The unit cell of spinel contains 8 formula units. The 32 oxygen atoms are arranged in a cubic close packed lattice. in spinel, the 8 aluminum ions occupy octahedral positions and the 8 magnesium ions occupy tetrahedral positions. The presence of trivalent aluminum ions (in y—Al203) in lieu of divalent magnesium ions (in spinel) necessitates that only 5 1/3 of the original 8 tetrahedral sites be occupied per unit cell in the y—AI203 structure. Overall, the ratio of the number of aluminum ions in octahedral to tetrahedral sites is 3:1. 63 Recently, density functional theory calculations have revealed the presence of hydrogen in the spinel lattice of y-Al2033. The y-Al203 phase is formed upon the dehydration of the aluminum oxidehydroxide boehmite at temperatures ranging from 400 to 700 °C. During this process, porosity is generated as water is eliminated from the structure, creating cracks and void space in the alumina. Along the progression from hydroxide phases to oxide phases, there is an initial rise in surface area and porosity. However, heating at temperatures above 500 °C leads to diminished textural properties. Conventional forms of y—Al203 typically exhibit a BET surface area below 250 mZ/g and a pore volume less than 0.50 cc/gz. The properties of y- Al203 for applications in catalysis and adsorption are determined in large part by these textural parameters. In view of the recent advancements realized for mesostructured forms of silicas4'5, similar improvements may be anticipated in the properties of mesostructured aluminas. However, comparatively limited progress has been made in this area. Since Bagshaw’s initial disclosure of the supramolecular synthesis of mesostructured alumina from aluminum alkoxides and PEG surfactantss, others have synthesized similar materials using both ionic7'9 and 10’“ surfactants as structure directors. A common feature of all these non-ionic mesostructured aluminas is the amorphous framework walls that limit their thermal and hydrothermal stability, greatly compromising their usefulness in potential catalytic applications. Recently, a mesostructured alumina containing a transition alumina phase has been reported”. However, the true composition of this material remains unclear without evidence of a boehmite precursor, which, as described below, is the crucial step leading to the formation of mesostructured y-AI203. This work reports on the synthesis, characterization, and stability of crystalline mesostructured y-aluminas prepared from the hydrolysis of aluminum sec- butoxide in the presence of either PEO or amine surfactants as porogens. The novelty of the synthesis is the resulting formation of a mesostructured surfactant/boehmite composition, denoted MSU-SIB. Upon thermal transformation, MSU-SIB is converted to crystalline, mesostructured y-Al203, denoted MSU-y. It will be demonstrated that these materials are structurally stable and afford textural properties exceeding those of conventional transition aluminas. 3.2. Experimental 3.2.1. Synthesis The assembly of the MSU-SIB surfactant-boehmite mesostructure was achieved by hydrolysis of aluminum sec-butoxide in an alcoholic aqueous solution of the desired surfactant. Syntheses could be performed over a wide range of reaction conditions. Two general routes to mesostructured y-Al203 are identified in sections 3.2.1.1 and 3.2.1.2 below. The surfactants selected for study are as follows: 65 Surfactant Formula Pluronic P84 (EO)19(PO)39(EO)19 Pluronic P65 (EO)19(PO)30(EO)19 Pluronic L64 (EO)13(PO)30(EO)13 Pluronic L61 (EO)2(PO)30(EO)2 Tergitol 15-s-9 C7H15CH2((EO)90H)CyH15 Tergitol 15-S-12 C7H15CH2((EO)120H)CyH15 3.2.1.1. Synthesis from PEO Surfactants A typical MSU-y synthesis using a PEO surfactant as a porogen was carried out as follows. Pluronic P84 and aluminum sec-butoxide were dissolved in 2- butanol. To this solution water was added to hydrolyze the alkoxide. The solution was stirred on a magnetic stirrer until homogenous, sealed in a glass vessel, and placed in a 100 °C oven under static conditions for a period of ~20 hours. The resulting MSU-S/B was recovered by vacuum filtration, washed with 2-butanol and water, and air dried. Conversion to MSU-y was achieved by calcining the as- synthesized MSU-S/B at 500 “C for 4 hours. Alternatively, during the synthesis, the water, alcohol, and surfactant could be mixed first, adding the alkoxide last. Reaction mixtures in syntheses involving lesser amounts of solvent would become quite thick. in these instances, a laboratory grade blender was used to mix the reactants. 66 3.2.1.2. Synthesis from Amine Surfactants and MSU-B Materials A second method used to synthesize MSU-y materials is to calcine either the MSU-SIB surfactant-boehmite compositions made from amine surfactants or surfactant free MSU-B boehmites, described in the previous chapter, at temperatures ~500 9C. Both methods afford mesostructured MSU-y aluminas. 3.2.1 .3. Synthesis of Doped MSU-y Aluminas MSU-y aluminas were also doped with metal ions. In these syntheses, the desired amount of dopant was administered as a salt which was dissolved in the 2-butanol and water surfactant solution. La(N03)3 - 9 H20 and (NH4)2Ce(N03)6 were used to supply La and Ce, respectively, in the doping reactions. The processing of the reaction mixtures was performed as described in section 3.2.1.1 above. 3.2.2. Characterization of MSU-y Aluminas Powder diffraction patterns were collected on a Rigaku Rotoflex Diffractometer using Cu K, radiation (A = 0.154 nm). Intensity was calculated in counts per second every 0.02 degrees 26. Sweep rates were varied from 2 - 8 degrees ZBlmin. Nitrogen adsorption - desorption isotherms were collected on a Micromeritics TriStar 3000 sorptometer at 77K. Samples were degassed at least 12 hours at 150 °C under a vacuum of 10’6 torr prior to analysis. Surface areas and pore sizes were determined by the BET13 and BJH“ methods, respectively. 67 TEM images were collected on a JEOL 1OOCX microscope with a 0936 filament at an accelerating voltage of 120 kV. Samples were prepared by sonicating a suspension of the powder in ethanol for 20 min. One to two drops of the suspension were then placed on a carbon coated holey film that was supported on a 3mm, 300 mesh copper grid. 27Al NMR spectra were collected on a Varian 400 VRX solid — state NMR spectrometer. Samples were placed in a zirconia rotor and spun at 4 kHz with a pulse width of 0.5 ps and a pulse delay of 0.5 s. Chemical shifts were referenced to AI(H20).33+ which is assigned a value of 0 ppm. Steam testing was performed by passing 20% H20 in nitrogen over the alumina at 600 ‘-’C for 5 hours. Exposed aluminas were air dried before further analysis. 3.3 Results and Discussion 3.3.1. MSU-y Aluminas from PEO Surfactants 3.3.1.1. General Properties of Materials Representative wide angle powder XRD patterns of an as-synthesized MSU- S/B surfactant-boehmite composition made from aluminum sec-butoxide and P84 and the corresponding MSU-y alumina are presented in Figure 3.1. In the case of these materials, peaks assignable to boehmite and y-Al203 for MSU-SIB and MSU-y, respectively, are evident. As was the case for the surfactant-free MSU-B boehmite materials described in the previous chapter, a low angle diffraction 68 MSU-y 440 311 40° Intensity MSU-SIB illlllfiTlrrlrlIIIWTIUWITIIYTjIUIIII 10 20 30 4o 50 so 70 so Degrees (26) Figure 3.1. Representative wide angle powder XRD patterns of an MSU-SIB surfactant-boehmite composition made from aluminum sec-butoxide and the PEO surfactant P84 under reaction conditions described in Table 3.1, line 3 and the corresponding MSU-y alumina prepared from calcination of the as- synthesized MSU-S/B at 500 9C for 4 hr. 69 peak in the XRD patterns was not always observed. However, when low angle peaks were observed, the d-spacings of these peaks have recurring values in multiples of ~8 A. The most common values, 48 and 56 A, correspond to an average of 6 to 7 building units in the particle. Nitrogen adsorption — desorption isotherms and pore size distributions of a representative MSU-y alumina and a similar alumina prepared without surfactant (blank) are shown in Figure 3.2. it can be seen that the pore volume and pore size are increased in the MSU-y as compared to the material prepared in absence of surfactant. Porosity in conventional y-AI203 is generated as cracks and voids form upon the dehydration of the oxidehydroxide precursor, so it is reasonable to expect some adsorption in the blank sample. However, use of surfactants as porogens imparts considerably larger pore volumes and sizes, as is reflected in the isotherms. Typical surface areas, pore sizes, and pore volumes for MSU-y aluminas range from 300 — 380 m2/g, 10 — 20 nm, and 0.75 — 1.6 cc/g, respectively. These same values for the blank sample are 259 m2/g, 5 nm, and 0.47 cc/g, which are in line with those expected for textural properties of conventional y-aluminas. The textural properties of these materials made from PEO surfactants are summarized in Table 3.1. In addition to the XRD data, further evidence of the presence of y-Al203 in the framework walls was provided by the 27Al NMR spectrum of MSU-y. As described in section 3.1, the unit cell of y-Al203 contains aluminum ions in both octahedral and tetrahedral coordination, in a ratio of 3:1. Accordingly, 27AI NMR spectra of y- Al203 show peaks consistent with this distribution of aluminum ions”. 70 1 000 dVIdD - o 5 1o 15 20 25 so Pore Diameter (nm) Volume Adsorbed (cc/g) Blank PIP Figure 3.2. Nitrogen adsorption-desorption isotherm and BJH adsorption pore size distribution (inset) of an MSU-y alumina prepared from aluminum sec- butoxide and the PEO surfactant P84 as described in Table 3.1, line 3. included for comparison is the same data for an alumina prepared without surfactant as described in Table 3.1, line 20. Isotherms are offset 200 cc/g for clarity and PSD curves are displayed in the same order as the isotherms. 71 Table 3.1. Synthetic parameters and textural properties of MSU-y aluminas prepared from aluminum sec-butoxide and PEG surfactants. Al/Surl/s-BuOH/HZO sag Pore Size” Pore Volume Surfactant Molar Ratioa (m2/g) (nm) (cc/g) 1 P84 100: 1 :27 : 544 334 15 1.17 2 P84 75 : 1 :27 : 544 332 17 1.23 3 P84 66: 1 :27 : 544 378 13 1.09 4 P84 50 : 1 :27 : 544 365 15 1.25 5 P84 25 : 1 :27 : 544 366 17 1.39 6 P84 66 : 1 :249 : 544 339 10 0.76 7 P84 66 : 1 : 160 : 544 341 13 0.93 8 P84 66 : 1 : 107 : 544 381 4, 17 0.84 3 P84 66 : 1 :27 : 544 378 13 1.09 9 P84 66 : 1 : 12 : 544 342 14 1.06 10 P84 66 :1 :0 : 544 325 10 0.92 3 P84 66: 1 :27 : 544 378 13 1.09 11 P84 66 : 1 :46 : 933 355 11 1.14 12 P84 66 : 1 :65 : 1322 338 11 1.04 13 P84 66:1 :103 2100 338 12 1.12 14 P84 66: 1 :218 : 4433 314 9 0.85 15 15-S-9 66 : 1 : 17 : 285 367 20 1.63 16 15-S-12 66 : 1 : 14 : 232 368 20 1.66 17 L61 66 : 1 :41 :670 279 12 0.78 18 L64 66 :1 :70 : 1160 300 10 0.76 19 P65 66 : 1 :83 : 1360 324 6 0.69 20 None 66 : 0 : 65 : 1322 259 5 0.47 ‘ Due to the large molar masses of PEO surfactants, molar ratios of aluminum alkoxide to PEO surfactants are much higher than when amine surfactants are used. The mass ratios, however, are comparable (section 3.3.1.2). b BJH pore diameter as measured by the adsorption branch 72 The 27Al NMR spectrum of MSU-y is provided in Figure 3.3. Two resonances are observed, one at 72 ppm and another at 6.7 ppm, corresponding to the four- and six- coordinate aluminum, respectively. From the relative intensity of these peaks, it can be seen that there is good agreement between the observed tetrahedral occupancy with the theoretical values. Furthermore, in the NMR spectra of amorphous aluminas, an additional resonance at ~35 ppm is often observed, which is attributable to 5- coordinate aluminum centers. The lack of any such resonance in the spectrum of MSUJY is evidence that no appreciable amount of amorphous alumina exists in this material. Perhaps the best evidence for the confirmation of a mesostructured alumina with crystalline walls is attained from TEM. Bright field and dark field images of MSU-y prepared from P84 and aluminum sec-butoxide are shown in Figure 3.4. The same fiber like morphology evident in MSU-B is also seen here. This indicates that there is retention of morphology on going from the boehmite to the gamma phase. From the image in Figure 3.4 A, it appears that the randomly oriented fibers have a length of approximately 30 nm and an average thickness of 5 nm. The porosity of MSU-y alumina is defined by the space between the overlapping fibers. In part B of Figure 3.4, a lower magnification image and the corresponding selected area electron diffraction pattern is provided for an MSU-y prepared from aluminum sec-butoxide and P84 as described in Table 3.1, line 3. The presence of the rings in the diffraction pattern, corresponding to hkl reflections of [311], 73 6.72 71 .96 I 11" [III 1111 III III] I I I I W IIII II IIIIIIIII 200 150 100 50 0 -50 -100 -150 -200 mm Figure 3.3. 27AI NMR of calcined (500 9C) MSU-y alumina prepared from aluminum sec-butoxide and P84 as described in Table 3.1, line 3. The two resonances at 6.7 ppm and 72 ppm are assigned to aluminum in octahedral and tetrahedral coordination, respectively. The peaks are in the approximate 75:25 ratio consistent with y-Al203. 74 Figure 3.4. Bright field TEM image of (A) calcined (500 °C) MSU-y alumina prepared from aluminum sec-butoxide and P84 as described in Table 3.1, line 3. (B) Lower magnification image of the same product and the corresponding electron diffraction pattern, with diffuse rings assignable to the [311], [400], and [440] crystallographic planes. (C) Dark field image of the same particles as in (B), but formed from an ~7° arc of the [440] diffraction ring. The crystal orientations contributing to the [440] ring appear as bright sports in the image. 75 76 [400], and [440], is a direct consequence of the polycrystalline nature of the material. Their presence confirms the crystalline nature of these materials. Dark field imaging in TEM is accomplished by using only diffracted electrons to form images. Additionally, an aperture can be inserted into the electron beam to select only electrons that are diffracted by specific crystallographic planes. Part C of Figure 3.4 displays such an image, the dark field image of the same particles as in Part B, created from ~79 arc of the [440] diffraction ring of MSU-y. The crystal orientations contributing to the [440] ring of the diffraction pattern appear as bright spots in the dark field image. The appearance of these bright spots uniformly across the image confirms that y—Al203 nanoparticles comprise the framework walls in MSU-y. 3.3.1 .2. Variations of Synthetic Parameters As was discussed in the previous chapter for mesoporous boehmites prepared from aluminum sec-butoxide and amine surfactants, these new MSU aluminas are synthesized through a particle assembly pathway. Consequently, controlling the textural properties of the products as is practiced with templated mesostructured materials is not straightfonrvard. Although the pore size distributions do not respond to differences in surfactant size, the y-Al203 compositions reported here can be made over a range of synthetic conditions while exhibiting textural properties in excess of for conventional aluminas. The textural properties are summarized in Table 3.1. 77 The aluminum to surfactant ratio was altered to range from ratios Al/P84 = 100 to Al/P84 = 25. It should be noted that the molar ratios of Al to surfactant employed in the syntheses leading to the formation of MSU-y aluminas are considerably greater than in the syntheses leading to the formation of MSU-B boehmites. The molar masses of the PEO surfactants are generally much larger than those of the amine surfactants. While the reported molar ratios differ greatly, the mass ratios of the two reagents are comparable. For example, at AVP84 = 66, the mass of aluminum sec-butoxide to the mass of P84 = 3.91. For comparison, an AVDDA = 5 corresponds to an aluminum sec-butoxide to DDA mass ratio of 6.65. While the textural properties of the alumina prepared at Al/Surf =66 appear to be out of line, there is a general increase in both surface area and pore volume with decreasing Al/Surf ratios (Table 3.1, lines 1-5). There is no apparent trend in pore size. While aluminas prepared at other ratios appear better, most syntheses were carried out at an Ai/Surf = 66 ratio for two reasons. First, reaction mixtures at high surfactant concentrations (low Al/Surf ratios) tend to be very thick, sticky, and difficult to process. Second, at the time of this work, the origin of the low angle diffraction peak was unclear, and aluminas exhibiting a low angle peak were expected to have more desirable textural properties than those that did not. The differences in textural properties for reaction products that exhibit a low angle XRD reflection and those that do not are not substantial. Perhaps there exists a certain ratio where there is an optimum coverage of surfactant by alumina, above which any additional alumina is not mesostructured into new 78 fibers, but rather goes toward increasing the thickness of existing MSU-y fibers. Void space that would contribute to pore volume would, above this ratio, be sacrificed as alumina fibers thickened. Nevertheless, syntheses carried out at this ratio proved fruitful. Recently, the synthesis of y-Al203 nanofibers prepared from NaAlOz and PEG surfactants was reported”. Molar ratios of aluminum to surfactant in syntheses leading to the y-AI203 fibers were ~ 2:1, much lower than the ratios employed in the current work. Although no claims of a mesostructure were made, fibers of comparable length and thickness to those reported here were obtained. Additionally, the large amount of surfactant used in the syntheses provided for the formation of an alumina with pore volumes in the range of 1.41 - 1.95 cc/g. A direct relationship between mass of surfactant used in the synthesis and pore volume of the resulting product was noted in Zhu’s work. These findings are not in agreement with observed trends in the syntheses of MSU-B boehmites reported in the previous chapter and in the syntheses of MSU-y aluminas, as described in the following paragraphs. Solvent polarity was also adjusted to examine the effect on the textural properties of MSU-y aluminas. From the nitrogen adsorption - desorption isotherms and pore size distributions in Figure 3.5 and the data in Table 3.1, lines 6 — 11, it can be seen that the textural properties go through a maximum for aluminas prepared at a water: 2-butanol ratio of 4 : 1 (80/20 v/v). It could be that as the polarity decreases at high a 2-butanol content, the surfactant ceases to form micelles capable of directing the structure. At the other extreme, of low 2- 79 8 ‘ S 0%I In 2000 - /77 a _ /////‘19% B d 0 5 10 15 20 25 30 :/// ,f/ 3 : Pore Diameter (nm) .._...,.- , 20% .8 1500‘ P I L O m '5 < 0) E 2 O > PIP Figure 3.5. Nitrogen adsorption-desorption isotherms of MSU-y aluminas prepared from aluminum sec-butoxide and P84 as described in Table 3.1, lines 3 and 6-10 showing the effect of solvent polarity on the textural properties. The volume percent of 2-butanol in the solvent is labeled for each isotherm. isotherms are offset 300 cc/g for clarity. The BJH adsorption pore size distributions (inset) are in the same order as the isotherms. 80 butanol content, the total amount of solvent may be decreasing to point that is not conducive to mesostructure formation. Nitrogen adsorption - desorption isotherms and pore size distributions for MSU-y aluminas prepared from aluminum sec-butoxide and P84 solutions of varying concentration are shown in Figure 3.6. It was observed that the surface areas, pore volumes, and pore sizes decreased slightly with decreasing concentration of the surfactant employed in the synthesis. The effect of surfactant concentration is most pronounced at a concentration of 5%. Aside from ease of processing the reaction mixture, there seems to be little advantage in diluting the reaction mixture. The surfactant concentration does not appear to be a critical factor affecting the textural properties and pore structure. This is the same phenomenon that was observed in the syntheses of MSU-SIB precursors and MSU-B boehmites from aluminum sec-butoxide and amine surfactants as described in the previous chapter. Finally, MSU-y aluminas could be prepared with other nonionic PEO surfactants in which Figure 3.7 shows the nitrogen adsorption — desorption isotherms and the corresponding pore size distributions for these aluminas. It seems odd that the smallest molecular weight surfactants yield aluminas with the largest pore volumes and sizes. However, in no case does the amount of surfactant used in the synthesis correspond to the pore size or volume observed in the surfactant free products. Again, the lack of a direct relationship between surfactant properties and textural properties of MSU-y aluminas is evidence of a particle assembly pathway. 81 a N 2 . , [5% ’15 I / \ _ 3 150° 0W. v _ Pore Diameter (nm) _ 13 fl/x"/” 01 .9 L- O m '0 < 0) E 2 O > Figure 3.6. Nitrogen adsorption-desorption isotherms and BJH adsorption pore size distributions (inset) for MSU-y aluminas prepared from aluminum sec- butoxide and P84 as described in Table 3.1, lines 3 and 11 -14 showing the effect of surfactant concentration (mass P84 / mass solvent) on the textural properties. The mass percent of P84 is labeled for each isotherm. isotherms are offset 300 cc/g for clarity. The PSD curves are displayed in the same order as the corresponding isotherms. 82 2000 Volume Adsorbed (cc/g) PIP Figure 3.7. Nitrogen adsorption-desorption isotherms and BJH adsorption pore size distributions (inset) for MSU-y aluminas prepared from aluminum sec- butoxide and differing PEO surfactants as described in Table 3.1, lines 3 and 11- 14. isotherms are offset 200 cc/g for clarity. The PSD curves are displayed in the same order as the corresponding isotherms. 83 3.3.1.3. Effect of Calcination Procedure As described in the previous chapter, the lower boiling points of the amine surfactants allowed for their removal from the as-synthesized MSU-S/B surfactant-boehmite compositions at mild temperatures with the retention of the boehmite phase during MSU-B synthesis. Such is not the case with the PEO surfactants. Temperatures above 400 9C are required for surfactant removal via calcination, causing the boehmite in the MSU-SIB hybrid composition to be converted to y—Ai203. y-Al203 is but one of at least seven identified transition aluminas. Calcination at elevated temperatures may be a general route to transition aluminas. The transformation sequence of thermal dehydration of aluminas dictates that 6-Al203 is the next phase to form once y-Al203 is heated at temperatures above ~750 QC‘. MSU-SIB aluminas prepared from PEO surfactants were calcined at 800 9C for 2hr to achieve the conversion to a mesostructured 6-Al203, denoted MSU-6. This transition alumina also has a structure based on the spinel structure and consequently, evolution from y-A|203 to 6-Al203 can be performed without significant changes in the pore structure of the alumina‘m’. Therefore, it is not unreasonable to expect the retention of desirable textural properties in these materials, as evidenced from the nitrogen adsorption — desorption isotherms (Figure 3.8). Due to the similarities in 6- and y-AI203, it is somewhat difficult to distinguish between these two transition aluminas. XRD diffraction can be used to some extent to differentiate phases, but the patterns of all transition aluminas are strikingly similar. A better tool is to use is 27Al NMR, however in this case, site 84 occupancy of Al3+ ions is nearly identical. The inset of Figure 3.8 provides the wide angle XRD pattern of the reaction product formed through the calcination of an MSU-SIB surfactant-boehmite composition at 800 QC. The case could be made that these peaks are assignable to 6—Al203. A comparison of textural properties of these two mesostructured aluminas (Table 3.2) reveals that while the pore volumes and sizes are comparable, there is a marked decrease in the surface area. This could be due to sintering of y-Al203 fibers upon transformation to 6-Al203. The 27AI NMR spectra (not shown) looks nearly identical to that of the MUS-v alumina. Due to the lack of definitive proof of 0-A1203 phase formation, it could be argued that the material claimed here to be a mesostructured 6-AI203 may only be a y- Al203 whose pore structure is beginning to collapse. Regardless, this material exhibits exceptional textural properties having been calcined at 800 9C. 3.3.2. MSU-y Prepared from Amine Surfactants MSU-y aluminas could also be prepared from amine surfactant-boehmite compositions synthesized as described in the previous chapter. This was achieved by directly calcining MSU-S/B surfactant-boehmite precursors, or the surfactant-free MSU-B boehmite aluminas that had been previously calcined, at temperatures sufficient to initiate conversion to y-AI203 (>400 ‘2C). The resulting products exhibited XRD patterns consistent with y-AI203, signifying the crystalline nature of these materials. As expected, these materials have the same morphology as other MSU-y aluminas. The bright field TEM image and 85 700 'I 600- a: "a “C ‘2 -.s 500- 400- Volume Adsorbed (cc/g) MSU-8 10 20 30 40 50 60 70 00 Degrees (26) 0 I I I I I I I I I I I l I I I I I T T 0.6 0 0.2 0.4 PIP O 0.8 1 Figure 3.8. Nitrogen adsorption-desorption isotherm and wide angle powder XRD pattern (inset) of a calcined (800 QC) MSU-6. The initial MSU-SIB surfactant-boehmite intermediate was prepared from aluminum sec-butoxide and the PEO surfactant P84 under conditions described in Table 3.1, line 13. 86 Table 3.2. Comparison of textural properties of calcined MSU-y formed at 500 9C and calcined MSU-6 alumina formed at 800 90a Calcination . b Surface Area Pore Size Pore Volume Temperature ( 2/ ) ( ) (cc/ ) rn nm (QC) 9 9 1 500 306 18 1.14 800 251 16 0.98 2 500 338 12 1.04 800 249 16 1.00 ‘3 The initial MSU-SIB surfactant-boehmite intermediate was prepared from aluminum sec-butoxide and the PEO surfactant P84 under conditions described in Table 3.1, line 13. b BJH pore diameter as measured by the adsorption branch 87 corresponding electron diffraction pattern shown in Figure 3.9 for an MSU-y prepared from an amine-mediated MSU-B boehmite verified that an open framework was created upon the assembly of the small crystalline particles into a mesostructure. Nitrogen adsorption — desorption isotherms (Figure 3.10) also show behavior similar to MSU-y aluminas derived from PEO surfactants. The textural properties of these materials are summarized in Table 3.3. The pore volumes and pore sizes are approximately equal (at a given temperature) for MSU-y aluminas prepared directly from MSU-SIB or from previously calcined MSU-B. The surface areas are greater for MSU-y aluminas prepared through the calcination of MSU-SIB than for MSU-B precursors. Perhaps there is some structural degradation of MSU-B caused by adsorption of water onto the surface, rendering a poorer quality MSU-y to be formed upon a second calcination. it was observed that pore volumes for MSU-B materials decreased at higher aging temperature (Table 2.1). That trend is observed here as well. Finally, it is noted that data provided in Table 3.3 for the MSU-Iy aluminas from MSU-SIB and MSU- B indicate that the textural properties are enhanced compared to those in MSU-y aluminas prepared from PEO surfactants. 88 Fig boe COIT at a tem by c Dam IIIGS and Figure 3.9. TEM image of MSU-y prepared through the calcination of MSU-B boehmite at 500 °C for 4 hr. The initial MSU-S/B surfactant-boehmite composition was prepared from aluminum sec-butoxide and dodecylamine (DDA) at a reaction stoichiometry of Al:DDA:EtOH:H20 = 5:1:68.0:73.4 and a reaction temperature of 20 °C. The surfactant free MSU-B boehmite phase was obtained by calcining the MSU-SIB precursor at 325 °C for 10 hr. The electron diffraction pattern (inset) demonstrates the crystallinity of the framework of the mesostructure, as indicated by the diffuse rings attributable to the [311], [400], and [440] crystallographic planes of y—Ai203. 89 9O dVIdD a .1 3 1200 - 3 Porse Diameter (Iism) 'U . O .O - L O - 8 o . E 2 .1 O > - 400 - o 0 re .° 4:. 0 ca o on —L Figure 3.10. Nitrogen adsorption-desorption isotherms and BJH adsorption pore size distributions (inset) for MSU-y aluminas prepared through the calcination (500 9C) of amine-mediated MSU-S/B surfactant-boehmite precursors (samples A and C) and surfactant free MSU-B precursors (samples B and D). The AlzDDAzEtOHzH20 reaction stoichiometry was 5:1:68.0:73.4 for all samples. Reaction temperature was 20 9C for samples A and B and 45 ‘2C for samples C and D. Isotherms are offset 300 cc/g for clarity. The PSD curves are displayed in the same order as the corresponding isotherms. 91 Table 3.3. Textural properties of MSU-y aluminas prepared through the calcination of MSU-SIB surfactant-boehmite compositions and MSU-B boehmites at 500 9C.” Surface Area Pore Sizeb Pore Volume Precursor ("IQ/9) (nm) (cc/9) A MSU-S/B 440 7 1 .07 B MSU-B 385 7 0.97 C MSU-S/B 529 3 0.86 D MSU-B 431 4.5 0.89 a The MSU-S/B and MSU-B precursors were prepared from aluminum sec- butoxide and dodecylamine. The AlzDDAzEtOHzH20 reaction stoichiometry was 5:1:68.0:73.4 for all samples. Reaction temperature was 20 ‘2C for samples A and B and 45 9C for samples C and D. b BJH pore diameter as measured by the adsorption branch 92 3.3.3 Stability of MSU-y Aluminas 3.3.3.1 Doped MSU-y Aluminas it is known that the incorporation of certain elements into y—Al203 has a stabilizing effect on the structure. While transition metals have been doped into aluminas toward this end, more commonly either lanthanum or cerium is used as the dopant. It has been demonstrated that these ions stabilize the y-A1203 phase ”2‘ on the toward conversion to d-Al203 by forming a metal aluminate species surface of the alumina. Due to the nature of the spinel structure of y-Al203 and other transition aluminas, metal ions can also stabilize alumina by filling the vacancies in the iatticezz'za. Most studies are aimed at delaying the temperature at which y-Ai203 is converted to G-A1203 or other transition aluminas. Many important industrial catalytic applications, of alumina are performed in a steam environment, so aluminas must also be stable under these conditions if they are to function effectively. A study of the ability of lanthanum to stabilize alumina in a steam environment suggested that the lanthanide aluminate mechanism is at work here as well“. MSU-X aluminas were doped with rare earth metals to improve their thermal stability”. While MSU-y aluminas have proven to be thermally stable, improving their hydrothermal stability would increase the possibility for effective use in catalytic applications. It was also possible to dope the MSU-y aluminas of the present work with cerium and lanthanum. As shown in Table 3.4, MSU-y aluminas could be doped with an assortment of metal ions with the products 93 Table 3.4. Textural properties of doped MSU-y aluminas prepared from aluminum sec-butoxide and P84. in each synthesis, the dopant was administered as a metal salt which was dissolved in the surfactant solution prior to addition of the aluminum sec-butoxide. Surface Area Pore Sizea Pore Volume Dopant 2 (m /0) (nm) (cc/0) Cu 437 1 0 1 .05 Ni 426 1 0 1 .08 Cr 438 3-1 5 0.96 Co 390 3-1 5 0.88 Fe 380 1 0 0.93 Mn 392 7 0.81 Ba 322 1 3 0.91 La 386 7 0.68 Ce 41 6 9 0.99 8‘ BJH pore diameter as measured by the adsorption branch 94 exhibiting favorable textural properties. While the introduction of transition metals may introduce catalytically active sites to MSU-y aluminas, steam testing (below) was conducted for the lanthanum and cerium doped materials, keeping with conventional methodologies. It was concluded that in amorphous MSU-X aluminas, direct incorporation of rare earths into the framework was a feature of the formation. The method of incorporation in MSU-y, however, may involve both the formation of an aluminate coating and the incorporation of the dopant into vacancies of the y-Al203 lattice. Peaks assignable to a CeAlOa phase in the wide angle XRD pattern of 5% cerium doped MSU-y indicates that formation of an aluminate surface species is involved during the incorporation process (Figure 3.11). However, no such peaks are evident in the pattern of the 5% lanthanum doped material, indicating that direct substitution into the spinel lattice of y-Ai203 may also be involved. Nitrogen adsorption — desorption isotherms (Figure 3.12) indicate that the pore volume of MSU-y is greater in the pristine sample than in either doped variation, although the pores are smaller and more uniform in the doped materials. The TEM images of these materials (Figure 3.13 A) reflect this as well; larger pores are evident in the clean sample (A) than in the lanthanum doped (C) or cerium doped (E) sample. 3.3.3.2. Steam Testing of MSU-Iv Aluminas These three aluminas were tested for hydrothermal stability by exposing each to a 20% steam atmosphere at 600 SC for 5 hr. Interestingly, the pristine MSU-y 95 intensity Pristine jIII rIII IIII IIII IIII IIII IIII I I I I I I I 10 20 30 40 50 60 70 80 Degrees (26) Figure 3.11. Wide angle powder XRD patterns of MSU-y aluminas prepared from aluminum sec-butoxide and P84: pristine and doped with either lanthanum or cerium. Peaks marked with an asterisk are attributable to a cerium aluminate phase, CeAlOa. The doped MSU-y aluminas were prepared through the calcination (500 9C) of doped MSU-SIB precursors prepared as described in Table 3.1, line 13. 96 . 5% Ce 1400' a - 2 > . 1: A 1200- m - m E I 0 P10 DIO :0 40 50 z 1000- ore ame er (nm) 77 g I 5% La 3 u in 800- .0 . < 2 “E’ 600- 2 I g I 400- 200_ Pristine o I I I I I I I l l I I l I l I I I I I 0 0.2 0.4 0.6 0.8 1 P1P 0 Figure 3.12. Nitrogen adsorption-desorption isotherms and BJH adsorption pore size distributions (inset) of MSUJY aluminas: pristine and doped with either lanthanum or cerium. The doped MSU-y aluminas were prepared through the calcination (500 °C) of doped MSU-SIB precursors prepared as described in Table 3.1, line 13. isotherms are offset 400 cc/g for clarity. The PSD curves are in the same order as the corresponding isotherms. 97 alumina appears to have the best hydrothermal stability. it retains 85% of the original surface area while the lanthanum doped and cerium doped MSU-y aluminas retain only 57% and 55%, respectively. A major concern under these conditions is the sintering of the nanoparticles. A comparison of TEM images of the three samples before steaming (Figure 3.13) reveals that there is more void space between fundamental building units of the framework in the undoped alumina(A). it may be that the greater distance between particles keeps them from sintering. The particles of 5% La doped (C) and Ce doped (E) aluminas may be more easily sintered as they are initially in closer proximity. The TEM images of the steamed aluminas (B,D,F) support this view. While still lath-like in morphology, the particles appear shorter and thicker in all cases, but more so for the doped aluminas than for the clean sample. Sintering leads to loss of surface area, which is observed in all cases. it is possible that doping of aluminas leads to the formation of micropores in the structures, which contribute greatly to the surface area. Upon steaming and subsequent sintering, micropores are lost as larger pores are formed. It is noted that the average pore sizes in the doped aluminas increase after steaming (Figure 3.14). The y-Al203 phase is preserved for all materials. Electron diffraction patterns reveal the diffuse rings, confirming crystallinity of the frameworks. However, this Structural integrity of the y—Al203 phase is greatest in the pure alumina as judged by XRD patterns (Figure 3.15). While all samples exhibit peaks consistent with y- 98 Figure 3.13. TEM images of MSU-y aluminas prior to (A,C,E) and after steaming (B,D,F) at 600 °C for 5 hr in a 20% steam atmosphere. Images A and B are for pristine MSU-y. Images C, D and E, F are for MSU-y doped with 5% La3+ and 5% 09‘“, respectively. The MSU-y aluminas were prepared as described in the caption to Figure 3.11. The scale bar represents 100 nm in all figures. 99 ~ 5%.! .5 O O O l dVIdD a . E 800' 0 10 20 30 40 50 5%La v .. Pore Diameter (nm) '5 . 0 Q . '5 in 600 _. 'U . < . 0 E 1 3 400- O . > Pristine Figure 3.14. Nitrogen adsorption-desorption isotherms and BJH adsorption pore size distributions (inset) of pristine and doped MSU-y aluminas after steaming for 5 hr at 600 9C in a 20% steam atmosphere. Samples were prepared as described in the caption to Figure 3.11. Isotherms are offset 300 cc/g for clarity. The PSD curves are in the same order as the corresponding isotherm. lOl Intensi Pristine IIIIIIIIIIIIIIIIIIIIIIIIIIIIIITIIII1 10 20 30 40 50 60 70 80 Degrees(2®) Figure 3.15. Wide angle powder XRD patterns of pristine and doped MSU-y aluminas after steaming for 5 hr at 600 9C in a 20% steam atmosphere. Peaks marked with an asterisk are attributed to CeAi03, cerium aluminate. Samples were prepared as described in the caption to Figure 3.11. 102 A1203, the peak intensity is greatest in the pristine alumina. Peaks assignable to the CeAl03 phase are again evident for the cerium doped sample. Traditionally, lanthanides are stabilizers of aluminas toward conversion to d- Al203. in the case of MSU-y aluminas, they may also provide stabilization toward d-Al203, but the textural properties are compromised in comparison to the pristine alumina. Perhaps compared against a pure alumina synthesized without the aid of a surfactant, the lanthanides would improve hydrothermal stability. Ironically, as the pore structure is relatively unchanged after steaming, the most open framework appears to be the most stable. The results of the study are summarized in Table 3.5. 3.4. Conclusions Synthesis of the first mesostructured y—Ai203 has been accomplished through a particle assembly pathway from the hydrolysis of aluminum alkoxide in the presence of electrically neutral nonionic PEG and amine surfactants as porogens. The formation of an MSU-S/B precursor is key in. attaining these mesostructures. It has been demonstrated that these aluminas possess textural properties superior to those of conventional aluminas, exhibiting surface areas, pore sizes, and pore volumes in excess of 400 m2/g, 15 nm, and 1.00 cc/g, respectively. The lack of a correlation between the size, mass, or molar mass of surfactant employed in the synthesis with the textural properties of the surfactant 103 Table 3.5. Textural properties of MSU-y aluminas prior to and after treatment at 600 9C in a 20% steam atmosphere for 5 hr.al Surface Area Pore Sizeb Pore Volume Sample Condition 2 (m lg) (nm) (cc/0) Pristine Fresh 371 13 1.1 1 Steamed 31 6 1 2 0.85 5% La Fresh 386 7 0.68 Steamed 220 12 0.77 5% Ce Fresh 416 9 0.99 Steamed 228 14 0.77 a The samples were prepared as described in the caption to Figure 3.11. b BJH pore diameter as measured by the adsorption branch 104 free products is evidence that these materials are not assembled through a supramolecular templating pathway. The open framework resulting from the assembly of fundamental alumina fibers was verified through TEM imaging. MSU-y aluminas can be doped with transition metals and lanthanides without compromising the structure or surface area of the MSU-y alumina. Additionally, these materials show exceptional thermal and hydrothermal stability, up to 600 9C for 5 hr in a 20% steam environment. These properties make MSU-y excellent candidates for catalysts or catalyst supports. 3.5. Future Directions While one example of a mesostructured higher temperature transition alumina was shown here, it would be of interest to synthesize more of these phases. Pore structures in these materials seem to remain intact at high temperatures, leading one to inquire into the possibility of making a mesostructured d-Alzog, or at least an d-Al203 that demonstrates marked improvements in textural properties. As was the case for mesostructured boehmite, finding a suitable catalytic application for MSU-y is also of interest. However, noting that y-A1203 is perhaps the world’s most widely used catalyst support, the possibilities for making an impact in this area are enormous. One such endeavor is the topic of the next chapter. 105 3.6. References (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11) (12) (13) (14) Wefers, K.; Misra, C. "Oxides and Hydroxides of Aluminum," Alcoa Laboratories, 1987. Misra, C. American Chemical Society Monograph, Vol. 184: Industrial Alumina Chemicals, 1986. Sohlberg, K.; Pennycook, S. J.; Pantelides, S. T. Journal of the American Chemical Society 1999, 121, 7493-7499. Kresge, C. T.; Leonowicz, M. E.; Roth, W. J.; Vartuli, J. C.; Beck, J. S. Nature (London) 1992, 359, 710-712. 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.; et al. J. Am. Chem. Soc. 1992, 114, 10834-10843. Bagshaw, S. A.; Pinnavaia, T. J. Angew. Chem, Int. Ed. Engl. 1996, 35, 1102-1105. Vaudry, F.; Khodabandeh, S.; Davis, M. E. Chem. Mater. 1996, 8, 1451- 1464. Yada, M.; Machida, M.; Kijima, T. Chem. Commun. (Cambridge) 1996, 769-770. Valange, S.; Guth, J. L.; Kolenda, F.; Lacombe, S.; Gabelica, Z. Microporous Mesoporous Mater. 2000, 35-36, 597-607. Cabrera, 8.; El Haskouri, J.; Alamo, J.; Beltran, A.; Beltran, D.; Mendioroz, S.; Marcos, M. D.; Amoros, P. Adv. Mater. (Weinheim, Ger.) 1999, 11, 379-381. Yang, P.; Zhao, D.; Margolese, D. I.; Chmelka, B. F.; Stucky, G. D. Chem. Mater. 1999, 11 , 2813-2826. Gonzalez-Pena, V.; Diaz, l.; Marquez-Alvarez, C.; Sastre, E.; Perez- Pariente, J. Microporous Mesoporous Mater. 2001, 44-45, 203-210. Brunauer, S.; Emmett, P. H.; Teller, E. J. Am. Chem. Soc. 1938, 60, 309- 319. Barrett, E. P.; Joyner, L. G.; Halender, P. P. Journal of the American Chemical Society 1951, 73, 373. 106 (I5) (15) (17) (18) (19) (20) (21 ) (22) (23) (24) (25) John, C. 3.; Alma, N. C. M.; Hays, G. R. Appl. Catal. 1983, 6, 341-346. Zhu, H. Y.; Riches, J. D.; Barry, J. C. Chemistry of Materials 2002, 14, 2086-2093. Wolverton, C.; Hass, K. C. Physical Review B: Condensed Matter and Materials Physics 2001 , 63, 024102/024101-024102/0241 16. Krokidis, X.; Raybaud, P.; Gobichon, A.-E.; Rebours, B.; Euzen, P.; Toulhoat, H. Journal of Physical Chemistry B 2001, 105, 5121-5130. Schaper, H.; Doesburg, E. B. M.; Van Reijen, L. L. Appl. Catal. 1983, 7, 211-220. Beguin, B.; Garbowski, E.; Primet, M. Appl. Catal. 1991, 75, 119-132. Oudet, F.; Courtine, P.; Vejux, A. J. Catal. 1988, 114, 112-120. Mizukami, F.; Maeda, K.; Watanabe, M.; Masuda, K.; Sano, T.; Kuno, K. Stud. Surf. Sci. Catal. 1991, 71, 557-568. Braun, S.; Appel, L. G.; Zinner, L. B.; Schmal, M. Br. Ceram. Trans. 1999, 98, 77-80. Schaper, H.; Amesz, D. J.; Doesburg, E. B. M.; Van Reijen, L. L. Appl. Catal. 1984, 9, 129-132. Zhang, W.; Pinnavaia, T. J. Chem. Commun. (Cambridge) 1998, 1185- 1186. 107 Chapter 4 Hydrodesulfurization Over MSUdy Supported Catalysts 4.1. introduction The hydrodesulfurization (HDS) reaction presents a challenge in the oil refinement process. HDS is the removal of sulfur from sulfur-containing organic molecules by treatment with H2 at elevated temperatures and pressures. HDS has been performed in refineries since the 1950’s. Since that time, however, improvement of the process has been constantly targeted. When burned, fuels that are high in sulfur release SOx pollutants into the atmosphere, contributing to smog and acid rain. in addition, the presence of sulfur in automobile exhaust reduces the effectiveness of catalytic converters. Sulfur, if not removed, also has a detrimental effect on downstream catalysts used in the refinement process. Providing more motivation, the EPA has currently set up guidelines to restrict the amount of sulfur allowed in diesel products to 15 ppm by the year 2010‘. This corresponds to a 97% reduction over a ten year period. Complicating matters is that relatively clean crude oil stocks have been consumed, leaving only heavier crudes laden with an increased amount of sulfur. Clearly, advances in HDS technology must be made to satisfy both economic and environmental concerns. Transition metal sulfides (TMS) have proven to be effective catalysts for HDS. Within this group of compounds, there are select metals whose sulfides are particularly active. The position of the metal in the periodic table and the 108 morphology of the metal sulfide play a role in determining the relative catalytic activity in the HDS reaction. A number of review articles discuss periodic trends of TMS in HDS catalysisz's. Ruthenium disulfide is perhaps the most active TMS. However, molybdenum disulfide is the catalyst of choice for commercial use. The lower activity of M082 is compensated by its reduced cost. In addition, the activity of TMS can be promoted by the presence of another metal sulfide, commonly those of cobalt and nickel. The role of the promoter is not entirely clear, but it has been demonstrated that they increase catalytic activity by increasing the electron density on the primary metal or lowering the Mo-S bond strength9, allowing for the creation of more active sites once sulfur atoms are removed by hydrogen. Catalytically active sites on M082 are characterized by sulfur vacancies, or coordinatively unsaturated metal ions, at the edge planes of the TMS. Utilizing a support can help to disperse the catalyst and expose more of these edge planes. Studies of catalytic activity have been performed for many different supports 1012 13,14 ranging from carbon , mesoporous synthetic clays and aluminosiiicates15 17, and zeolites”. Still, alumina is the most common support employed for HDS catalysts. Alumina is stable under the conditions encountered during the reaction. It is inert; hence it does not interfere with the HDS reaction. Aluminas are inexpensive, are readily available, and exhibit desirable surface and textural properties. Many aluminas are porous, allowing for better dispersal of catalyst and increased access to the catalytically active sites. The transition alumina y- Al203 is one such alumina with the features described above; it is the most 109 common support used in HDS chemistry. In fact, many commercial catalysts are mixtures of molybdenum and cobalt sulfides on a y-Al203 support. The use of mesoporous supports may lead to improved pathways to sulfur removal. The increased surface area and pore volume in these materials should allow for better dispersal of the catalytic phases and better access to the active sites on these catalysts. Additionally, the pore sizes in these supports can accommodate larger sulfur containing substrate molecules such as dibenzothiophene-based molecules, which mimic more closely the sulfur bearing contaminates in real oil stocks. To this end, mesostructured MSU-y aluminas have been used in the present work as catalyst supports in HDS reactions. MSU- y aluminas exhibiting different particle morphologies have been recently synthesized‘g. Aluminas with scaffold or lath morphology could be attained from different aluminum containing precursors. Here, the difference in morphology is used as a basis for comparison of performance in HDS reactions. 4.2. Experimental 4.2.1. Sample Preparation The MSU-y aluminas of scaffold morphology were synthesized as described in the previous chapter. MSU-y with lath morphology was prepared similarly, but from hydrolysis by ammonia of either aluminum cations or the Al13 oligomeric cation. The MSU-y aluminas used in this study were prepared as follows. 110 4.2.1.1. MSU-y Exhibiting Scaffold Morphology A 10% mass solution of Pluronic P84, a nonionic PEO surfactant, was dissolved in 2-butanol and water (20/80 v/v) under stirring on a magnetic stirrer. To this solution, aluminum sec-butoxide was added and stirred until homogenous, to achieve an overall reaction stoichiometry Al:P84:s-BuOH:H20 of 66:1:103:2100. The reaction mixture was removed from stirring and allowed to age at 100 °C for 20 hr in a sealed vessel to yield the MSU-SIB surfactant- boehmite precursor. This intermediate was recovered by vacuum filtration, washed with water and 2-butanol, and air dried. The surfactant free MSU-y alumina was obtained through calcination of the MSU-SIB material at 500 °C for 4 hr. 4.2.1.2. MSU-y Exhibiting Lath Morphology The assembly of a MSU-SIB mesostructure was achieved by aging an aqueous mixture of a [Al13O4(OH)24(H20)12]Cl-, and P84 at 80 °C for 6 - 10 h and then hydrolyzing the mixture with concentrated NH4OH. The precipitate formed upon hydrolysis was aged at 80 °C for 6 h, then at 100 °C for 24h, and air dried. The resulting mixture of MSU-SIB and NH4C| was calcined at 325 °C for 3h, then at 550 °C for 4h, to form the surfactant free MSU-y alumina. 4.2.1 .3. Catalyst Loading The MSU-y alumina supports were loaded with catalyst by incipient wetness impregnation of molybdenum and cobalt salts followed by thermal degradation to 111 the oxides. The supports were loaded with the volume of solution necessary to fill 90% of the available pore volume. For example, to achieve a 6% loading of Mo on an alumina having pore volume of 1.0 cc/g, 0.90 mL (per gram of Al203) of a solution of ammonium heptamoiybdate tetrahydrate would be required. The amount of salt dissolved in this solution is determined by the desired loading. in this case, a 6% Mo loading would require that 0.06 gram of Mo per gram of A1203, or 0.110 gram of the Mo salt, be impregnated. The requisite volume of this solution was added dropwise to the MSU-y alumina, with intermittent agitation by either manual or mechanical shaking, to provide even wetting. The M0 impregnated alumina was calcined at 400 °C for 2 hr to decompose the salt to the oxide. The pore volume of this intermediate material was obtained through nitrogen adsorption. The cobalt was loading was determined to provide a 3:1 molar ratio of Mo:Co. Cobalt (ll) nitrate hexahydrate was impregnated onto the support in the same manner as described for the Mo loading. Another calcination at 400 °C for 2 hr yielded the catalyst in its non-active, oxidic form. Conversion to the catalytically active sulfidic form was achieved by heating the materials to 400 °C for 1 hr in a tube furnace while a stream of 10% H28 in nitrogen was passed over the material. In each case, this resulted in the formation of a black powder that was readied for use by pressing into a pellet. In this study, the lath and scaffold aluminas were separately loaded with 6%, 9%, and 12% M0 by mass. These samples are denoted by their morphology and loading. For example, the scaffold alumina with a 9% Mo loading is referred to as 112 Scaffold-9. The loading of the commercial Crosfield 465 catalyst was in the range of 10-20% M0 by mass. It was sulfided as received. 4.2.2. Reactor Design and Experimental Conditions The HDS pilot plant used in this study is housed in the Chemical Technology Division of Argonne National Laboratory, Argonne, IL. The automated, continuous-flow unit is made primarily of 316 stainless steel tubing. Liquid feed in all experiments was 0.8 wt% S as dibenzothiophene (DBT) in hexadecane. This feed is representative of the kind of sulfur present in a middle distillate oil. The gas feed is a hydrogen and nitrogen mixture. Both the liquid and gas feeds are separately pumped through the system, eventually mixing together prior to entering the preheater furnace. At this point, the liquid is vaporized and all feed exits the preheater in the gas state. The feed is next passed through the reactor furnace and over the catalyst bed. The bed contains approximately 19 of catalyst diluted with 29 SiC, resting between plugs of glass wool. The reaction products are swept out of the reactor and separated. Liquid products are condensed and collected while gas products, mainly H28, are passed through a NaOH scrubber solution before being vented. A schematic of the system is provided in Figure 4.1. The typical reaction conditions were as follows. All liquid lines were heated to 35 °C to prevent the hexadecane from plugging the line. The preheater furnace was set to 350 °C and the reactor furnace to 400 °C. The nitrogen flow was set at 200 cc/min and hydrogen at 500 cc/min. The system was purged and the liquid 113 Vent Catalyst V ‘— Preheater V Bed Liquid NaOH Collection Figure 4.1. Schematic of the HDS reactor employed in the study. The liquid feed of dibenzothiophene in hexadecane was introduced to the gas feed, a mixture of hydrogen and nitrogen. The two feeds are passed through the preheater (set to 350 °C), where all feed is vaporized. Then the feed reactants are passed over the catalyst bed at 400 °C. Liquid products are condensed and collected while gas products are scrubbed in 2M NaOH prior to venting. 114 feed was introduced. The system pressure was set to 400 psi and was controlled by a back pressure regulator. Liquid products were collected at desired intervals. 4.2.3. Characterization Catalysts were characterized by powder x-ray diffraction (XRD), nitrogen adsorption, and transmission electron microscopy (TEM). XRD patterns were obtained on a Rigaku Rotoflex diffractometer with Cu K, radiation of 1.54 A. Nitrogen adsorption — desorption isotherms were collected on a Micromeritics TriStar 3000 sorptometer at 77K. Samples were degassed at least 12 hours at 150 °C under a vacuum of 10'6 torr prior to analysis. Surface areas and pore sizes were determined by the BET” and BJH21 methods, respectively. TEM images were collected on a JEOL 1OOCX microscope with a CeBe filament at an accelerating voltage of 120 kV. Samples were prepared by sonicating a suspension of the powder in ethanol for 20 min. One to two drops of the suspension were then placed on a carbon coated holey film that was supported on a 3 mm, 300 mesh copper grid. Liquid collections from the HDS reactor were diluted in hexane for GC-MS analysis. The products were separated on a DBS—MS column and analyzed using an HP 5890 GC-MS Series it Plus. The conversion was determined as the percentage of DBT converted to products. The selectivity was defined as the percentage of product that is biphenyl. 115 4.3. Results and Discussion 4.3.1 . Catalytic Reactions in all cases, the product mixture of the HDS reactions contained biphenyl, cyclohexylbenzene (CHB), 4H-DBT (the ring hydrogenated product of DBT), and unreacted DBT. Reaction pathways are provided in Figure 4.2. From the conversion and selectivity data presented in Table 4.1, it can be seen that conversions roughly increase with metal loading. It is interesting that for both aluminas, the 9% loaded samples show a drop in conversion from that of the 6% loaded samples. As expected, conversions are highest for the 12% loaded samples, with a maximum of 87% for the scaffold-12 catalyst. Comparing selectivities, those of the scaffold catalysts are consistently higher than those of the lath catalysts. As the 12% loaded samples showed the highest conversions and were the most selective, these catalysts were studied in more detail. it should be noted, however, that both MSU-y catalyst at all loadings show results comparable or superior to those obtained from the commercial catalyst tested under similar conditions. The Crosfield catalyst had a conversion of 77% and a selectivity for biphenyl of 61%, for a percent yield of biphenyl of 47%. The MSU-y lath-12 and scaffold-12 catalyst led to biphenyl yields as high as 49% and 84%, respectively. 116 Table 4.1. Results of HDS reactions over MSU-y catalysts exhibiting distinctly different morphologies. Time DBT Biphenyl Biphenyl Samplea (hr) Conversionb Selectivityc Yield % % % 1 53 43 23 Lath -6 2 79 62 49 3 69 62 43 W W 1 WWWéz’W "WS‘t—BWW "“3—6WW Lath -9 2 66 58 38 3 64 60 38 W W ' W W1“ W W72W 63 ”W45W’" Lath-12 2 78 61 48 3 77 64 49 W 1 " W 59”” WW52WWW W " W 31 ‘ Scaffold-6 2 76 73 55 3 74 89 66 WW 1 ' A W 61 W _ _ W 94 W W 57W— Scaffold-9 2 67 89 60 3 67 93 62 WW W W W1" 73 99 W W72W‘ Scaffold-1 2 2 87 93 81 3 86 98 84 ‘ The number refers to the loading of Mo, expressed as a mass percentage of material b percentage of initial DBT converted ° percentage of biphenyl in product 117 DBT 3P \ W“ _ 02:1.) —- /_\ \ / I (IQ-Gk) Figure 4.2. Reaction pathways of DBT in HDS reactions. The upper pathway involves direct desulfurization leading to the formation of biphenyl (BP). The lower pathway involves a hydrogenation step prior to desulfurization, leading to the formation of cyclohexylbenzene (CHB). 118 4.3.2. Structural and Textural Properties of MSU-y Catalysts Structural and textural properties were measured at several stages throughout the catalyst preparation and testing processes. Powder XRD patterns and nitrogen adsorption - desorption isotherms were obtained for each catalyst after each of the following: alumina synthesis, metal loading, sulfiding, and catalytic testing. The catalysts at each of these stages are referred to as pristine, loaded, sulfided, and spent, respectively. The properties of a commercial catalyst, Crosfield 465, are included for comparison. The wide angle powder XRD patterns for the lath-12, scaffold-12, and Crosfield catalysts are provided in Figures 4.3 — 4.5, respectively. The structural integrity of the MSU-y aluminas remained intact over the duration of the processing and HDS reaction; the y—Al203 phase is evident in each pattern. The Crosfield 465 catalyst shows different behavior. After sulfiding, the alumina peaks become increasingly broad, possibility due to the formation of mixed phases, or more likely, due to coating of the alumina surface by highly dispersed M082, leading to a decrease in the y-Al203 peak intensities. Restacking of the M082 may occur in the spent catalyst, exposing the y—Al203 support. The pattern of the spent catalyst shows peaks that are assignable to both the y-A1203 and poorly crystalline M082 phases. The absence of reflections consistent with M082 in any of the MSU-y catalysts’ diffraction patterns indicates that the active catalytic phase is present in small domains and is well dispersed on the supports. Similar behavior (not shown) is observed in the lower metal content samples as well. 119 Lath Spent Sulfided intensi Pristine IIIIIIIIIIIIIIIIIIIITIIIIIIIIIIIIlI—l 10 20 30 40 50 60 70 80 Degrees (26) Figure 4.3. Wide angle XRD patterns for lath-12 MSU-y alumina taken at various stages of the catalytic testing process. The y-Al203 phase is evident throughout the process and the lack of peaks assignable to M082 indicate that the active catalyst is present in small domains and is well dispersed on the support. 120 Scaffold Sulfided intensi Pristine IIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIITI 10 20 30 40 50 60 70 30 Degrees (26) Figure 4.4. Wide angle XRD patterns for scaffold-12 MSU-y aluminas taken at various stages of the catalytic testing process. The y-Al203 phase is evident throughout the process and the lack of peaks assignable to M082 indicate that the active catalyst is present in small domains and is well dispersed on the suppon. 121 Crosfield Spent Sulfided intensi Loaded IIIIllIIIIIIII'IIIIIIWIIIIIIIIITITI 10 20 30 40 50 60 70 30 Degrees (26) Figure 4.5. Wide angle XRD patterns for the commercial Crosfield catalyst, taken at various stages of the catalytic testing process. The decrease in intensity of the peaks assignable to y-Al203 in the pattern of the sulfided sample may be due to the coating of the alumina surface by small domains of M082. Peaks from y-A1203 and M082 in the pattern of the spent sample indicate that M032 particles may aggregate sufficiently to cause scattering and expose the surface of the support. 122 Upon inspection of the nitrogen adsorption — desorption isotherms of the 12% Mo loaded samples, Figures 4.6 and 4.7, it is clear that the metal loading, sulfiding, and catalytic testing have a deleterious effect upon the pore structure and textual properties of these catalysts. As the active catalytic phase is introduced onto the support, the pore size, pore volume, and surface area will naturally decrease. As shown in Table 4.2, in the lath-12 series, the pore volume is reduced to 25% of its original value after use in the HDS reaction. The scaffold-12 catalysts are the most resistant to the conditions imposed by this process as they retain approximately 45% of their original pore volumes. The effect is more pronounced at higher metal loading in each sample. A similar trend is observed in the surface areas of these materials. Again, the scaffold-12 alumina exhibits the highest degree of structural integrity, retaining approximately 65% of its original surface area while the lath-12 retains only 15%. in the case of the lath aluminas, at any point after loading the metal, the pores appear to collapse, leaving only a small amount of porosity. The pore size distributions are so broad that it is not possible to determine a mean pore size for the catalyst at any point after loading (inset, Figure 4.6). in the scaffold-12 catalysts, the pore size distribution shifts to a smaller average pore size throughout the process (inset, Figure 4.7). Comparing the MSU-y catalysts to the commercial catalyst, the scaffold-12 catalyst exhibits adsorption behavior most similar to the Crosfield catalyst (Figure 4.8). The textural properties of the lath, scaffold, and commercial catalyst are summarized in Table 4.2. 123 800: Lath 700: a . E ..>, L A 600j X; 9 : 1m 8 . ii 10 20 30 40 50 v _ Pore Diameter (nm) 500 '0 .. 0 . e . " a e I 8 400: Pristine U -l < . 0 -I E 3001 3 . Loaded '5 t > .. 200: 1 Sulfided 100: Spent / 5.......,............ 0 0.2 0.4 0.6 0.8 1 PIPO Figure 4.6. Nitrogen adsorption - desorption isotherms and BJH adsorption pore size distributions (inset) for lath-12 MSU-y catalysts taken at various stages of the catalytic testing process. The pore volume steadily decreases as the catalyst is loaded, sulfided, and spent. lsothems are offset 100 cc/g for clarity and PSD curves are displayed in the same order as the corresponding isotherms. 124 Scaffold W . 0 10 20 30 40 50 Pore Diameter (nm) Pristine Loaded Volume Adsorbed (cc/g) 200; Sulfided Spent PIPO Figure 4.7. Nitrogen adsorption — desorption isotherms and BJH adsorption pore size distributions (inset) for scaffold-12 MSU-y catalysts taken at various stages of the catalytic testing process. The pore volume decreases slightly as the catalyst is loaded, sulfided, and spent while the pore size is increased and the distribution broadened over the same conditions. Isothems are offset 0, 100, 100, and 200 cc/g from bottom to top for clarity and PSD curves are displayed in the same order as the corresponding isotherms. 125 - Crosfield Spent dVIdD 1 Sulfided b O O l 0 10 20 30 40 50 Pore Diameter (nm) 1 : Loaded Sulfided Volume Adsorbed (cc/g) 0.1 O O PIP Figure 4.8. Nitrogen adsorption - desorption isotherms and BJH adsorption pore size distributions (inset) for the commercial Crosfield catalyst taken at various stages of the catalytic testing process. The pore volume decreases slightly as the catalyst is sulfided and spent. However, the pore size and distribution remain relatively unaffected over the same conditions. lsothems are offset 100 cc/g for clarity and PSD curves are displayed in the same order as the corresponding isotherms. 126 Table 4.2. Textural progerties of MSUj alumina HDS catalysts. BET Surface BJH Pore _ Pore Volume Sample Condition Area Size — Ads. (cc/9) 2 (m lg) (nm) Pristine 0.60 309 1 0.0 Loaded 0.56 234 -- Lam's Sulfided 0.35 158 -- 2- _ 2 - Spent __. -928” 140 _ __ Pristine 0.60 309 1 0.0 Loaded 0.39 21 3 -- Lath‘g Sulfided 0.35 159 -- a - z -,_S_P_9m ___ __Q.2_-L_ 2 75 _ - _.:: __ Pristine 0 60 309 10.0 Loaded 0.33 1 82 -- Lath” Sulfided 0.31 127 -- g 2_ 2 Spent _ _Q.15 _ 47 -- Pristine 1 .23 368 13.0 Loaded 0.97 354 1 1.0 scafi°'d'6 Sulfided 0.77 270 1 1.0 2 v j W - (VS ent 22-22-. ____0.51_ _ _ 223 8.0 Pristine 1 .23 368 13.0 Loaded 0.58 335 4.0 scam'd'g Sulfided 0.43 271 4.0 _ SpenL ___ _ _ _0.44 - 249 4.0 Pristine 1 .23 368 13.0 Loaded 0.98 350 12.0 scafi°'d'12 Sulfided 0.58 261 9.0 ‘ Spent _ 0.56 238 9.0 Loaded 0.55 223 7.5 Crosfield Sulfided 0.24 100 1 1 .0 Spent 0.31 147 8.0 127 it is understandable that the scaffold and commercial catalyst have similar adsorption behavior as they exhibit similarities in morphology as well. Figures 4.9 through 4.11 provide a series of TEM images comparing the particle morphology of the loaded catalysts with that of the spent catalysts. The morphology of the scaffold-12 and commercial catalysts is largely unaffected after use in the HDS reaction. However, the long range order in the lath-12 catalyst has been destroyed to some extent. The insets in each case show an approximately 30 nm by 30 nm area. It can be seen that the local pore structure in each catalyst remains nearly unchanged. The decrease in pore volume, as determined from the isotherms, is reflected in these images. 4.3.3. Relationships Between Catalyst Structure and Performance While both MSUJY catalysts were effective in HDS reactions, the scaffold aluminas do show better performance than the lath aluminas. This can be explained by the differences in morphology of the alumina support. The pores in the lath aluminas are slit-like and run in the direction of the fundamental particles comprising the framework. Contrast that to the scaffold aluminas, where the pores are void spaces between randomly oriented fundamental particles. This leads to two different M032 particle sizes when supported on the aluminas. The layered structure of M082 is highly anisotropic. The sheets of M082 can extend a thousand-fold the distance in the stack direction, leading to the formation of a large, but catalytically inactive, basal plane. it has been realized that only the edge planes of the material contain the catalytically active sites. 128 Figure 4.9. Bright field TEM images of the scaffold-12 MSU-y catalyst (A) after loading and (B) after using in an HDS reaction for 3 hr at 400 °C and 400 psi. The scale bar represents 100 nm. The inset is an approximately 30 nm x 30 nm area of the catalyst. That the framework of the alumina is largely unaffected by the conditions of the HDS reaction provides evidence of its stability. 129 130 Figure 4.10. Bright field TEM images of the lath-12 MSU-y catalyst (A) after loading and (B) after using in an HDS reaction for 3 hr at 400 °C and 400 psi. The scale bar represents 100 nm. The inset is an approximately 30 nm x 30 nm area of the catalyst. The local pore structure (inset) is compromised to some degree, as is shown by the decreased void space in (B) as compared with (A). The long range order of the alumina laths is lost after use in the HDS reaction. 131 Figure 4.11. Bright field TEM images of the commercial Crosfield catalyst (A) after loading and (B) after using in an HDS reaction for 3 hr at 400 °C and 400 psi. The scale bar represents 100 nm. The inset is an approximately 30 nm x 30 nm area of the catalyst. Although the local pore structure (inset) appears to withstand the conditions encountered in the HDS reaction, it appears that there is some aggregation of the fundamental particles after use. 133 While the basal plane contributes greatly to the surface area of the catalyst, the surface area does not correlate with activity in HDS”. The edge planes contain the sulfur vacancies that are the active sites in HDS. Comparing the structures, there is a good match in morphology between the pore structure in the MSU-y lath catalyst and M082. Still, the pores will limit the extent of the growth of M082 in the direction perpendicular to the planes, leading to the formation of truncated slabs of the active sulfide on the MSU-y laths. This leads to highly dispersed, small domains of MoSz. The pore structure of the MSUJY scaffolds is not as ordered as in the MSU-y laths. Additionally, the fibers of alumina constituting the framework do not provide a good area for M082 to grow into large sheets. Consequently, the growth of M082 should be greater in the layer stacking direction on MSU-y scaffolds than on MSU-y laths. For the same molybdenum loading, a more highly stacked M082 particle will provide a greater number of edge sites than will a lesser stacked particle. This is one contributing factor to the greater conversion on MSU-y scaffolds than on MSU-y laths. Secondly, due to the nature of the open framework in the scaffold catalyst, access to these catalytically active edge sites may be increased as well. While the conversions on MSU-y catalysts are comparable, the selectivities are much better for the scaffolds than for the laths. The differences in selectivity are addressed by the rim — edge model of the M032 catalys123'24. If the M082 particle is viewed as a stack of discs, then the edge sites are divided into two categories, distinguished by their location within the stack (Figure 4.12). Sites on 135 :I edge It layers basal Figure 4.12. The Daage-Chianelli rim-edge model of the M082 catalyst particle. Sulfur removal takes place on rim and edge sites while hydrogenation of the aromatic ring is restricted to rim sites only“. 136 the top and bottom discs are designated rim sites and all others as edge sites. The taller the stack, the higher the edge / rim ratio. In the development of this model, Daage and Chianelli found that hydrogenation rates were inversely proportional to the stack height. It was therefore concluded that hydrogenation takes place only on the rim sites where sulfur removal (HDS) occurs on both sites. On MSU-y catalysts, a greater stack height on the scaffold catalysts is in accord with the increased selectivity. Having shorter but wider particles of M032 on the lath catalysts also agrees with the model. While it is argued above that merely increasing the stack height will lead to a higher conversion due to an increased number of edge sites, it is clear that particle size effects have a greater influence on selectivity than conversion. As an example, consider the blocks in Figure 4.13 to represent M082 particles. Keeping the overall amount of M032 the same, but chopping the slab in half and stacking the two pieces on top of each other, a larger number of edge sites are created. The more pronounced effect is that the ratio of rim sites to rim plus edge sites is decreased. This is in line with the data; selectivities are much better on scaffolds than on laths, while conversions are comparable at a given loading. 137 \3 fix. 100 Total Edge Sites: 660 Rim / (Rim + Edge) = 0.66 10 j 6 50 Total Edge Sites: 720 Rim / (Rim + Edge) = 0.33 Figure 4.13. influence of morphology on the ratio of rim sites to total sites available on the edges of MoS2 particles. While increasing the stack height leads to a slight increase in the overall number of edge sites, the more pronounced effect is the decrease in the relative number of rim sites. 138 4.4. Conclusions MSU-y aluminas were used effectively as catalyst supports for HDS catalysis. These crystalline mesostructured aluminas are capable of dispersing up to a 12% loading of molybdenum and able to withstand the harsh conditions encountered during the reaction. MSU-y aluminas with lath morphology had conversions are selectivities comparable to a commercial catalyst while the aluminas with scaffold morphology had conversions and selectivities superior to the commercial catalyst. The MSU-y alumina supports may lead to more economical catalysts, as greater conversions and selectivities are obtained with lesser quantity of active catalyst. The difference in performance of the two MSU-y aluminas is attributed to morphological matching with the catalytically active MoS2 phase. The laths promote growth of MoS2 in a sheet-like morphology while scaffolds lead to increased stacking in a direction perpendicular to the layers of MoS2. Additionally, the open framework and larger pore volume of the scaffold alumina as compared to the lath alumina may provide easier access to catalytically active edge sites of M082. Viewed in the rim-edge model of MoS2 catalysts, these results are reasonable. 4.5. Future Directions This initial venture into utilization of MSUJy aluminas in HDS reactions was fruitful, but there are some aspects that can be explored further. Can MSU-y aluminas be loaded with an increased mass of molybdenum? Some industrial 139 catalysts contain upwards of 20% molybdenum. The method of incorporation employed in this study was incipient wetness impregnation. Supports are loaded based upon their pore volume. The limiting factor is the solubility of the metal salt in the volume available with which to work. The higher pore volume of the scaffold materials may be better able to accommodate a higher loading. Othentvise, an alternative method will be needed. HDS reactions were limited to 3 hour runs. What is the longevity of the MSU-y catalyst? Can it remain on stream for extended periods of time? The average life of industrial catalysts may be approaching twp years. Are MSU-y catalysts poisoned after a certain time? If so, can they be recycled or regenerated? How do MSU-y catalysts perform with other substrates? As oil crudes become increasingly dirtier, larger molecules with sterically hindered sulfur atoms, such as 4,6-dimethyldibenzothiophene become more prevalent. The larger molecules present the toughest challenge in HDS. How do MSU-y aluminas perform with real oil feeds? There are reports in the literature employing lanthanide doped aluminas as supports for catalysts”. Noting that similar doping of MSU-y aluminas actually decreased the hydrothermal stability of the materials, it is possible that such an experiment would be of little value here. Finally, there has been a report of a new HDS catalyst claiming to have superior activity as compared to any catalyst currently in existence. This new catalyst, marketed under the name Nebula, was developed jointly by Akzo-Nobel, ExxonMobil, and Nippon Ketjen“. Nebula catalysts are not based upon an 140 alumina support as many of it predecessors are. These catalysts are already being used in European refineries to produce low sulfur diesel fuels. Consequently, there may be a shift in direction in research and development of HDS catalysts. What fate that holds for alumina based catalysts is yet to be determined. 141 4.6. References (1 I (2) (3) (4) (5) (6) (7) (8) (9) (10) (11) (12) (13) (14) (15) (16) (17) (18) in New York Times: New York, NY, 2000. Chianelli, R. R. NATO ASI Ser., Ser. C 1983, 105, 361-378. Topsoe, H.; Clausen, B. S. Catal. Rev. - Sci. Eng. 1984, 26, 395-420. Chianelli, R. R. ACS Symp. Ser. 1985, 279, 221-234. Prins, R.; De Beer, V. H. J.; Somorjai, G. A. Catal. Rev. - Sci. Eng. 1989, 31, 1-41. Harris, 8.; Chianelli, R. R. Theor. Aspects Heterog. Catal. 1990, 206-255. Chianelli, R. R.; Daage, M.; Ledoux, M. J. Adv. Catal. 1994, 40, 177-232. Hensen, E. J. M.; De Beer, V. H. J.; Van Santen, R. A. NATO ASI Ser., Ser. 31998, 60, 169-188. Nielsen, L. P.; Christensen, 8. V.; Topsoe, H.; Clausen, B. S. Catalysis Letters 2000, 67, 81 -85. Hensen, E. J. M.; Brans, H. J. A.; Lardinois, G. M. H. J.; De Beer, V. H. J.; Van Veen, J. A. R.; Van Santen, R. A. J. Catal. 2000, 192, 98-107. Farag, H.; Whitehurst, D. D.; Sakanishi, K.; Mochida, i. Catal. Today1999, 50, 9-17. Ledoux, M. J.; Michaux, 0.; Agostini, G.; Panissod, P. J. Catal. 1986, 102, 275-288. Carrado, K. A.; Xu, L.; Marshall, C. L.; Wei, 0.; Seifert, S.; Bloomquist, C. A. A. Stud. Surf. Sci. Catal. 2000, 129, 417-424. Carrado, K. A.; Marshall, C. L.; Brenner, J. R.; Song, K. Microporous Mesoporous Mater. 1998, 20, 17-26. Yue, Y.; Sun, Y.; Gao, Z. Catalysis Letters 1997, 47, 167-171. Song, 0.; Reddy, K. M. Appl. Catal., A1999, 176, 1-10. Ramirez, J.; Contreras, R.; Castillo, P.; Klimova, T.; Zarate, R.; Luna, R. Appl. Catal., A 2000, 197, 69-78. Okamoto, Y. Catalysis Today 1997, 39, 45-59. 142 (19) (20) (21) (22) (23) (24) Zhang, Z.; Hicks, R. W.; Pauly, T. R.; Pinnavaia, T. J. Journal of the American Chemical Society 2002, 124, 1592-1593. Tauster, S. J.; Pecoraro, T. A.; Chianelli, R. R. J. Catal. 1980, 63, 515- 519. Daage, M.; Chianelli, R. R.; Ruppert, A. F. Stud. Surf. Sci. Catal. 1993, 75, 571-584. Daage, M.; Chianelli, R. R. J. Catal. 1994, 149, 414-427. Miciukiewicz, J.; Laniecki, M.; Domka, F. Catalysis Letters 1998, 51, 65- 68. lnternet publication currently available on the ExxonMobil website: http://www.exxonmobiI.com/refiningtechnologies/pdf/h7_nebula.pdf 143 Chapter 5 Concluding Remarks A Commentary As mentioned in the introduction, there has been little progress made in the development of mesostructured aluminas relative to other metal oxides. At the start of this work, all known mesostructured aluminas were amorphous in nature. This work has demonstrated a considerable achievement in reporting at least two new crystalline forms of mesostructured alumina. Alumina has uses in numerous applications as a catalyst support or as a component in catalyst system. HDS was just one arena in which the utility of these new materials could have been demonstrated. The improvement in textural properties, along with the stability of these materials, provides great potential to MSUJY aluminas for use in a variety of catalytic applications. To fully reach this potential, eventually the syntheses reported here will need to be commercialized. That leads to one important drawback regarding these materials. They are made from relatively expensive precursors of polyalkylene oxides and aluminum alkoxide. Will other applications prove it necessary to scale up this production of this material? Are the increased costs balanced by increased performance? These are questions that will be addressed at some time in the future. On a personal note, this project has been very challenging. There were times that l wondered if anything would come out of it. Of course, in the end it has been 144 very rewarding. I am satisfied with the contributions that i have made to build upon existing knowledge. It has only been ten years since Mobil’s initial disclosure of the M418 family of materials. So while the entire field of mesostructured materials is relatively new, the subset of mesostructured aluminas is in infancy. I’m excited that I have had the opportunity to create some of the pioneering work in this field. MSU and the Pinnavaia group have been at the forefront of this from the beginning. The first mesostructured alumina and now the first crystalline mesostructured aluminas have developed in these labs. Finally, I am proud to be in the lineage of scientists that have contributed to this project. It started with Steve Bagshaw, continued with Peter Tanev, Wenzhong Zhang, myself, and is now left in the capable hands of Zhaorong Zhang (and future additions to the Pinnavaia group). There is still progress to be made in this field and i look fonrvard to the future contributions from this group. 145 IIItI'IIjflIjIIIII'IIjIIIIjIIIjIII 3