1:. y . .... . ..¢.€v:..;.. .2... ,2 :1. 3K . , V y u.... , , ‘ .11.... wgwuué‘rw "‘3th 5.... . a ‘1'. nah-nu“ ‘11:.0 “bikini ‘ .3:;.!..: £452.11... .. .4 oq Lmrsty This is to certify that the dissertation entitled REGIO AND CHEMO SELECTIVE HYDROGENATION BY PALLADIUM NANOPARTICLES EMBEDDED IN POLYELECTROLYTE FILMS presented by Somnath Bhattacharjee has been accepted towards fulfillment of the requirements for the Ph.D. degree in Chemistry 10/20/03 Date MSU is an Affirmative Action/Equal Opportunity Institution 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 REGIO AND CHEMO SELECTIVE HYDROGENATION BY PALLADIUM NANOPARTICLES EMBEDDED IN POLYELECTROLYTE FILMS By Somnath Bhattachaijee A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Chemistry 2008 ABSTRACT REGIO AND CHEMO SELECTIVE HYDROGENATION BY PALLADIUM NANOPARTICLES EMBEDDED IN THE POLYELECTROLYTE FILMS By Somnath Bhattacharjee Transition-metal nanoparticles (NPs) are attractive catalytic materials because they possess a high surface area to volume ratio and electronic properties that are a function of particle size. Without a suitable support, however, aggregation of NPs tends to reduce their surface area and alter their electronic properties. This research examines the catalytic properties of NPs embedded in multilayer polyelectrolyte films. The polyelectrolyte matrix restricts NP aggregation, and layer-by-layer assembly on alumina supports yields a heterogeneous catalyst that can be easily recycled. Encapsulated NPs exhibit remarkable catalytic selectivities that are a strong function of NP diameter. Pd NPs embedded in multilayer polyelectrolyte films can be easily prepared through layer-by—layer adsorption of poly(acrylic acid) (PAA) and poly(ethylenimine)- Pd2+ (PEI-Pd(II)) complexes followed by chemical reduction of Pd(II) with NaBH4. Transmission electron microscopy confirms the formation of Pd particles with diameters of 1-3 nm. Remarkably, [PAA/PEI-Pd(0)]3PAA films catalyze the hydrogenation of monosubstituted alkenes with turnover frequencies (TOFs) that are as much as IOO-fold higher than TOFs for hydrogenation of multiply substituted double bonds. These selectivities are higher than those of Wilkinson’s catalyst. Regioselective hydrogenation of monosubstituted C=C bonds occurs when substrate molecules contain both mono and multiply substituted C=C bonds. The combination of the encapsulating polyelectrolyte film and small N'Ps likely hinders access of multiply substituted double bonds to catalytic sites. NP-containing films can also be prepared through deposition of PAA-Pd(II)/PEI films and subsequent reduction of Pd(II). The use of different ratios of PAA to Pd(II) in deposition solutions gives a series of films with Pd NPs whose average diameters range from 2.2 to 3.4 nm, and the catalytic selectivities of these NPs vary dramatically with their size. TOFs for the hydrogenation of monosubstituted unsaturated alcohols increase with decreasing average NP size, whereas multisubstituted unsaturated alcohols show the opposite trend. Hence, the ratio of TOFs for the hydrogenation of allyl alcohol and crotyl alcohol is 39 with average particle diameters of 2.2 nm and only 1.3 with average particle diameters of 3.4 nm. Ratios of TOFs for hydrogenation of allyl alcohol and B-methallyl alcohol are as high as 240. Increasing selectivity with decreasing average particle size occurs with both films deposited on alumina powder and NPs stabilized by polyelectrolytes in solution. Presumably, the highest selectivities occur on the smallest NPs because hydrogenations of monosubstituted and multisubstituted double bonds take place on different types of atoms [PAA-Pd(0)/PEI]3 films also catalyze chemoselective hydrogenation of aromatic nitro compounds that contain other readily reducible functional groups. The resulting amines may serve as intermediates in the synthesis of a wide range of important organic compounds. In the reduction of 4—nitrobenzaldehyde to 4-aminobenzaldehyde, [PAA- Pd(O)/PEI]3 films are more selective than either Pd/C or Pd/alumina. Chemoselective hydrogenation of aromatic nitroamides, nitroesters, nitroaldehydes, and nitronitriles is also possible with these catalysts. TO MY FAMILY iv ACKNOWLEDGEMENT I wish to express my gratitude with humility to some special people who contributed enormously to my success as a graduate student. First I want to thank my thesis advisor Dr. Merlin Bruening for his encouragement, enthusiastic words, guidance and invaluable suggestions during this fascinating study. I especially thank him very much for accepting me into his group during a tough time as a graduate student and allowing me to independently explore my ideas. I primarily owe my achievement in this dissertation to him. I wish I could be able to grasp everything that he tried to teach me. I am extremely grateful to Dr. Babak Borhan for not only accepting me to be part of his group but also his generous guidance and invaluable suggestions during my stay in his group as well as the time after I lefi his group. Moreover, I convey my thanks to him for serving on my PhD committee. I express my special gratitude to him for providing me an opportunity to work with Dr. Jennifer Schomaker to learn several great tricks regarding organic synthesis. I am very thankful to Dr. Schomaker for her suggestions, guidance and enthusiastic words. I express my gratitude to her especially for sharing her experience regarding chemistry and life during my stay in Dr. Borhans’s group. I thank Dr. Gregory Baker and Dr. Kevin Walker for being on my PhD committee. I am grateful to Dr. Baker in particular for his motivating words and invaluable suggestions whenever we talked. I am thankful to all past and present colleagues whom I learn a lot from about chemistry and life. I would like to thank David Dotzauer, Sebastian Grajales, Dr. Srividhya Kidambi, Maneesha Adusumill, Lu Ouyang, Fei Xu, Weihan Wang, Yujing Tan, Guanqing Lin, Dr. Jamie Dunn, Dr. Lei Sun, Dr. Malaisamy Ramamoorthy, Dr. Jinhua Dai, Parul Jain, Randy Benedict as well as Dr. Chrysoula Vasileiou, Dr. Courtney Olmsted, Dr. Tao Zheng, Dr. Marina Tanasova, Dr. Montserrat Rabago Smith, Dr. Jun Yan, Zhihua Shang, Stewart Hart, Daniel Whitehead, Xiaoyong Li, Abra, Kate. Dr. Rachael Crist, Dr. Ben Travis, Dr. Qifei Yang. I especially would like to thank Dr. Srividhya Kidambi in particular for generously sharing her experience and knowledge of her research project with me. I would like to convey my special thanks to David Dotzauer for being an awesome colleague as well as a great fiiend and provide me all possible help and sharing his word of wisdom during my stay as a Dr. Bruening’s group member. I would also like to thank Sebastian Grajales very much for his help and wonderful discussion on several issues including chemistry. Last, but not least, I owe the most gratitude to my family members for their constant support, patience and encouragement. I express my special gratitude with humility to bapi, ma, baramama, puspita, didi, chordibhai, goutamda, subhanduda, didibhai, baba and ma (in-laws), mesomasai (Dr. Shyama Prasad Bhattacharyya) and many others for always being there for me. Finally, I gratefully acknowledge the generous financial support from Department of Chemistry, Michigan State University, US Department of Energy, American Chemical Society Petroleum Research Funding. vi TABLE OF CONTENTS LIST OF TABLES ................................................................................. X LIST OF FIGURES .............................................................................. XII LIST OF SCHEMES ............................................................................. XVI LIST OF ACRONYMS ....................................................................... XVIII CHAPTER 1: BACK GROUND INFORMATION ........................................... 1 1.1. Introduction .................................................................................... 1 . . Early Development of Catalytically Active Transition-Metal NPs. . . . . . . . . . . . . . ......4 1.3. Recent Developments in the Preparation of Stable, Catalytic Metal NPs. . . . . .......6 1.3.1. Immobilization of Metal NPs on Solid Supports ................................ 7 1.3.2. Solution-based Metal NP Stabilizers ............................................ 10 1.4. Layer-by-Layer Assembly of Oppositely Charged Polyelectrolytes.. .................. 15 1.5. Motivation and Research Goal ............................................................ 18 1.6. Outline of the Thesis ....................................................................... 20 1.7. References ................................................................................... 23 CHAPTER 2: SELECTIVE HYDROGENATION OF MONOSUBSTITUTED ALKENES BY Pd NANOPARTICLES EMBEDDED IN POLYELECTROLYTE FILMS ................................................................................................ 40 2.1. Introduction .................................................................................. 40 2.2. Experimental ................................................................................. 42 2.2.1. Materials ............................................................................ 42 2.2.2. Preparation of Pd NPs Encapsulated in Polyelectrolyte Multilayer Films ......................................................................................... 42 2.2.3. TEM of Pd NPs .................................................................... 43 2.2.4. Determination of the Amount of Pd in Catalyst and Deposition Solutions .................................................................................... 43 2.2.5. Hydrogenation Reaction .......................................................... 44 2.2.6. Recycling Experiment ............................................................ 45 2.2.7. Hydrogenation Reactions by Wilkinson’s Catalyst ........................... 45 2.2.8. Isolation of Products .............................................................. 46 vii 2.3. 2.4. 2.5. 2.6. 2.2.9. Transport Studies .................................................................. 47 Results and Discussions .................................................................... 48 2.3.1. NP Size .............................................................................. 48 2.3.2. Intermolecular Selectivity ........................................................... 50 2.3.3. Intramolecular Selectivity ........................................................ 54 Conclusion ................................................................................... 58 Appendix ..................................................................................... 59 References .................................................................................... 66 CHAPTER 3: SELECTIVITY AS A FUNCTION OF NANOPARTICLE SIZE IN THE CATALYTIC HYDROGENATION OF UNSATURATED ALCOHOLS......69 3.1. 3.2. 3.3. 3.4. 3.5. 3.6. Introduction .................................................................................. 69 Experimental Section ....................................................................... 7 1 3.2.1. Materials ............................................................................ 71 3.2.2. Preparation of Pd NPs Encapsulated in Polyelectrolytes .................... 71 3.2.2. 1. [PAA-Pd(0)/PEI]3 on Alumina ......................................... 71 3.2.2.2. [Pd(O)/PEI] on Alumina .................................................. 72 3.2.2.3. PAA-Pd(0)/PEI in Solution ........................................................... 72 3.2.3. Characterization of Pd NPs ....................................................... 73 3.2.4. Hydrogenation Reactions ............................................................ 76 Result and Discussion ...................................................................... 77 3.3.1. Catalyst Synthesis and Variation of Particles Size ............................ 77 3.3.2. Selective Hydrogenation with Films on Alumina Particles .................. 90 3.3.3. Hydrogenation with NP/Polyelectrolyte Complexes in Solution. . . . . .......98 3.3.4. Solvent Effects on Hydrogenation ................................................ 99 . 3.3.5. Hydrogenation of Other Unsaturated Alcohols .............................. 101 Possible Explanation of Selectivity ..................................................... 102 Conclusion .................................................................................. 105 References .................................................................................. 107 CHAPTER 4: CHEMOSELECTIVE NITROARENE HYDROGENATION CATALYZED BY PALLADIUM NPs EMBEDDED IN POLYELECTROLYTE MULTILAYERS .................................................................................. 110 4.1. Introduction ................................................................................. 110 4.2. Experimental ................................................................................ 114 4.2.1. Materials ........................................................................... 114 4.2.2. Preparation of Pd NPs on Alumina ............................................ 115 4.2.2.1. [PAA-Pd(0)/PEI]3 ....................................................... 115 4.2.2.2. Reduced PdC142‘ on alumina ........................................... 115 4.2.3. Characterization of Pd NPs ...................................................... 116 viii 4.2.4. Determination of the Pd Content of Catalysts Used in Hydrogenation .............................................................................. l 16 4.2.5. General Procedure for the Hydrogenation of Nitroaromatic Compounds ................................................................................. 116 4.3. Result&Discussion ....................................................................... 119 4.3.1. Catalyst Preparation and Characterization ...................................... 119 4.3.2. Hydrogenation of nitrobenzene as a function of solvent and nitrobenzene concentration. .............................................................................. 121 4.3.3. Hydrogenation of nitrobenzaldehyde .......................................... 123 4.3.4. Hydrogenation of Other Nitroarenes ............................................ 125 4.4. Conclusion .................................................................................. 127 4.5. Appendix ................................................................................... 128 4.6. References ................................................................................... 140 CHAPTER 5: CONCLUSIONS AND FUTURE WORK ................................ 144 5.1. Selective Catalysis ......................................................................... 144 5.2. Selective Hydrogenation of Unsaturated Alcohols as a Function of NP Diameter .................................................................................... 145 5.3. Chemoselective Hydrogenation of Aromatic N‘itro Compounds by Pd NPs encapsulated in polyelectrolyte films ................................................... 145 5.4. Future Work ................................................................................ 146 5.4.1. Effect of NP Size on the Chemoselective Hydrogenation of 4- nitrobenzaldehyde ......................................................................... 146 5.4.2. The Contribution of Size-Dependent Electronic Properties to Selective Hydrogenation .............................................................................. 146 5.4.3. Synergistic Effects Using Bimetallic NPs ..................................... 147 5.5. Summary Statement ....................................................................... 148 5.6. References .................................................................................. 149 ix LIST OF TABLES Table 2.1. TOFs for the hydrogenation of a series of unsaturated alcohols by different Pd catalysts ........................................................................................... 51 Table 2.2. TOFs for the hydrogenation of a series of unsaturated alcohols by Wilkinson’s catalyst ................................................................................. 54 Table 2.3. TOFs (moles hydrogenated/ mole Pd/ h) for selective hydrogenation of monosubstituted double bonds in the presence of di and tri substituted double bonds in the same molecules. The reaction was catalyzed by [PAA/PEI—Pd(0)]3PAA on alumina in a 3:1 mixture of water and ethanol ............................................................. 56 Table 2. A-l Time-dependent percentage of crotyl alcohol hydrogenation catalyzed by [PAA/PEI-Pd(0)]3PAA in water .................................................................. 60 Table 3.1. Amounts and Pd Content of Catalysts Used in Hydrogenation Reactions ............................................................................................. 75 Table 3.2. Calculation of the relative surface area of large (ranging from ~50-300 nm) particles and NPs (< 5 nm) in [PAA-Pd(0)/PEI]3 films and reduced [PdCl42'/PEI]3 films on carbon-coated TEM grids. The initial concentration of Pd(II) was 15 mM in the solution containing PAA-PdCl42' or PdCl42'. Large particles were not prevalent when using lower concentrations of Pd(II) .............................................................. 84 Table 3.3. TOFs and TOF ratios for unsaturated-alcohol hydrogenation catalyzed by [PAA-Pd(O)/PEI]3 films that contain Pd NPs with different diameters. The catalytic films were deposited on micron-sized alumina. Values in parentheses are the sum of TOFs for hydrogenation and isomerization .................................................................. 92 Table 3.4. TOFs and TOF ratios for unsaturated-alcohol hydrogenation catalyzed by [PAA-Pd(0)/PEI]3/(PAA/PEI)3 films that contain Pd NPs with different diameters. The films were deposited on micron-sized alumina. Values in parentheses are the sum of TOFs for hydrogenation and isomerization ...................................................... 93 Table 3 5. TOFs and TOF ratios for unsaturated-alcohol hydrogenation catalyzed by reduced [PdCl42'/PEI]3 films that contain Pd NPs with different diameters. The films were deposited on micron-sized alumina. Values in parentheses are the sum of TOFs for hydrogenation and isomerization .................................................................. 97 ' Table 3.6. TOFs and TOF ratios for unsaturated-alcohol hydrogenation catalyzed by a solution of PAA—Pd(O)/PEI that contains Pd NPs with different diameters. Values in parentheses are the sum of TOFs for hydrogenation and isomerization ..................... 99 Table 3.7. TOFs and TOF ratios for unsaturated-alcohol hydrogenation catalyzed by [PAA-Pd(O)/PEI]3 films in different solvents. The films contain Pd NPs with different diameters and are deposited on micron-sized alumina. Values in parentheses are the sum of TOFs for hydrogenation and isomerization ................................................ 100 Table 3.8. TOFs and TOF ratios for unsaturated-alcohol hydrogenation catalyzed by [PAA-Pd(0)/PEI]3 films that contain Pd NPs with different diameters. The catalytic films were deposited on micron-sized alumina. Values in parentheses are the sum of TOFs for hydrogenation and isomerization ................................................................ 102 Table.3.9. Formula used to calculate m, the number of shells for NPs having cuboctahedron geometry, and the number of Pd atoms at the various positions available . on NP surfaces ...................................................................... ' ............... 104 Table 4.1. TOFs for nitrobenzene hydrogenation catalyzed by alumina-supported [PAA-Pd(II)/PEI]3 films in different solvents ................................................. 122 Table 4.2 TOFs (moles hydrogenated/moles Pd/h) for nitrobenzene hydrogenation by different catalysts in EtOH ..................................................................... 122 Table 4.3. Product distributions from 4—nitrobenzaldehyde hydrogenation (>99% conversion) using several catalysts and two solvent systems ................................ 124 Table 4.4. Catalytic hydrogenation of various aromatic nitro compound by [PAA- Pd(O)/PEI]3 films and reduced [PdCluz'] on alumina .......................................... 126 xi LIST OF FIGURES Figure 1.1. Number of publications found in a search for the key word “nanoparticles” using SciFinder Scholar (Chemical Abstracts search tool) ...................................... 1 Figure 1.2. Number of publications found for the key phrase “Nanoparticles for catalysis” in a SciFinder Scholar (American Chemical Society Search tool) search ........ 6 Figure 1.3. Encapsulation of metal nanoparticles in a dendrimer matrix ................ 14 Figure 1.4. Formation of bilayer on the substrate through layer-by-layer deposition of polyelectrolytes ...................................................................................... 16 Figure 1.5. Intermolecularly selective hydrogenation of allyl alcohols by [PAA/PEI- Pd(0)]3PAA on alumina ............................................................................ 19 Figure 2.1. Home-built diffusion dialysis set-up for transport experiments ............ 47 Figure 2.2. TEM image of [PAA/PEI/Pd(0)]3PAA on carbon-coated cupper grid. . . ..48 Figure 2.3. TEM image of a [PAA/PEI-Pd(0)]3PAA film on a carbon-coated copper grid. This image was taken at a lower magnification than Figure 2.4 in to better show the distribution of NPs .................................................................................. 49 Figure 2.4. Histogram of Pd NP diameters in several TEM images. The average particle size is 2.2 nm ................................................................................ 49 Figure 2.5. Time evolution of the receiving phase concentrations of allylic alcohols during diffusion dialysis of 1 (squares), 2 (triangles), 3 (diamonds), and 4 (circles) through porous alumina membranes coated with a [PAA/PEI]7 film. The source phase initially contained 20 mM of the allylic alcohols, and the receiving phase was initially deionized water ...................................................................................... 58 Figure 2. A-l Example gas chromatograms of the reaction mixture from hydrogenation of crotyl alcohol (3) after 10 min (black line) and 40 min (red line) of reaction time. The reaction was catalyzed by [PAA/PEI-Pd(0)]3PAA in water. Integration was performed using instrument software .......................................... 61 Figure 2. A-2. Plot of time-dependent percentage of crotyl alcohol hydrogenation catalyzed by [PAA/PEI-Pd(0)]3PAA in water ................................................... 61 Figure 2. A—3 . 1H NMR spectrum of isolated crude 2-heptene-4-ol (compound 6) in CD3OD. Impurity peaks are indicated with a * ................................................ 62 xii Figure 2. A-4 '3 C NMR spectrum of isolated crude 2-heptene-4-ol (compound 6) in CD3OD. Impurity peaks are indicated with a * ................................................ 63 Figure 2. A-S. IH NMR spectrum of purified 3,7-dimethyl-6-octen—3-ol (compound 8) in CD3OD. Impurity peaks are indicated with a * .......................................... 64 Figure 2. A-6. '3 C NMR spectrum of purified 3,7-dimethyl-6-octen-3-ol (compound 8, chapter 2) in CD301). Impurity peaks are indicated with a * .............................. 65 Figure 3.1. (a) TEM image of a [PAA-Pd(0)/PEI]3 film on a carbon-coated copper grid. 1 mM Pd(II) was present in the PAA-KdeCl4 solution used for film deposition. A few of the NPs are circled. (b) Histogram of Pd NP diameters in several TEM images. The average particle size is 2.2:t0.6 nm. The area-average particle size is 2.5i0.7 nm...78 Figure 3.2. (a) TEM image of a [PAA-Pd(O)/PEI]3 film on a carbon-coated copper grid. 4 mM Pd(II) was present in the PAA-KdeCl4 solution used for film deposition. A few NPs are circled. (b) Histogram of Pd NP diameters in several TEM images. The average particle size is 3.0i0.8 nm. The area-average particle size is 3.1d:0.5 nm. . .......79 Figure 3.3. (a) TEM image of a [PAA-Pd(O)/PEI]3 film on a carbon-coated copper grid. 8 mM Pd(II) was present in the PAA-KdeCl4 solution used for film deposition. (b) Histogram of Pd NP diameters in several TEM images. The average particle size is 3.2:t0.6 nm. The area-average particle size is 3.5:t0.7 nm .................................... 80 Figure 3.4. (a) TEM image of a [PAA-Pd(0)/PEI]3 film on a carbon-coated copper grid. 15 mM Pd(II) was present in the PAA-KdeCl4 solution used for film deposition. (b) Histogram of Pd NP diameters in several TEM images. The average particle size is 3.4i1.3 nm. The area-average particle size is 3.8i1.1 nm .................................... 81 Figure 3.5. Low-magnification TEM images of large particles and NPs in [PAA- Pd(0)/PEI]3 films on carbon-coated copper grids. The films were prepared with 15 mM Pd(II) in the PAA deposition solution. (a) Image showing large particles with diameters of 160d:100 nm, as determined from several TEM images. Such large particles appeared in ~30% of similar images. (b) Image of NPs. Larger spots are overlapping multiple particles. The dark spot in the center of low-magnification images is due to a defect in the lens where it was burned by the electron beam ............................................. 83 Figure 3.6. (a) TEM image of a dried PAA-Pd(0)/PEI solution on a carbon-coated copper grid. 1 mM Pd(II) was initially present in the PAA solution. (b) Histogram of Pd NP diameters in several TEM images. The average particle size is 2.6i0.8 nm. The area- average particle size is 3.1i0.8 nrn ............................................................... 86 Figure 3.7 . (a) TEM image of a [PAA-Pd(O)/PEI] solution on a carbon-coated copper grid. 4 mM Pd(II) was initially present in the PAA-PdCluz' solution. (b) Histogram of Pd xiii NP diameters in several TEM images. The average particle size is 3.2:t0.7 nm. The area- average particle size is 3.4:t0.7 nm ............................................................... 87 Figure 3.8. (a) TEM image of a dried PAA-Pd(0)/PEI solution on a carbon-coated copper grid. 8 mM Pd(II) was initially present in the PAA solution. (b) Histogram of Pd NP diameters in several TEM images. The average particle size is 3.5:t0.7 nm. The area- average particle size is 3.7i0.7 nm ............................................................... 88 Figure 3.9. (a) TEM image of a [PAA-Pd(O)/PEI] solution on a carbon-coated copper grid. 15 mM Pd(II) wasinitially present in the PAA-PdCl42‘ solution. (b) Histogram of Pd NP diameters in several TEM images. The average particle size is 5.1:t1.3 nm. The area-average particle size is 5.8:l:1.4 nm ......................................................... 89 Figure.3.10. (a) TEM image of a dried PAA-Pd(O)/PEI solution on a carbon-coated copper grid. 15 mM Pd(II) was initially present in the PAA solution. This image was taken at a relatively low magnification to show the distribution of NPs and the absence of large particles on the TEM grid. The dark spot in the center of low-magnification images is due to a defect in the lens where it was burned by the electron beam ..................... 90 Figure 3.11. (a) TEM image of a [Pd(0)/PEI]3 film on a carbon-coated copper grid. The concentration of PdCl42' used in depositing the precursor [PdCl42’/PEI]3 film was 1 mM. (b) Histogram of Pd NP diameters in several TEM images. The average particle size is 3.0:t0.9 nm. The area-average particle size is 3.5i1.0 nm .................................... 94 Figure 3.12. (a) TEM image of a [Pd(0)/PEI]3 film on a carbon-coated copper grid. The concentration of PdCl42' used in depositing the precursor [PdCl42'/PEI]3 films was 15 mM. (b) Histogram of Pd NP diameters in several TEM images. The average particle size is 3.4i0.8 nm. The area-average particle size is 3.8i0.9 nm ................................. 95 Figure 3.13. (a) Low-magnification TEM image of a reduced [PdCl42'/PEI]3 film on a carbon-coated copper grid. 15 mM PdCl42’ was initially present in the deposition solution. The average size of these large particles in several TEM images is 170i110 nm. Large particles are visible in about 60% of these low-magnification images, but we estimate that the surface area of the large particles constitutes ~33% of the surface area in the film (Table 3.2) ............................................................................................ 96 Figure 3.14. Plot of the calculated percentage of surface, defect and face atoms with respect to the total atoms for each particle size ................................................ 105 Figure 4.1. (a) TEM image of a [PAA-Pd(O)/PEI]3 film on a carbon-coated copper grid. 15 mM Pd(II) was present in the PAA-KdeCl4 solution used for film deposition ............................................................................................. 120 Figure 4.1. (b) Histogram of Pd NP diameters in several TEM image based on their area..... ............................................................................................................................ 121 xiv Figure 4. A-l. 1H NMR spectrum of purified 2-Aminotoluene (compound 7) in CDC13 ............................................................................................... 128 Figure 4. A-2. '3 C NMR spectrum of purified 2-Aminotoluene (compound 7) in CDC13 ............................................................................................................................... 129 Figure 4. A-3. 1H NMR spectrum of purified Methyl-4-aminobenzoate (compound 4) in CDC13 ......................................................................................... 130 Figure 4. A-4. '3 C NMR spectrum of purified Methyl-4-aminobenzoate (compound 4) in CDC13 ......................................................................................... 131 Figure 4. A-5. 1H NMR spectrum of purified 4-Aminobenzamide (compound 6) in CD3OD .............................................................................................. 132 Figure 4. A-6. '3 C NMR spectrum of purified 4-Aminobenzamide (compound 6) in CD30D .............................................................................................. 133 Figure 4. A-7. 1H NMR spectrum of purified 4-Aminobenzaldehyde and residual AcOEt (compound 2) in CDC13 .................................................................. 134 Figure 4. A-8. ’3 C NMR spectrum of purified 4-Aminobenzaldehyde and residual AcOEt (compound 2) in CDC13 .................................................................. 135 Figure 4. A-9. IH NMR spectrum of purified 4-Aminobenzonitrile (compound 3) in CDC13 ................................................................................................ 136 Figure 4. A-10. '3 C NMR spectrum of purified 4-Aminobenzonitrile (compound 3) in CDC13 ............................................................................................. 137 Figure 4 A-ll. 1H NMR spectrum of purified 4-Aminoacetophenone (compound 5) in CDCl3 ............................................................................................. 138 Figure 4. A-12. '3 C NMR spectrum of purified 4-Aminoacetophenone (compound 5) in CDC]; ......................................................................................... 139 XV LIST OF SCHEMES Scheme 1.1 Heck reaction of styrene and iodobenzene catalyzed by Pd/C and Pd NPs on single-wall carbon nanotubes ..................................................................... 8 Scheme 1.2. Oxidation of cinnamyl alcohol by Pd/C and Pd NPs on single-wall carbon nanotubes .............................................................................................. 8 Scheme 1.3. Hydrogenation of cinnamaldehyde by Pd/C and Pd NPs on single-wall carbon nanotubes ..................................................................................... 8 Scheme 1.4. Pauson-Khand reactions catalyzed by Co and Pd NPs supported on silica for a one-pot syntheses of bicyclic enones ......................................................... 9 Scheme 1.5. 3-component cascade process by palladium NPs generated in situ to form isoindolinone ......................................................................................... 1 l Scheme1.6. Regioselective silylation of sugars through Pd NP-catalyzed silane alcoholysis ............................................................................................ 1 1 Scheme 1.7. Selective hydrogenation of Dehydrolinalool to linalool by PS-b-P4VP-Pd on alumina ............................................................................................. 13 Scheme 1.8. Regio selective hydrogenation of Linalool to dihydrolinalool by [PAA/PEI-Pd(0)]3PAA films on alumina ........................................................ 20 Scheme 1.9. Selectivity as a function of NP size in the catalytic hydrogenation of unsaturated alcohols ................................................................................. 21 Scheme 2.]. [PAA/PEI-Pd(0)]3PAA film formation on Alumina .......................... 41 Scheme 3.]. Cuboctahedra Pd NP with different atoms .................................... 103 Scheme 4.1. Chemoselective hydrogenation of aromatic nitro compounds by a montrnorillonitesilylpalladium (II) complex ................................................... 111 Scheme 4.2. Preparation of a montmorillonitesilylpalladium (II) complex .............. 112 Scheme 4.3. Silicon hydride reduction of nitro benzaldehyde catalyzed by Pd NPs generated in-situ ............................................................................................................... 1 13 Scheme 4.4. Chemoselective hydrogenation of p-nitrobenzonitrile catalyzed by Pd NPs on a carbon nanofiber .............................................................................. 113 xvi Scheme 4.5. Hydrogenation of 4-nitrobenzaldehyde to several possible products. . 124 xvii LIST OF ACRONYMS LBL: Layer-by-layer PEI: Poly(ethylenimine) PAA: Poly(acrylic acid) TEM: Transmission Electron Microscopy AES: Atomic Emission Spectroscopy. NP: Nanoparticles DMF: Dimethylformamide MSI: Metal-Support Interaction. TOF: Turnover Frequencies PMHS: Polymethylhydrosiloxane EtOAc: Ethyl Acetate EtOH: Ethanol THF: Tetrahydrofuran xviii CHAPTER 1 BACK GROUND INFORMATION 1.]. Introduction During the last 15 years, the formation, characterization, and application of nanoparticles (N Ps) have become one of the most prominent areas of chemical research. The yearly number of publications in this field has grown exponentially as shown in Figure 1.1. Much of the interest in NPs stems fi'om the fact that they often manifest properties that are different fiom the properties of the bulk material of the same 30000 20000 -' 10000 ‘ # Hits for Nanoparticles 1990 1995 2000 2005 Publication Year Figure 1.1. Number of publications found in a search for the key word “nanoparticles” using SciFinder Scholar (Chemical Abstracts search tool). “Images in this dissertation are presented in color” composition,"2 and classical laws of physics are sometimes not applicable to small NPs. By definition, NPs have diameters that range from 1 to 100 nm, so they contain between about 10 and 106 atoms or molecules per particle.3’4 Because of their unique properties, NPs are being intensively studied for various applications in sensingf"l3 24-29 30-49 optoelectronicsy"18 tissue engineering,”"23 biotechnology, and catalysis. Because of their high surface area to volume ratio, perhaps the most natural application area for NPs is catalysis. Berzelius introduced the term catalysis in 1836 to classify certain types of chemical reactions whose progress is greatly affected by additional substances even though they are neither reactants nor products.50 Since then, a molecular understanding of catalytic processes has been emerging, and catalytic reactions constitute about 90% of chemical processes.“ Solid-state materials, including transition metals, are frequently the materials of choice for catalysis because they can be readily separated from a reaction mixture and frequently reused. One disadvantage of heterogeneous catalysts, however, is that typically only surface atoms serve as active sites for catalytic processes. The percentage of atoms that reside at the surface of a material increases with decreasing particle size, so small metal particles are highly reactive catalysts for many reactions. However, optimal use of metal atoms for catalysis requires extremely small particles. Even in a 20 nm-diameter particle, only 10% of the atoms reside at the particle surface. In contrast, in a 3 nm-diameter particle, 50% of the atoms are on the surface. The large surface areas of nanostructured metals allow efficient use of metal atoms in various catalytic reactions. During the twentieth century, chemists made tremendous achievements in heterogeneous catalysis, which is responsible for the vast majority of chemical - - 9 production.52 5 Nevertheless, heterogeneous catalysts sometimes show relatively low selectivity, and it is often difficult to elucidate catalytic mechanisms. On the other hand, homogeneous catalysts, which have been primarily developed since the early l970$,60'63 can sometimes be tailored to selectively catalyze specific reactions. Still, homogeneous catalysts are not often used on a large scale because they cannot be easily recovered. Transition-metal NPSM’65 might bridge the gap between homogeneous and heterogeneous catalysis because they can be either suspended in solution as homogeneous catalysts or fixed onto a support. Although by definition the diameters of NPs vary between one nm and 100 nm, the most active NPs in catalysis have diameters of only a few 11mm“58 NPs are also relevant to homogeneous catalysis because there is a full continuum between small metal clusters and larger NPs. Moreover, NP catalysts are frequently selective, efficient, and recyclable, and thus they meet the modern requirements for green catalysts. Already, numerous applications of transition-metal NP catalysts are emerging.”97 As discussed above, the high surface area of NPs is attractive for catalysis. However, the high fraction of active surface atoms also gives rise to a high surface energy that ofien causes metal NPs to aggregate. This aggregation yields larger, stable particles with less surface area and less catalytic activity and provides a serious challenge in catalysis where control over NP size is often vital for improving the reactivity and selectivity of a catalyst. Fortunately, several methods have been developed to stabilize metal NPs and prevent aggregation. Perhaps the most common of these is to cap the particles with ligands that reduce surface energy or provide repulsive interactions between particles.98"°l However, the capping ligands may isolate the particle surface and decrease catalytic activity. Other methods of NP stabilization that may be appropriate for catalysis include immobilization on solid supports and in polymer films.'02"03 This dissertation focuses on the layer-by-layer (LBL) assembly of polyanions and polycations to incorporate stable Pd NPs in a polyelectrolyte matrix. These NPs show remarkable regio and chemoselective hydrogenation, and selectivities vary dramatically with NP diameter. In this introduction I attempt to place my work in context by providing background on previous studies of NP catalysis and layer-by-layer deposition. Section 1.2 discusses the historical background of transition-metal NPs and their preliminary development for catalytic applications. Section 1.3 describes more recent work including stabilization of NPs using ligands, surfactants, dendrimers/polymers, and solid supports. Finally, the last section of this chapter (section 1.4) introduces the general procedure of layer-by-layer assembly, which can be implied to prepare stable and catalytically active homogeneous and heterogeneous metal NPs for catalytic applications. 1.2. Early Development of Catalytically Active Transition-Metal N Ps In the mid-nineteenth century, Michael Faraday was the first to record the synthesis of NPs.82"°4 Faraday noted the formation of deep red solutions upon the reduction of aqueous NaAuCl4 by phosphorous in C82. Faraday believed that the particles formed via his method had diameters smaller than wavelength of visible light. Unfortunately, at the time there were no tools for determining the actual size of the gold particles. In the 19805 J. M. Thomas reproduced the formation of gold particles according to F araday’s method and generated NPs with diameters ranging from 3-30 nm.105 However, these gold NPs were not used for catalytic applications. The use of transition-metal NPs as catalysts started in the second half of the twentieth century. Initial studies included a small but important group of reactions including hydrogenation, hydrosilylation and hydration of unsaturated organic substrates. In the 19405, Nord reported pioneering catalytic applications of NPs in the reduction of nitrobenze.'06"08 In 1970, Parravano used Au NP catalysts for hydrogen-atom transfer between benzene and cyclohexane, as well as oxygen-atom transfer between CO and C02.'09 However, the real breakthrough in NP catalysis came in the 19805 when Haruta observed that oxide-supported Au NPs catalyze CO oxidation by 02 at low temperatures and that the small size of gold is crucial for the tranforrnation.”0'”3 In the 19705 and 19805, NP catalysts were also used in the fields of redox catalysis, photocatalysis including photo-water splitting and photo-hydrogenation of C02 to methane, and hydrogenation and oxidation of unsaturated compoundsm'122 In the 19905, Reetz reported seminal studies of Pd NPs as catalysts for Heck coupling reactions between butyl acrylate and iodobenzene or aryl bromides and styrenem'I30 Moreover, Figure 1.2 illustrates that in the first few years of the twenty-first century, there has been enormous growth in the number of publications concerning the application of metal NPs for catalysis. The fundamental goals of recent research on NP catalysts include preparation of stable NPs; improvement of catalyst activities and selectivities; and understanding the catalytic mechanisms which cause high activity and selectivity. The next section discusses studies in these specific areas. # Hits for Nanoparticles for Catalysrs N O G O 1991 1996 2001 2006 Publication Year Figure 1.2. Number of publications found for the key phrase “Nanoparticles for catalysis” in a SciFinder Scholar (American Chemical Society Search tool) search. 1.3. Recent Developments in the Preparation of Stable, Catalytic Metal NPs As mentioned in section 1.1, the preparation of catalytically active, stable metal NPs is challenging due to their tendency to aggregate. To prevent aggregation, metal NPs are immobilized on various solid supports such as carbon and metal oxides, or stabilized by surfactants, ligands, or polymers. Immobilization on a support physically prevents NP aggregation, whereas ligands, polymers, and surfactants typically surround particles to lower surface energy or provide repulsive forces between particles. NPs stabilized by surfactants, ligands, and small polymers can serve as homogeneous catalysts, whereas NPs on solid supports act as heterogeneous catalysts that can be readily recovered. The following subsections focus on the different techniques for preparing stable metal NPs and the catalytic applications of these particles. 1.3.1. Immobilization of Metal N Ps on Solid Supports Carbon Supports Charcoal is a common solid support for catalysts, and Pd on activated C (Pd/C) is an archetypical catalyst. Activated carbon is produced by treatment of charcoal with acid and elution to remove ash and other contaminants, and to prepare a highly porous material. Purified activated carbon often provides optimal interactions with the precious metal catalysts and ensures dispersibility in the reaction media. A recent report suggests that Pd NPs on single wall carbon nanotubes are more active than commercially available Pd on activated carbon for the Heck reaction of styrene and iodobenzene (Scheme 1.1). However, Pd on activated C is the more active catalyst in the case of oxidation of cinnamyl alcohol to cinnamaldehyde by molecular oxygen and the hydrogenation of cinnamaldehyde to 3-phenylpropionaldehyde (Schemes 1.2 and 1.3).13' The average diameter of Pd NPs on single-wall carbon nanotubes is 5 nm, whereas the average diameter of Pd on activated carbon (1 wt %) is 14.5 nm. A number of studies show the high activity of metal NPs on carbon in a wide range of reactionsm"35 Another report shows that catalytic activities of cobalt NPs on graphite, mesoporous carbon, and charcoal are almost same for the Pauson-Khand reaction. This suggests that in this case, the catalytic activity depends on the NPs and not on the nature of support '36 | \ Tributylamine, DMF, 140°C 0 / O + s O Pd Catalyst Single-Wall Carbon Nanotube 60 min 100% conversion PdlC 60 min 65% conversion Scheme 1.1. Heck reaction of styrene and iodobenzene catalyzed by Pd/C and Pd NPs on single-wall carbon nanotubes. O ./\/\OH 02, EtOH, 80°C \ H Pd Catalyst Single-Wall Carbon Nanotube 10 h 2% yield PdIC 10 h 11% yield Scheme 1.2. Oxidation of cinnamyl alcohol by Pd/C and Pd NPs on single-wall carbon nanotubes. ' O O \ H H2, o-xylene, 80 C = O \ OH Pd Catalyst Single-Wall Carbon Nanotube 2 h 10% yield PdIC 2 h 100% yield Scheme 1.3. Hydrogenation of cinnamaldehyde by Pd/C and Pd NPs on single-wall carbon nanotubes. Metal Oxide Supports A large number of recent publications focus on catalysis with NPs supported on oxides of Si 137-152 A169.150.153.155 Ti I56-158 Zr 159-168 Ca 76 Mg 69,155,169-173 and Zn 174 These supports are attractive because they are stable at high temperatures and easy to remove from the reaction medium. The performances of these supported catalysts were examined for various reactions such as hydrogenation of unsaturated substrates,175 l24-I27,129,l30,l76-l79 93,126,]37,140,16I,l77.180,181 165,166,182 - 183 Heck, Stllle, Suzuki, Sonogashira, Negishi,184 and Kumada185 ’186 coupling reactions, oxidation of CO using molecular oxygen,77’82‘187"194 Mannich reactions,74 and Pauson—Khand reactions (Scheme 1.4) .141,195 Co NPs, Pd NPs CH3 MeozcL on Sio2 MeOfi)“: MeOZC WOAC MeOZC Scheme 1.4. Pauson—Khand reactions catalyzed by Co and Pd NPs supported on silica for a one-pot syntheses of bicyclicenones. Catalyst particles are anchored to oxide supports through a metal-support interaction (MSI), which prevents NP aggregation and can potentially influence the activity of the catalyst.”””’197 Studies indicate that the strength of the MSI depends on both the support and the metal. For example, the extent of aggregation of supported silver NPs in a vacuum decreases in the order Si02>A1203>C, while for Pt nanoparticles, the extent of interaction decreases in the order A1203> Si02>C.'98 In particular, alumina powders have surface areas that range from 0.1 to 400 m2/ g, pore volumes between 0.1 and 1.5 cm3/g, and average pore sizes between 2 nm and 170 199 pm The large surface area seem to help keep prevent NP aggregation and increase NP 200,201 loadin . Thus, aluminum oxide powder is erha s the most attractive su ort g P P PP material for NP catalysis. Additionally, alumina membranes can be used as a support to catalytic NPs.202’203 1.3.2. Solution-based Metal NP Stabilizers Surfactants Surfactants are common stabilizers of metal-NP suspensions. These molecules interact with small NPs to reduce their surface energy. A recent report indicates that suspensions of Pd NPs stabilized by N,N-dimethy1-N-cetyl-N-(2-hydroxyethyl) ammonium chloride successfully catalyze hydrogenation and dehalogenation of various halogenoarenes.204 Cetyltrimethylammonium bromide also stabilizes Cu NPs on several metal oxides, and these supported Cu NPs were employed for the selective dehydrogenation of methanol to produce formaldehyde.205 Metal oxide supports help to recover the catalyst at the end of the reaction. Ligands The use of ligands as NP stabilizers is especially interesting because variation of the ligand composition allows some optimization of the NP reactivity for use in specific catalytic reactions. Treatment of molecular Pd complexes with CO in DMF or toluene at room temperature produces Pd NPs that catalyze nucleophilic 10 substitution/carbonylation/amination to afford isoindolinones at room temperature as shown in the scheme 1.5.206 Pd NPs capped with special ligands such as cyclodextrins o l 1 mol % Pd(OAC)2. co (1 atm)/RNH2 cc - 3’ K2003, DMF \ I / CE/NHR Scheme 1.5. 3-component cascade process by palladium NPs generated in situ to form isoindolinone. are active for the catalysis of the hydrogenation of unsaturated substrates and of the Suzuki, Heck, and Stille coupling reactions.94'96 Dodecanethiolate-capped Pd NPs catalyze the Suzuki-Miyaura reaction of aryl halides, including, with phenylboronic acid at ambient temperature.93 Another simple method of stabilization involves addition of OH OSiMDBT HO O Pd-nanoparticles _ HO O HO TBDMSi-H, DMA TBDMSiO HO OCH3 25°C, -H2 H0 OCHs TBDMSiH= t-BuMeZSiH DMA = Dimethylacetamide Scheme1.6. Regioselective silylation of sugars through Pd NP-catalyzed silane alcoholysis. 11 silanes (R3SiH), such as tert—butyldimethylsilane, to PdC12 in N, N-dimethylacetamide. The black NP solution formed in this way catalyzes silane alcoholysis of sugars207 (Scheme 1.6), and selective cross—coupling of the silane with phenyl and vinyl thioethers gives the corresponding thiosilanes and silthianes.208 Polymers Polymers are particularly attractive stabilizers because they weakly bind metal nanoparticles via heteroatoms in the interior of the polymer and build a protective “shell” of polymer chains. In addition to stabilization, the steric bulk of the polymer framework may restrict access to metal NPs to impart selectivity. The polymer can be a soluble linear or branched macromolecule that “wraps” around the metal NP in solution to prevent metal aggregation and precipitation. The most commonly encountered polymer stabilizers are commercially available poly(N-vinyl-Z-pyrrolidone) (PVP),209'2'2 polyurea,213 polyacrylonitrile,”4 poly(acrylic 215 5 acid),214 polysiloxane, oligosaccharides,l4 poly(4-vinylpyridine),216 poly(N,N- l9 dialkylcarbodiimide),217 polyethylene glycol,218 chitosan,2 and a hyperbranched aromatic polyamide.220 Metal NPs protected in these polymers are used as soluble catalysts, and they can also be immobilized in ionic liquids for various kinds of reactions. Hirai’s pioneering work in the 19705 and 19805 is especially impressive. He generated Rh NPs using aqueous methanol or NaOH in methanol as a reducing agent for RhC13-3H20 to produce polyvinyl alcohol (PVA)-stabilized Rh NPs.22"222 Toshima pursued research on metal-NP catalysts in the early 19905 with polymer-stabilized bimetallic NPs generated by direct co-reduction of salts of two different metals (Pt and 12 Au for example), by sequential reduction (the initial NPs serve as seeds for the surface condensation of the second metal), or by galvanometric reduction of a metal salt by NPs of another more easily reduced metal.2°9’223’224 In the late 19905, Andres et al. synthesized PVP-stabilized Pt, Pd and Rh NPs by ethanolic reduction of the corresponding metal halide. These NPs were immobilized in an ionic liquid, 1-n-butyl-3- methylimidazolium hexafluorophosphate ([BMI][PF6]), and found to be very efficient catalysts for benzene hydrogenation at 40 °C.2'0 The polymer-stabilized metal NPs could also be recycled without loss of activity. Additionally, Pd and bimetallic (PdAu, PdPt) NPs were stabilized in polystyrene-block-poly-4-vinylpyridine (PS-b-P4VP) micelles and studied in the selective hydrogenation of dehydrolinalool as shown in the scheme 1.7.225’226 Cyclohexene hydrogenation was catalyzed with Pd NPs stabilized by highly branched amphiphilic polyglycerol.227 \ - - - \ \ Catalyst. PS b-P4VP Pd/AIzOgA' / \ OH H2 (1 atm), Toluene, 90°C 0” selectivi 99.8"/ b GC . Dehydrolinalool ty ° y llnalool Scheme 1.7. Selective hydrogenation of Dehydrolinalool to linalool by PS-b-P4VP-Pd on alumina. Dendrimers Dendrimers are a unique class of polymeric materials that often find application as templates to stabilize metal NPs. The formation of NPs stabilized by dendrimers was 228.229 230,231 232-234 reported in 1998 by the research groups of Crooks, Tomalia and Esumi. Crooks pioneered the use of dendrimer-encapsulated metal NPs for catalysis. In his 13 approach, metal cations such as Pd2+, Pt“, or Au3+ that are complexed with the nitrogen atoms of interior tertiary amines are reduced to NPs upon treatment with NaBH4 as Metal Cation NaBH4 — fl 0 Dendrimer—metal cation Dendrimer-metal NP Dendrimer . = Metal NP Figure 1.3. Encapsulation of metal nanoparticles in a dendrimer matrix.235 3228335440 The procedure used by Crooks and Tomalia illustrated in figure 1 encapsulated metal NPs inside the dendrimers, whereas Esumi’s process incorporated NPs at the dendrimer periphery. Dendrimer-encapsulated metal NPs are attractive for catalysis for several reasons. First, the dendrimer templates are fairly uniform structures, and they yield well-defined NP5.229'24"242 Second, NPs are stabilized by encapsulation within the dendrimer through weak interactions with amines so they retain their reactivity but do not agglomeratem'244 Third, the dendrimer branches can be used as selective gates to limit the access of large substrates to the encapsulated nanoparticlesumu Finally, the terminal groups on the dendrimer periphery can be tailored to control the solubility of the hybrid nanocomposite?“249 1.4. Layer-by-Layer Assembly of Oppositely Charged Polyelectrolytes The previous sections described a number of ways to stabilize NPs, including stabilization with polymers. Because this research specifically employs NPs in multilayer polyelectrolyte films, this section discusses the layer-by-layer assembly of such films. In 1966, Iler reported the formation of layers of charged particles by adsorption from solutions.250 In the early 1990’525"252 Decher and coworkers rediscovered surface modification through layer-by-layer deposition, and greatly expanded the technique by using sequential assembly of oppositely charged polymers. This pioneering work spawned thousands of subsequent studies because of the simplicity and versatility of the approach. The stepwise growth of multilayer polyelectrolyte films (Figure 1.4) is an entropy driven process due to the release of salt counter ions that were 253.254 originally neutralizing the polyelectrolytes in solution. The stability of each polyanion/polycation “bilayer” is presumably maintained through the formation of 5 (Other interactions multiple ion pairs between oppositely charged polyelectrolytes.25 such as hydrogen bonding can also be employed for layer-by-layer deposition,256 but this work focuses solely on electrostatic interactions.) The procedure for layer-by-layer deposition of polyelectrolyte films is very simple and begins with immersion of a charged substrate in a solution containing an oppositely charged polyelectrolyte. The polymer adsorbs on the surface, presumable due to a gain in entropy through the release of many counter ions, as mentioned above. The presence of the single-layer coating on the surface reverses the sign of the substrate charge. After rinsing to remove loosely attached polyelectrolyte, immersion of the substrate into a solution containing a polyelectrolyte of opposite charge gives a bilayer consisting of two oppositely charged polyelectrolytes, as 15 depicted schematically in the Figure 1.4.257 Additional bilayers are added similarly (Figure 1.4). Polyanion solution Polycation solution ”l i) Polyanion ii) Rinse L > i) Polycation ii) Rinse C _/ Substrate 1-bilayer Formation Figure 1.4. Formation of bilayer on the substrate through layer-by-layer deposition of polyelectrolytes.257 Multilayer polyelectrolyte films should consist of stratified structures because they are formed via sequential adsorption of polycations and polyanions. However, layers are highly intertwined in most cases due to the loops and tails in polyions and interactions between polyelectrolytes in several different layersm'259 These tails and loops are partially responsible for the charge overcompensation at the film surface. 260 Additionally, the entanglement of neighboring layers also aids charge compensation 16 within the multilayer films. Several studies did not detect any counter ions in the bulk of polyelectrolyte films, which suggested the complete compensation of opposite charge in the films.260’261 However, more recent studies show that charge compensation depends on the compositions of the film as well as deposition and post-deposition treatments.262 The layer-by-layer deposition method has several attractive features. 1. It is an easy and convenient method that usually involves a simple dip and rinse procedure. 2. Film thickness can be controlled at the nm scale by varying the number of layers 257363 and the ionic strength or the pH of the deposition solutions.263'266 The pH controls the charge density of weak polyelectrolytes by changing the degree of protonation or deprotonation, and polyelectrolytes with high charge densities typically give thin films with polymer chains are highly extended. A high concentration of salt in the deposition solution usually increases the film thickness because it shields neighboring charges on the polyelectrolytes and allows the formation of loops and tails. 3. Films can form on any surface that will adsorb polyelectrolytes. This requirement is easily met because most common solid substrates are charged. Adsorption is also relatively independent of 267 substrate geometry and can occurs on materials ranging from small particles to test 268 3 tubes to spongy membranes20 . 4. Many different polyelectrolytes are available for deposition. This technique initially started primarily with polyamines and polysulfonates, but subsequent reports deomonstrated layer-by-layer deposition of electro-active 269.270 polymers, quantum dots,271'273 DNA,274‘276 charged viruses,277’278 inorganic 269,279,280 281-284 285,286 sheets, proteins, and enzymes. 5. Film formation is relatively environmentally friendly in that it typically uses water as the solvent. Due to the 17 advantages of layer-by-layer assembly, these films. are being examined for applications in areas such as sensors, electronics, tissue engineering, and biotechnology. 1.5. Motivation and Research Goal Section 1.3 of this chapter indicates that a plethora of methods were developed to obtain stable metal NPs for use in catalysis. However, systematic studies of selective catalysis with a series of well-controlled, highly reactive NPs are lacking. There are two important challenges to prepare homogeneous/heterogeneous NPs for use as efficient catalysts. Firstly, to obtain maximum catalytic efficiency, the NPs must be stabilized to eliminate aggregation without blocking most of the active sites on the NP surfaces. Secondly, the stabilizer should afford control over NP size and reactivity to manipulate catalytic activity. We envisioned that the incorporation of metal NPs in thin polyelectrolyte multilayers prepared by layer-by-layer deposition could address these challenges quite effectively. To form NPs in these films, we first adsorb polyelectrolyte complexes of transition-metal ions and subsequently reduce these ions to form metal NPs. Variation of the concentration of metal ions in the film affords control over the NP size, and the polyelectrolyte film may also serve as a filter to enhance selectivity. Previous work in our group demonstrated the preparation of Pd NPs incorporated in poly(acrylic acid)(PAA)/polyethylenimine (PEI) films on alumina ([PAA/PEI-Pd(0)]3) and their catalytic activity (Figure 1.5). Those studies demonstrated moderate catalytic selectivity with allylic alcohols that had different substituents at the position a to the double bond as shown in the figure 1.5. However, that research did not investigate the 18 effect of NP size on selectivity or investigate a very wide range of hydrogenation reactions. K 1 + H2 + H2 Figure 1.5. Interrnolecularly selective hydrogenation of allyl alcohols by [PAA/PEI- Pd(0)]3PAA on alumina. This research examines selective, NP-catalyzed hydrogenation of a much broader series of compounds including allylic alcohols with various substituents on the double bond, molecules with more than one double bond, and nitroaromatic compounds. The NP catalysts show selectivities >100 for hydrogenation of mono over disubstituted double bonds, and these selectivities vary dramatically with NP size. These studies are predicated on careful formation and characterization of films with controlled NP size. Both intramolecular and intermolecular selectivities can be achieved. Additionally, the NP films are chemoselective in the hydrogenation of nitroaromatic compounds. These studies certainly demonstrate the unique properties of NP catalysts in selective reactions. 1.6. Outline of the Thesis In this introduction, I have attempted to provide background on NP stabilization and catalysis with NPs to show how my work fits in the broader context of this field. Chapter 2 expands on our previous work with Pd NPs polyelectrolyte films and shows unprecedented selectivities in hydrogenation of mono and multisubstituted double bonds. Initially, the chapter shows that Pd NPs embedded in multilayer polyelectrolyte films can be easily prepared through layer-by-layer adsorption of poly(acrylic acid) (PAA) and poly(ethylenimine)—Pd(II) (PEI-Pd(II)) complexes followed by reduction of Pd(II) with NaBI-Ia. Transmission electron microscopy confirms the formation of Pd particles with Scheme 1.8. Regio selective hydrogenation of Linalool to dihydrolinalool by [PAA/PEI- Pd(0)]3PAA films on alumina diameters of 1-3 nm. Remarkably, [PAA/PEI-Pd(0)]3PAA films catalyze the hydrogenation of allyl alcohols (monosubstituted unsaturated alcohols) with turnover frequencies (TOFs) that are as much as lOO-fold higher than turnover frequencies for hydrogenation of 3-methyl-2-buten-1-ol (multisubstituted unsaturated alcohols). Intramolecular selectivity for the hydrogenation of monosubstituted unsaturated alcohols also occurs when substrate molecules contain both mono and multiply substituted double 20 bonds, and one example is illustrated in scheme 1.8. The combination of the encapsulating polyelectrolyte film and small NPs apparently results in hindered access of multiply substituted double bonds to catalytic sites. Chapter 3 investigates selectivity in hydrogenation as a function of NP size, and shows that in the hydrogenation of mono and multisubstituted double bonds, selectivity increases dramatically as NP diameter decreases from 3.9 to 2.2 nm. The use of different ratios of poly(acrylic acid) to Pd(II) in deposition solutions for PAA-Pd(0)/PEI films gives a series of catalysts with Pd NPs whose average diameters range from 2.2 to 3.4 nm. Turnover frequencies (TOFs) for the hydrogenation of monosubstituted unsaturated 1 k 240 relative_ relative- krelative=3 k Scheme 1.9. Selectivity as a function of NP size in the catalytic hydrogenation of unsaturated alcohols. alcohols increase with decreasing average NP size, whereas multisubstituted unsaturated alcohols show the opposite trend. Hence, the ratios of TOFs for hydrogenation of allyl alcohol and B-methallyl alcohol are as high as 240 with the smallest NPs as illustrated in Scheme 1.9. Presumably, high selectivities occur on the smallest NPs because 21 hydrogenations of monosubstituted and multisubstituted double bonds take place on different types of atoms (e. g., terraces versus edges) on the Pd NPs. Chapter 4 expands the scope of [PAA-Pd(O)/PEI]3 catalysts and demonstrates chemoselective hydrogenation of aromatic nitro compounds. Chemoselective hydrogenation of nitroarenes is often challenging due to the formation of unwanted side products. The [PAA-Pd(0)/PEI]3 catalysts allow hydrogenation of various nitroarenes to the corresponding amines with functional group tolerance in high yield (>90 %) and short reaction times (15 to 60 mins). Higher chemoselectivity is observed when the hydrogenation reaction is performed in a mixture of water and Ethanol than in pure Ethanol. In chapter 5, I provide concluding remarks along with future prospects for these catalysts. 22 1.7. References (l) Haruta, M. Chemical Record 2003, 3, 75. (2) Zanella, R.; Louis, C.; Giorgio, S.; Touroude, R. Journal of Catalysis 2004, 223, 328. (3) Klabunde, K. J. In Free Atoms, Clusters, and Nanoscale Particles; Academic Press: San Diego, 1994, pp 2 & 36. (4) Klabunde, K. J .; Mohs, C. In Chemistry of Advance Materials: An Overview; Interrante, L. V.; Hampden-Smith, M. J., Eds., Wiley-VCH: New York, 1998, pp 271. (5) Stringer, R. C.; Schommer, S.; Hoehn, D.; Grant, S. A. Sensors and Actuators, B: Chemical 2008, 134, 427. (6) Scodeller, P.; F lexer, V.; Szamocki, R.; Calvo, E. J .; Tognalli, N.; Troiani, H.; F ainstein, A. Journal of the American Chemical Society 2008, 130, 12690. (7) Du, D.; Chen, 8.; Song, D.; Li, H.; Chen, X. Biosensors & Bioelectronics 2008, 24, 475 (8) Liu, X.; Niu, W.; Li, H.; Han, 8.; Nu, L.; Xu, G. Electrochemistry Communications 2008, 10, 1250. (9) Tallury, P.; Payton, K.; Santra, S. Nanomedicine 2008, 3, 579. (10) Guo, W.; Yuan, J.; Li, B.; Du, Y.; Ying, B; Wang, E. Analyst 2008, 133, 1209. (11) Xiang, C.; Zou, Y.; Sun, L.-X.; Xu, F. Analytical Letters 2008, 41, 2224. (12) Kim, G.-Y.; Kang, M.-S.; Shim, J .; Moon, S.-H. Sensors and Actuators, B: Chemical 2008, 8133, l. (13) Rezaei-Zarchi, S.; Saboury, A. A.; Javed, A.; Barzegar, A.; Ahmadian, S.; Bayandori-Moghaddam, A. Journal of Biosciences 2008, 33, 279. (14) Heltzel, A. J .; Qu, L.; Dai, L. Nanotechnology 2008, 19, 245702/ 1. (15) Salsamendi, M.; Marcilla, R.; Dobbelin, M.; Mecerreyes, D.; Pozo-Gonzalo, C.; Pomposo, J. A.; Pacios, R. Physica Status Solidi A: Applications and Materials Science 2008, 205, 1451. (16) Silva, L. B.; Baptista, P.; Raniero, L.; Doria, G.; Martins, R.; Fortunate, E. Sensors and Actuators, B: Chemical 2008, BI 32, 508. 23 (17) Lee, C.-W.; Lee, K.-W.; Lee, J .-S. Materials Letters 2008, 62, 2664. (18) Kim, B.-H.; Cho, M.-S.; Kim, S.-Y.; Kim, Y.-J.; Woo, H.-G.; Kim, D.-H.; Sohn, H. Journal of Nanoscience and Nanotechnology 2007, 7, 3926. (19) Sultana, N.; Wang, M. Journal of Experimental Nanoscience 2008, 3, 121. (20) Sasaki, T.; Iwasaki, N.; Kohno, K.; Kishimoto, M.; Majima, T.; Nishimura, S.-l.; Minami, A. Journal of Biomedical Materials Research, Part A 2008, 86A, 969. (21) Oliveira, J. M. Kotobuki, N. ,,Marques A. P.; Pirraco, R. P.; Benesch, J. Hirose, M. ,,Costa S. A. ,Mano, J. F, Ohgushi, H. ,Reis, R. L. Advanced Functional Materials 2008,181840. (22) Yang, S.-Y.; Sun, J.-S.; Liu, C.-H.; Tsuang, Y.-H.; Chen, L.-T.; Hong, C.-Y.; Yang, H.-C.; Homg, H.-E. Artificial Organs 2008, 32, 195. (23) Hosseinkhani, H.; Hosseinkhani, M.; Gabrielson, N. P.; Pack, D. W.; Khademhosseini, A.; Kobayashi, H. Journal of Biomedical Materials Research, Part A 2008, 85A, 47. (24) Elaissari, A. Colloidal Nanoparticles in Biotechnology; Wiley-VC H: New York, 2008. (25) Jaffrezic-Renault, N.; Martelet, C.; Chevolot, Y.; Cloarec, J .-P. Colloidal Nanoparticles in Biotechnology 2008, 169. (26) Ni, P. Colloidal Nanoparticles in Biotechnology 2008, 31. (27) Parsons, J. G.; Peralta-Videa, J. R.; Gardea-Torresdey, J. L. Developments in Environmental Science 2007, 5, 463. (28) Majewski, P.; Thierry, B. Critical Reviews in Solid State and Materials Sciences 2007, 32, 203. (29) Weber, A.; Gruber-Traub, C.; Herold, M.; Borchers, K.; Tovar, G. E. M. NanoS 2006, 2, 20. (30) Vallribera, A.; Molins, E. Nanoparticles and Catalysis 2008, 161. (31) Astruc, D. Nanoparticles and Catalysis 2008, 1. (32) Li, D.; Dunlap, J. R.; Zhao, B. Langmuir 2008, 24, 5911. (33) Liu, Z.; Reed, D.; Kwon, G.; Shamsuzzoha, M.; Nikles, D. B. Journal of Physical Chemistry C 2007, 111, 14223. 24 (34) Srimani, D.; Sawoo, S.; Sarkar, A. Organic Letters 2007, 9, 3639. (35) Merga, G.; Wilson, R.; Lynn, G.; Milosavljevic, B. H.; Meisel, D. Journal of Physical Chemistry C 2007, 111, 12220. (36) Singla, M. L.; Negi, A.; Mahajan, V.; Singh, K. C.; Jain, D. V. S. Applied Catalysis, A: General 2007, 323, 51. (37) Pong, W.-Y.; Chang, H.-Y.; Liang, C.-H.; Chen, H.-I. Advanced Materials Research 2007, 15-17, 553. (38) Zhang, Z.; Zha, Z.; Gan, C.; Pan, C.; Zhou, Y.; Wang, Z.; Zhou, M.-M. Journal of Organic Chemistry 2006, 71, 4339. (39) Corma, A.; Sema, P.; Garcia, H. Journal of the American Chemical Society 2007, 129, 6358. (40) Park, K. H.; Jang, K.; Kim, H. J .; Son, S. U. Angewandte Chemie International Edition 2007, 46, 1152. (41) Mei, Y.; Lu, Y.; Polzer, F.; Ballauff, M.; Drechsler, M. Chemistry of Materials 2007, I9, 1062. (42) Park, C. M.; Kwon, M. S.; Park, J. Synthesis 2006, 3790. (43) Oyamada, H.; Akiyama, R.; Hagio, H.; Naito, T.; Kobayashi, S. Chemical Communications 2006, 4297. (44) Yan, W.; Wang, R.; Xu, Z.; Xu, J .; Lin, L.; Shen, Z.; Zhou, Y. Journal of Molecular Catalysis A: Chemical 2006, 255, 81. (45) Corma, A.; Sema, P. Science 2006, 313, 332. (46) Jiang, Y.; Gao, Q. Journal of the American Chemical Society 2006, 128, 716. (47) Mei, Y.; Sharma, G.; Lu, Y.; Ballauff, M.; Drechsler, M.; lrrgang, T.; Kempe, R. Langmuir 2005, 21, 12229. (48) Narayanan, R.; El-Sayed, M. A. Journal of Catalysis 2005, 234, 348. (49) Oh, S.-K.; Niu, Y.; Crooks, R. M. Langmuir 2005, 21, 10209. (50) Berzelius, J. J. New Philosophy of Journal 1836, 200, 223. (51) In R&D Magazine 2005, pp 20. 25 (52) Alexeev, O. 8.; Gates, B. C.Industrial & Engineering Chemistry Research 2003, 42 1571. (53) Thomas, J. M.; Johnson, B. F. G.; Raja, R.; Sankar, G.; Midgley, P. A. Accounts of Chemical Research 2003, 36, 20. (54) Kralik, M.; Biffis, A. Journal of Molecular Catalysis A: Chemical 2001, 1 77, 113. (55) St. Clair, T. P.; Goodman, D. W. Topics in Catalysis 2000, 13, 5. (56) Henry, C. R. Applied Surface Science 2000, 164, 252. (57) Bond, G. C. Heterogeneous Catalysis: Principles and Applications; Oxford Science: Oxford, 1987. (58) Handbook of Heterogeneous Catalysis; Ertl, G.; Kndzinger, H.; Weitkamp, J., Eds, Wiley -VCH: Weinheim, 1997. (59) Catalysis by Metals and Alloys; Ponec, V.; Bond, G. C., Eds., Elsevier: Amsterdam, 1995; Vol. 95. (60) Herrmann, W. A.; Comils, B. Angewandte Chemie International Edition 1997, 36, 1048. (61) Astruc, D. Organometallic Chemistry and Catalysis; Springer: Heidelberg, 2007, part 4 and 5. (62) Applied Homogeneous Catalysis with Organometallic Compounds; Comils, B.; Herrmann, W. A., Eds.,Wiley-VCH: Weinheim, 1996; Vol. 1 and 2. (63) Metal-Catalyzed Cross-Coupling Reactions; Meijere, A. D.; Diederich, F ., Eds., Wiley-VCH Weinheim, 1998. (64) Lewis, L. N. Chemical Reviews 1993, 93, 2693. (65) Bradley, J. S. In Clusters Colloids; Schmid, G., Ed., Wiley-VCH: Weinheim, 1994, pp 459. (66) Omelas, C.; Me'ry, D.; Blais, J. -C; Cloutet, E.; Ruiz, J .; Astruc, A.; Astruc, D. Angewandte Chemie International Edition 2005, 44, 7399. (67) Vries, J. G. D. Dalton Transactions 2006, 421. (68) Astruc, D. Inorganic Chemistry 2007 46, 1884. 26 (69) Schauermann, S.; Hoffmann, J .; Johanek, V.; Hartmann, J.; Libuda, J .; Freund, H.-J. Angewandte Chemie International Edition 2002, 41, 2532. (70) Bertarione, S.; Scarano, D.; Zecchina, A.; Johanek, V.; Hoffmann, J .; Schauermann, S.; Libuda, J .; Rupprechter, G.; F reund, H.-J. Journal of Catalysis 2004, 223, 64. (71) Riahi, G.; Guillemot, D.; Polisset-Thfoin, M.; Khodadadi, A. A.; Fraissard, J. Catalysis Today 2002, 72, 115. (72) Scire, S.; Minico, S.; Crisafirlli, C.; Satriano, C.; Pistone, A. Applied Catalysis B: Environmental 2003, 40, 43. (73) Yu, K. M. K.; Yeung, C. M. Y.; Thompsett, D.; Tsang, S. C. Journal of Physical Chemistry B 2003, 107, 4515. (74) Manabe, K.; Mori, Y.; Wakabayashi, T.; Nagayama, S.; Kobayashi, S. Journal of the American Chemical Society 2000, 122, 7202. (75) Soulivong, D.; Christophe. C.; Thivolle-Cazat, J .; Basset, J. -M.; Maunders, B. M.; Richard, B. A.; Sunley, G. J .; Sunley, G. J. Angewandte Chemie International Edition 2004 43, 5366. (76) Mori, K.; Hara, T.; Mizugaki, T.; Ebitani, K.; Kaneda, K. Journal of the American Chemical Society 2004, 126, 10657. (77) Campbell, C. T. Science 2004, 306, 234. (78) Bertoux, F .; Monflier, E.; Castanet, Y.; Mortreux, A. Journal of Molecular Catalysis A: Chemical 1999, 143, 23. (79) He, J .; Ichinose, 1.; Kunitake, T.; Nakao, A.; Shiraishi, Y.; Toshima, N. Journal of the American Chemical Society 2003, 125, 11034. (80) Maye, M. M.; Luo, J.; Lin, Y.; Engelhard, M. H.; Hepel, M.; Zhong, C.-J. Langmuir 2003, 19, 125. (81) Roth, C.; Hussain, 1.; Bayati, M.; Nichols, R. J .; Schiffrin, D. J. Chemical Communications 2004, 1532. (82) Daniel, M.-C.; Astruc, D. Chemical Reviews 2004, 104, 293. (83) Liu, Z.; Ling, X. Y.; Su, X.; Lee, J. Y. Journal of Physical Chemistry B 2004, 108, 8234. 27 (84) Bock, C.; Paquet, C.; Couillard, M.; Botton, G. A.; MacDougall, B. R. Journal of the American Chemical Society 2004, 126, 8028. (85) Veith, G. M.; Lupini, A. R.; Pennycook, S. J .; Ownby, G. W.; Dudney, N. J. Journal of Catalysis 2005, 231, 151. (86) Harman, W. D. Chemical Reviews 1997, 97, 1953. (87) Lin, Y.; F inke, R. G. Inorganic Chemistry 1994, 33, 4891. (88) Yang, H.; Gao, H.; Angelici, R. J. Organometallics 2000, I9, 622. (89) Pellegatta, J .-L.; Blandy, C.; Colliere, V.; Choukroun, R.; Chaudret, B.; Cheng, P.; Philippot, K. Journal of Molecular Catalysis A: Chemical 2002, I 78, 55. (90) Schulz, J .; Roucoux, A.; Patin, H. Chemical Communications 1999, 535. (91) Schulz, J .; Levigne, S.; Roucoux, A.; Patin, H. Advanced Synthesis & Catalysis 2002, 344, 266. (92) Roucoux, A.; Schulz, J .; Patin, H. Advanced Synthesis & Catalysis 2003, 345, 222. (93) Lu, F .; Ruiz, J .; Astruc, D. Tetrahedron Letters 2004, 45, 9443. (94) Liu, J .; Alvarez, J .; Ong, W.; Roman, E.; Kaifer, A. E. Langmuir 2001, 17, 6762. (95) Alvarez, J .; Liu, J .; Roman, E.; Kaifer, A. E. Chemical Communications 2000, 1151. (96) Strimbu, L.; Liu, J .; Kaifer, A. E. Langmuir 2003, I 9, 483. (97) Schmid, G.; Maihack, V.; Lantermann, F .; Peschel, S. Journal of the Chemical Society, Dalton Transactions: Inorganic Chemistry 1996, 589. (98) Sharma, J .; Chhabra, R.; Yan, H.; Liu, Y. Chemical Communications 2008, 2140. (99) Arai, M.; Miyake, M.; Yamada, M. Journal of Physical Chemistry C 2008, 112, 1953. (100) Vilain, C.; Goettmann, F .; Moores, A.; Le Floch, P.; Sanchez, C. Journal of Materials Chemistry 2007, I 7, 3509. (101) Bell, P. W.; Anand, M.; Fan, X.; Enick, R. M.; Roberts, C. B. Langmuir 2005, 21 , 11608. 28 (102) Kidambi, S.; Dai, J.; Li, J .; Bruening, M. L. Journal of the American Chemical Society 2004, 126, 2658. (103) Kidambi, S.; Bruening, M. L. Chemistry of Materials 2005, 17, 301. (104) Faraday, M. Philosophical Transactions of the Royal Society London 1857, 183. (105) Thomas, J. M. Pure and Applied Chemistry 1988, 60, 1517. (106) Rapino, L. D.; Nord, F. F. Journal of the American Chemical Society 1941, 63, 2745. (107) Rapino, L. D.; Nord, F. F. Journal of the American Chemical Society 1941, 63, 3268. (108) Kavanagh, K. E.; Nord, F. F. Journal of the American Chemical Society 1943, 65, 2121. (109) Parravano, G. Journal of Catalysis 1970, I8, 320. (110) Haruta, M.; Kobayashi, T.; Sano, H.; Yamada, N. Chemistry Letters 1987, 405. (111) Haruta, M.; Yamada, N.; Kobayashi, T.; Iijima, S. Journal of Catalysis 1989, I I 5, 301. (112) Haruta, M.; Tsubota, S.; Kobayashi, T.; Kageyama, H.; Genet, M. J .; Delmon, B. Journal of Catalysis 1993, 144, 175. (113) Haruta, M. Cattech 2002, 6, 102. (114) Moradpour, A.; Amouyal, E.; Keller, P.; Kagan, H. Nouveau Journal de Chimie 1978, 2, 547. (115) Henglein, A. Journal of Physical Chemistry 1979, 83, 2858. (116) Henglein, A.; Lilie, J. Journal of the American Chemical Society 1981, 103, 1059. (117) Kurihara, K.; F endler, J. H.; Ravet, I.; Nagy, J. B. Journal of Molecular Catalysis. 1986, 34, 325. (118) Boutonnet, M.; Kizling, J.; Touroude, R.; Maire, G.; Stenius, P. Applied Catalysis 1986, 20, 163. (119) Degani, Y.; Willner, 1. Journal of the Chemical Society, Perkin Transactions 2 Physical Organic Chemistry 1986, 37. 29 (120) Willner, I.; Maidan, R.; Mandler, D.; Duerr, H.; Doerr, G.; Zengerle, K. Journal of the American Chemical Society 1987, 109, 6080. (121) Bradley, J. S.; Hill, E.; Leonowicz, M. E.; Witzke, H. Journal of Molecular Catalysis 1987, 41 , 59. (122) Larpent, C.; Patin, H. Journal of Molecular Catalysis 1988, 44, 191. (123) Nanoparticles and Nanostructured Films Preparation, Characterizations and Applications; F endler, J. H., Ed., Wiley-VCH: Weinheim, 1998. (124) Reetz, M. T.; Helbig, W. Journal of the American Chemical Society 1994, I 16, 7401. (125) Reetz, M. T.; Quaiser, S. A. Angewandte Chemie International Edition 1995, 34, 2240. (126) Reetz, M. T.; Breinbauer, R.; Wanninger, K. Tetrahedron Letters 1996, 37, 4499. (127) Reetz, M. T.; Lohmer, G. Chemical Communications 1996, 1921. (128) Reetz, M. T.; Westermann, E; Lohmer, R.; Lohmer, G. Tetrahedron Letters 1998, 39, 8449. (129) Reetz, M. T.; Maase, M. Advanced Materials 1999, 11, 773. (130) Reetz, M. T.; Westermann, E. Angewandte Chemie International Edition 2000, 39, 165. (131) Corma, A.; Garcia, H.; Leyva, A. Journal of Molecular Catalysis A: Chemical 2005, 230, 97. (132) Mohr, C.; Hofmeister, H.; Claus, P. Journal of Catalysis 2003, 213, 86. (133) Son, S. U.; Lee, S. 1.; Chung, Y. K.; Kim, S.-W.; Hyeon, T. Organic Letters 2002, 4, 277. (134) Son, S. U.; Park, K. H.; Chung, Y. K. Organic Letters 2002, 4, 3983. (135) Son, S. U.; Park, K. H.; Chung, Y. K. Journal of the American Chemical Society 2002, 124, 6838. (136) Son, S. U.; Lee, S. 1.; Chung, Y. K. Angewandte Chemie International Edition 2000, 39, 4158. 30 (137) Lang, H.; Maldonado, S.; Stevenson, K. J .; Chandler, B. D. Journal of the American Chemical Society 2004, 126, 12949. (138) Lang, H.; May, R. A.; Iversen, B. L.; Chandler, B. D. Journal of the American Chemical Society 2003, 125, 14832. (139) Niederer, J. P. M.; Arnold, A. B. J .; Holderich, W. F .; Spliethof, B.; Tesche, B.; Reetz, M.; Bonnemann, H. Topics in Catalysis 2002, 18, 265. (140) Kim, S.-W.; Kim, M.; Lee, W. Y.; Hyeon, T. Journal of the American Chemical Society 2002, 124, 7642. (141) Park, K. H.; Son, S. U.; Chung, Y. K. Organic Letters 2002, 4, 4361. (142) Abu-Reziq, R.; Avnir, D.; Blum, J . Journal of Molecular Catalysis A: Chemical 2002, I87, 277. (143) Huang, A.; Liu, Y.; Chen, L.; Hua, J. Journal of Applied Polymer Science 2002, 85, 989. (144) Molnar, A.; Papp, A.; Miklos, K.; Forgo, P. Chemical Communications 2003, 2626. (145) Sanji, T.; Ogawa, Y.; Nakatsuka, Y.; Tanaka, M.; Sakurai, H. Chemistry Letters 2003, 32, 980. (146) Dominguez-Quintero, 0.; Martinez, S.; Henriquez, Y.; D'Ornelas, L.; Krentzien, H.; Osuna, J. Journal of Molecular Catalysis A: Chemical 2003, 197, 185. (147) Yuranov, I.; Moeckli, P.; Suvorova, E.; Buffat, P.; Kiwi-Minsker, L.; Renken, A. Journal of Molecular Catalysis A: Chemical 2003, 192, 239. (148) Corain, B.; Guerriero, P.; Schiavon, G.; Zapparoli, M.; Kralik, M. Journal of Molecular Catalysis A: Chemical 2004, 211, 237. (149) Kralik, M.; Kratky, V.; Centomo, P.; Guerriero, P.; Lora, S.; Corain, B. Journal of Molecular Catalysis A: Chemical 2003, 195, 219. (150) Johanek, V.; Laurin, M.; Hoffmann, J.; Schauermann, S.; Grant, A. W.; Kasemo, B.; Libuda, J .; Freund, H. J. Surface Science 2004, 561, L218. (151) Galow, T. H.; Drechsler, U.; Hanson, J. A.; Rotello, V. M. Chemical Communications 2002, 1076. (152) Bedford, R. B.; Singh, U. G.; Walton, R. 1.; Williams, R. T.; Davis, S. A. Chemistry of Materials 2005, 1 7, 701. 31 (153) Kormann, H.-P.; Schmid, G.; Pelzer, K.; Philippot, K.; Chaudret, B. Zeitschrift fuer Anorganische und Allgemeine Chemie 2004, 630, 1913. (154) Yashima, M.; Falk, L. K. L.; Palrnqvist, A. E. C.; Holmberg, K. Journal of Colloid and Interface Science 2003, 268, 348. (155) Heemeier, M.; Carlsson, A. F.; Naschitzki, M.; Schmal, M.; Baumer, M.; Freund, H.-J. Angewandte Chemie International Edition 2002, 41, 4073. (156) Choi, K.-M.; Akita, T.; Mizugaki, T.; Ebitani, K.; Kaneda, K. New Journal of Chemistry 2003, 27, 324. (157) Ebitani, K.; Fujie, Y.; K. Kaneda, K. Langmuir 1999, 15, 3557. (158) Guczi, L.; Beck, A.; Horvath, A.; Koppany, Z.; Stefler, G.; Frey, K.; Sajo, 1.; Geszti, 0.; Bazin, D.; Lynch, J. Journal of Molecular Catalysis A: Chemical 2004, 204, 545. (159) Wagner, M.; Kohler, K.; Djakovitch, L.; Muhler, M. Topics in Catalysis 1999, 13, 319. ' (160) Kohler, K.; Wagner, M.; Djakovitch, L. Catalysis Today 2001, 66, 105. (161) Djakovitch, L.; Koehler, K. Journal of the American Chemical Society 2001, 123, 5990. (162) Nejjar, A.; Pinel, C.; Djakovitch, L. Advanced Synthesis & Catalysis 2003, 345, 612. (163) Djakovitch, L.; Wagner, M.; Hartung, C. G.; Beller, M.; Koehler, K. Journal of Molecular Catalysis A: Chemical 2004, 219, 121. (164) Prockl, S. S.; Wolkgang, K.; Gruber, M. A.; Kohler, K. Angewandte Chemie International Edition 2004, 43, 1881. (165) Djakovitch, L.; Rollet, P. Advanced Synthesis & Catalysis 2004, 346, 1782. (166) Djakovitch, L.; Rollet, P. Tetrahedron Letters 2004, 45 , 1367. (167) Chouzier, S.; Gruber, M.; Dj akovitch, L. Journal of Molecular Catalysis A: Chemical 2004, 212, 43. (168) J. Penzien, Haessner, C.; Jentys, A.; Kohler, K.; Muller, T. E.; Lercher J. A. Journal of Catalysis 2004, I 21 , 302. 32 (169) T. Nishimura, T.; Kakiuchi, N.;Inoue, M.; Uemura, S. Chemical Communications 2000, 1245. (170) Kakiuchi, N.; Maeda, Y.; Nishimura, T.; Uemura, S. Journal of Organic Chemistry 2001, 66, 6220. (171) Choudary, B. M.; Madhi, S.; Chowdari, N. S.; Kantam, M. L.; Sreedhar, B. Journal of the American Chemical Society 2002, 124, 14127. (172) Mastalir, A.; Kiraly, Z. Journal of Catalysis 2003, 220, 372. (173) Pfeifer, P.; Schubert, K.; Liauw, M. A.; Emig, G. Applied Catalysis, A: General 2004, 270, 165. (174) Bertarione, S.; Scarano, D.; Zecchina, A.; Johanek, V.; Hoffmann, J.; Schauermann, S.; Libuda, J .; Rupprechter, G.; Freund, H.-J . Journal of Catalysis 2004, 223, 64. (175) Huang. J.; Jiang. T.; Gao,X.; Han, B. -X.; Liu, Z. -M.; Wu, W. -Z.; Chang, Y. -H.; Zhao, G.-Y. Angewandte Chemie International Edition 2004, 43.1397. (176) Reetz, M. T.; Breinbauer, R; Wedemann, P.; Binger, P. Tetrahedron 1998, 54, 1233. (177) Reetz, M. T.; de Vries, J. G. Chemical Communications 2004, 1559. (178) Little, A. F .; Fu, G. C. Angewandte Chemie International Edition 2002, 41 , 4176. (179) Choudary, B. M.; Chowdari, N. S.; Jyothi, K.; Kumar, N. S.; Kantam, M. L. Chemical Communications 2002, 586. (180) Thathagar, M. B.; Beckers, J .; Rothenberg, G. Journal of the American Chemical Society 2002, 124, 11858. (181) de Vries, J. G.; de Vries, A. H. M. European Journal of Organic Chemistry 2003, 799. (182) Son, S. U.; Jang, Y.; Park, J.; Na, H. B.; Park, H. M.; Yun, H. J.; Lee, J.; Hyeon, T. Journal of the American Chemical Society 2004, 126, 5026. (183) Zhao, D.; Fei, Z.; Geldbach, T. J .; Scopelliti, R.; Dyson, P. J . Journal of the American Chemical Society 2004, 126, 15876. (184) Lipshutz, B. H.; Blomgren, P. A. Journal of the American Chemical Society 1999, 121, 5819. 33 (185) Styring, P.; Grindon, C.; Fisher, C. M. Catalysis Letters 2001, 7 7, 219. (186) Lipshutz, B. H.; Tasler, S.; Chrisman, W.; Spliethoff, B.; Tesche, B. Journal of Organic Chemistry 2003, 68, 1177. (187) Jansen, J. F. G. A.; de Brabander van den Berg, E. M. M.; Meijer, E. W. Science 1994, 266, 1226. (188) Tomalia, D. A.; Naylor, A. M.; Goddard, W. A. Angewandte Chemie International Edition 1990, 29, 138. (189) Lopez, N.; ‘Norskov, J. K. Journal of the American Chemical Society 2002, 124, 11262. (190) Schimpf, 8.; Lucas, M.; Mohr, C.; Rodemerck, U.; Bruckner, A.; Radnik, J .; Hofmeister, H.; Claus, P. Catalysis Today 2002, 72, 63. ( 191) Zwijnenburg, A.; Saleh, M.; Makkee, M.; Moulijn, J. A. Catalysis Today 2002, 72, 59. (192) Date, M.; Ichihashi, Y.; Yamashita, T.; Chiorino, A.; Boccuzzi, F.; Haruta, M. Catalysis Today 2002, 72, 89. (193) Cuenya, B. R.; Baeck, S. -H.; Jaramillo, T. F.; McFarland E.W. Journal of the American Chemical Society 2003, 125, 12928. (194) Date, M.; Okumura, M.; Tsubota, S.; Haruta, M. Angewandte Chemie International Edition 2004, 116, 2129. (195) Moreno-Manas, M.; Pleixats, R. Accounts of Chemical Research 2003, 36, 638. (196) Tauster, S. J. F.; Baker, R. T. K.; Horsley, J. A. Science 1981, 211, 1121. (197) Bond, G. C. Catalysis 1983, 6, 27. (198) Arai, M. 1.; Nakayama, T.; Nishiyama, Y. Journal of Colloid and Interface Science 1984, 97, 254. (199) Augustine, R. L. Heterogeneous Catalysis for the Synthetic Chemist; Marcel: New York, 1996. (200) Cassell, A M.; Raymakers, J. A.; Kong, J .; Dai, H. Journal of Physical Chemistry B 1999, 103, 6484. (201) Harutyunyan, A. R.; Pradhan, B. K.; Kim, U. J .; Chen, G.; Eklund, P. C. Nano Letters 2002, 525. 34 (202) Schmidt, A.; Haidar, R.; Schomacker, R. Catalysis Today 2005, 104, 305. (203) Dotzauer, D. M.; Dai, J .; Sun, L.; Bruening, M. L. Nano Letters 2006, 6, 2268. (204) Leger, B.; Nowicki, A.; Roucoux, A; Rolland, J. -P. Journal of Molecular Catalysis A: Chemical 2007, 266, 221. (205) Tada, M.; Bal, R.; Namba, S.; Iwasawa, Y. Applied Catalysis A: General 2006, 307, 78. (206) Grigg, R.; Zhang, L.; Collard, S.; Keep, A. Tetrahedron Letters 2003, 44, 6979. (207) Chung, M.-K.; Orlova, G.; Goddard, J. D.; Schlaf, M.; Harris, R.; Beveridge, T. J .; White, G.; Hallett, F. R. Journal of the American Chemical Society 2002, 124, 10508. (208) Chung, M.-K.; Schlaf, M. Journal of the American Chemical Society 2004, 126, 7386. (209) Toshima, N. In Fine Particles Sciences and T echnolog From Micro to New Particles; Pellizzetti, E., Ed., Kluwer: Dordrecht; 1996, pp. 371 (210) Andres, R. P.; Bielefeld, J. D.; Henderson, J. 1.; lanes, D. B.; Kolagunta, V. R.; Kubiak, C. P.; Mahoney, W. J .; Osifchin, R. G. Science 1996, 273, 1690. (211) Narayanan, R.; El-Sayed, M. A. Journal of the American Chemical Society 2003, 125, 8340. (212) Li, Y.; Boone, E.; El-Sayed, M. A. Langmuir 2002, 18, 4921. (213) Ley, S. V.; Mitchell, C.; Pears, D.; Ramarao, C.; Yu, J .-Q.; Zhou, W. Organic Letters 2003, 5, 4665. (214) Demir, M. M.; Gulgun, M. A.; Menceloglu, Y. Z.; Erman, B.; Abramchuk, S. S.; Makhaeva, E. E.; Khokhlov, A. R.; Matveeva, V. G.; Sulman, M. G. Macromolecules 2004, 37, 1787. (215) Chauhan, B. P. S.; Rathore, J. S.; Bandoo, T. Journal of the American Chemical Society 2004, I 26, 8493. (216) Drelinkiewicza, A.; Waksmundzka, A.; Makowski, W.; Sobczak, J. W.; Krol, A.; Zieba, A. Catalysis Letters 2004, 94, 143. (217) Liu, Y.; Chalermchai. K.; Hu, J. Chemical Communications 2004, 398. 35 (218) Pillai, U. R.; Sahle-Demessie, E. Journal of Molecular Catalysis A: Chemical 2004, 222, 153. (219) Adlim, M.; Abu Bakar, M.; Liew, K. Y.; Ismail, J. Journal of Molecular Catalysis A: Chemical 2004, 212, 141. (220) Tabuani, D.; Monticelli, 0.; Chincarini, A.; Bianchini, C.; Vizza, F .; Moneti, S.; Russo, S. Macromolecules 2003, 3 6, 4294. (221) Hirai, H.; Nakao, Y.; Toshima, N. Journal of Macromolecular Science, Chemistry 1979, I3, 727. (222) Hirai, H.; Nakao, Y.; Toshima, N. Journal of Macromolecular Science, Chemistry 1978, A12, 1117. (223) Toshima, N.; Yonezawa, T. New Journal of Chemistry 1998, 22, 1179. (224) Shiraishi, Y.; Toshima, N. Journal of Molecular Catalysis A: Chemical 1999, I 41 , 187. (225) Sulman, E.; Bodrova, Y.; Matveeva, V.; Semagina, N.; Cerveny, L.; Kurtc, V.; Bronstein, L.; Platonova, O.; Valetsky, P. Applied Catalysis, A: General 1999, 1 76, 75. (226) Sulman, E.; Bronstein, L.; Matveeva, V.; Sulman, M.; Lakina, N.; Doluda, V.; Valetsky, P. Nanocatalysis 2006, 51. (227) Sablong, R.; Schlotterbeck, U.; Vogt, D.; Mecking, S. Advanced Synthesis Catalysis 2003, 345, 333. ‘ (228) Crooks, R. M.; Zhao, M.; Sun, L.; Chechik, V.; Yeung, L. K. Accounts of Chemical Research 2001, 34, 181. (229) Zhao, M.; Sun, L.; Crooks, R. M. Journal of the American Chemical Society 1998, 120, 4877. (230) Balogh, L.; Tomalia, D. A. Journal of the American Chemical Society 1998, 120, 7355. (231) Balogh, L.; Swanson, D. R.; Tomalia, D. A.; Hagnauer, G. L.; McManus, A. T. Nano Letters 2001, 1, 18. (232) Esumi, K.; Isono, R.; Yoshimura, T. Langmuir 2004, 20, 237. (233) Esumi, K.; Sato, K.; Suzuki, A.; Torigoe, K. Shikizai Kyokaishi 2000, 73, 434. 36 (234) Esumi, K.; Suzuki, A.; Aihara, N.; Usui, K.; Torigoe, K. Langmuir 1998, 14, 3157. (235) Scott, R. W. J .; Wilson, 0. M.; Crooks, R. M. Journal of Physical Chemistry B 2005, 109, 692. (236) Knecht, M. R.; Weir, M. G.; Myers, V. S.; Pyrz, W. D.; Ye, H.; Petkov, V.; Buttrey, D. J .; Frenkel, A. 1.; Crooks, R. M. Chemistry of Materials 2008, 20, 5218. (237) Ye, H.; Crooks, J. A.; Crooks, R. M. Langmuir 2007, 23, 11901. (238) Ye, H.; Crooks, R. M. Journal of the American Chemical Society 2007, 129, 3627. (239) Ye, H.; Crooks, R. M. Journal of the American Chemical Society 2005, I 2 7, 4930. (240) Niu, Y.; Yeung, L. K.; Crooks, R. M. Journal of the American Chemical Society 2001, 123, 6840. (241) Zhao, M.; Crooks, R. M. Angewandte Chemie International Edition 1999, 38, 364. (242) Zhao, M.; Crooks, R. M. Advanced Materials 1999, 11, 217. (243) Chechik, V.; Crooks, R. M. Jaurnal of the American Chemical Society 2000, 122, 1243. (244) Chechik, V.; Zhao, M.; Crooks, R. M. Journal of the American Chemical Society 1999, 121, 4910. (245) Zhao, M.; Hideo, T.; Crooks, R. M. Angewandte Chemie International Edition 1997, 36, 2596. (246) Hierlemann, A.; Campbell, J. K.; Baker, L. A.; Crooks, R. M.; Ricco, A. J. Journal of the American Chemical Society 1998, 120, 5323. (247) Tokuhisa, H.; Crooks, R. M. Langmuir 1997, 13, 5608. (248) Tokuhisa, H.; Zhao, M.; Baker, L. A.; Phan, V. T.; Dermody, D. L.; Garcia, M. E.; Peez, R. F.; Crooks, R. M.; Mayer, T. M. Journal of the American Chemical Society 1998, 120, 4492. (249) Wells, M.; Crooks, R. M. Journal of the American Chemical Society 1996, I 18, 3988. 37 (250) Iler, R. K. Journal of Colloid and Interface Science 1966, 21, 569. (251) Decher, G; Hong, J. D. Makromolekulare Chemie, Macromolecular Symposia 1991,46, 321. (252) Decher, G.; Schmitt, J. Progress in Colloid & Polymer Science 1992, 89, 160. (253) Bucur, C. B.; Sui, Z.; Schlenoff, J. B. Journal of the American Chemical Society 2006, 128, 13690. (254) Schlenoff, J. B. Multilayer Thin Films: Sequential Assembly of Nanocomposite Materials; Wiley-VCH: Weinheim, 2003. (255) Jeon, J.; Panchagnula, V.; Pan, J .; Dobrynin, A. V. Langmuir 2006, 22, 4629. (256) Chen, J .; Cao, W. Chemical Communications 1999, 1711. (257) Decher, G. Science 1997, 277, 1232. (258) Schlenoff, J. B. Ly, H.; Li. M. Journal of the American Chemical Society 1998, 120, 7626. (259) Joly, 8.; Kane, R.; Radzilowski, L.; Wang, T.; Wu, A.; Cohen, R. E.; Thomas, E. L.; Rubner, M. F. Langmuir 2000, I6, 1354. (260) Lowack, K.; Helm, C. A. Macromolecules 1998, 31, 823. (261) Finke, R. G. In Metal Nanoparticles. Synthesis, Characterization and Applications; F eldheim, D. L.; Foss, C. A.; Eds., Marcel Dekker: New York, 2000, pp 17. 262) Neff, P. A.; Wunderlich, B. K.; Klitzing, R. V.; Bausch, A. R. Langmuir 2007, 23, 4048. (263) Dubas, S. T.; Schlenoff, J. B. Macromolecules 1999, 32, 8153. (264) Yoo, D.; Shiratori, S. S.; Rubner, M. F. Macromolecules 1998, 31, 4309. (265) Harris, J. J .; Bruening, M. L. Langmuir 2000, I6, 2006. (266) Caruso, F.; Niikura, K.; Furlong, D. N.; Okahata. Y. Langmuir 1997, 13, 3422. (267) Caruso, F. Advanced Materials 2001, I3, 11. (268) Dai, J.; Bruening, M. L. Nano Letters 2002, 2, 497. 38 (269) Keller, S. W.; Yonemoto, E. H.; Brigham, E. S.; Saupe, G. B.; Mallouk T. E. Advances in Chemistry Series 1998, 254, 359. (270) Laurent, D.;Schlenoff, J. B. Langmuir 1997, 13, 1552. (271) Wang, D.; Rogach, A. L.; Caruso, F. Nano Letters 2002, 2, 857. (272) Lowman, G. M.; Nelson, S. L.; Graves, S. M.; Strouse, G. F.; Buratto, S. K. Langmuir 2004, 20, 2057. (273) Komarala, V. K.; Bradley, A. L.; Byme, S. J.; Corr, S. A.; Gun'ko, Y. K. Nanotechnology 2006, 1 7, 4117. (274) F redin, N. J .; Zhang, J.; Lynn D. M. Langmuir 2007, 23, 2273. (275) Jewell, C. M.; Zhang, J.; Fredin, N. J.; Wolff, M. R.; Hacker, T. A.; Lynn, D. M. Biomacromolecules 2006, 7, 2483. (276) Ren, K.; Shen, J. Biomaterials 2006, 2 7, 1152. (277) Yoo, P. J.; Nam, K. J.; Qi, J.; Lee, S. -K.; Park, J.; Belcher, A. M.; Hammond, P. T. Nature Materials 2006, 5, 234. (278) Fischlechner, M.; Toellner, L.; Messner, P.; Grabherr, R.; Donath, E. Angewandte Chemie International Edition 2006, 45, 784. (279) F eldheim, D. L.; Grabar, K. C.; Natan, M. J.; Mallouk, T. E. Journal of the American Chemical Society 1996, 118, 7640. (280) Keller, S. W.; Johnson, S. A.; Brigham, E. S.; Yonemoto, E. H.; Mallouk, T. E. Journal of the American Chemical Society 1995, 11 7, 12879. (281) Leguen, E.; Chassepot, A.; Decher, G.; Schaaf, P.; Voegel, J .-C.; Jessel, N. Biomolecular Engineering 2007, 24, 33. (282) Haynie, D. T.; Zhao, W.; Rudra, J. S. Nanomedicine 2006, 2, 150. (283) Kovacevic, D.; Peran, N. Colloids and Surfaces A: Physicochemical and Engineering Aspects 2006, 2 7 7, 177. (284) Salloum, D. S.; Schlenoff, J. B. Biomacromolecules 2004, 5, 1089. (285) Caruso, F.; Schuler, C. Langmuir 2000, I6, 9595. (286) Lvov, Y.: Caruso, F. Analytical Chemistry 2001, 73, 4212. 39 CHAPTER 2 SELECTIVE HYDROGENATION OF MONOSUBSTITUTED ALKENES BY Pd NANOPARTICLES EMBEDDED IN POLYELECTROLYTE FILMS 2.1. Introduction As shown in chapter 1, metal NPs are attractive catalytic materials because of their high surface area to volume ratio and their size-dependent, tunable electronic 6 properties." However, aggregation of metal NPs reduces surface area and restricts control over particle size. Thus, practical catalysts that rely on these materials generally ll-15 l6-l8 19-38 contain a support, such as carbon,7"° metal oxides, zeolites, or polymers, onto which the NPs are deposited to prevent particle aggregation and facilitate catalyst recovery. In particular, polymeric supports are attractive because they provide unique possibilities for modifying the environment around catalytic sites and controlling access to these sites. Incorporation of NPs in multilayer polyelectrolyte film523’24 or dendrimersfs‘26 for example, allows size-based selectivity in the hydrogenation of olefins. This chapter demonstrates that catalytic NPs in multilayer polyelectrolyte films also show remarkable intramolecular selectivity in the hydrogenation of alkenes based on the degree of substitution of the double bond. Selectivity in the hydrogenation of monosubstituted double bonds over di or trisubstituted double bonds is. equal to or greater than that of Wilkinson’s catalyst, tris(triphenylphosphine)rhodium(I) chloride, which is the prototype homogenous catalyst for such reactions.39 Additionally, alumina-supported NP/polyelectrolyte catalysts can be easily separated from the reaction mixture and recycled. Several reports show that the layer-by-layer deposition of polyelectrolyte films . . . . . . . . 4.4 ‘4 provrdes a convenient method for syntheSls or immoblllzatlon of nanopartrcles.23’2 ’ 0 4 40 In the present case, we prepare selective catalysts using layer-by-layer adsorption of poly(acrylic acid) (PAA) and poly(ethylenimine)—Pd(11) (PEI-Pd(II)) complexes on alumina particles and subsequent reduction of Pd(II) by NaBI-Ia to form NPs as illustrated schematically in scheme 2.1.23'24 The resulting embedded NPs contain catalytic sites that show unusually high selectivities, partly because the polymer restricts access to catalytic sites. . ._ J Repeat step 1-4 3 11:11:; 3) PE] ,‘ 36mg followed J a Alumina ’ [PAA-Pd(ID/PEIbPAA 3) NaBH4 4) Rinse -—————. [PAA-Pd(oyPE113PAA Scheme 2.1. [PAA/PEl-Pd(0)]3PAA film formation on Alumina. Our previous research reported that the size-based selectivity of these catalysts in hydrogenation of allyl alcohols containing different alkyl substituents in the position a to the double bond?”4 Smaller molecules exhibited higher hydrogenation rates, presumably because of more rapid transport to catalytic sites. The work reported in this chapter shows that [PAA/PEI-Pd(0)]3PAA films on alumina also exhibit high selectivities based on the substitution on the double bond itself. Moreover, intramolecular selectivity demonstrates that specificity cannot be explained solely by differences in the rate of diffusion through the polyelectrolyte film to catalytic sites. 41 2.2. Experimental 2.2.1. Materials Poly(ethylenimine) (PEI, MW = 25 000 Da), poly(acrylic acid) (PAA, Mw = 90 000 Da), palladium (5 wt % on alumina powder), a-alumina (100 mesh, typical particle size 75-100 um), allyl alcohol (99%), crotyl alcohol (97%, mixture of isomers), 2- methyl-2-propen-l-ol (98%), 3-methy-2-buten-1-ol (99%), propionaldehyde, 3-methyl-1- butanol (98.5+%), isovaleraldehyde (97%), 2-heptene-4-ol, 1,5-heptadien—4-ol, 3,7- dimethyl-l,6-octadien-3-ol (97%), potassium tetrachloropalladate(II) (99.99%), and sodium borohydride (98%) were purchased fi'om Aldrich and used as received. l-Butanol (J .T. Baker), n-propyl alcohol (Columbus Chemical), iso-butyl alcohol (Spectrum), ethanol (100%, Pharmco), and calcium hydride were also used as received. Benzene was purchased from EMD chemicals and distilled over calcium hydride. Aqueous solutions were prepared with deionized water (Milli-Q, 18.2 M!) cm), and hydrogen (99.9%) was obtained from AGA gases. 2.2.2. Preparation of Pd NPs Encapsulated in Polyelectrolyte Multilayer Films Preparation of [PAA/PEI-Pd(0)]3PAA films on alumina began with vigorous stirring of 15 g of til-alumina in 100 mL of 20 mM PAA (pH adjusted to 4.0 with 0.1 M NaOH, molarity of PAA is given with respect to the repeating unit) for 10 min. Subsequently, the alumina was allowed to settle, the supernatant was decanted, and the alumina was washed with three 100-mL aliquots of deionized water (stirring in water for 5 min each time) to remove excess PAA. The polycation layer was deposited by vigorous stirring of PAA-coated alumina in a 100-mL solution containing the PEI-Pd(II) complex (1 mg/mL PEI, 2 mM KdeCla, pH adjusted to 9.0 with 0.1 M HCl), followed 42 by washing with three lOO-mL aliquots of deionized water. This procedure was repeated until 3.5 bilayers were deposited on alumina. Reduction of Pd(II) was performed by stirring the [PAA/PEI-Pd(II)]3PAA films on alumina in 100 mL of freshly prepared 0.1 mM NaBH4 for 30 min, followed by washing with three 100-mL aliquots of deionized water. The water was decanted, and the residual water was removed under reduced pressure to afford a catalytic powder. 2.2.3. TEM of Pd NPs For TEM imaging, films were deposited on carbon-coated copper grids that were treated in a UV/ozone cleaner for 1 min. [PAA/PEI-Pd(0)]3PAA films were deposited as described above, except the grid was simply immersed for 5 min in solutions without stirring, and rinsing between each step consisted of a 1-min washing with deionized water fi'om a squirt bottle. Imaging was performed on a JEOL lOOCX microscope using an accelerating voltage of 100 kV. Particle sizes were determined manually on digital images. 2.2.4. Determination of the Amount of Pd in Catalyst and Deposition Solutions Calculation of turnover frequencies (TOFs), i.e., the moles of substrate that are hydrogenated per mole of Pd per hour, requires knowing the amount of Pd in the catalyst. To determine Pd content, Pd in the catalyst was dissolved using aqua regia, and atomic emission spectroscopy (Varian Spectra AA-200) was employed to analyze these solutions as described previously.23‘24 43 2.2.5. Hydrogenation Reactions Catalytic hydrogenation was performed in a 100-mL, three-neck, round-bottomed flask. During the reaction, hydrogen at a gauge pressure of 50 kPa was bubbled through a frit at the bottom of a solution that was vigorously stirred. Prior to the reaction, suspended catalyst (250 mg of [PAA/PEI-Pd(0)]3PAA on alumina or 10 mg of commercial 5% Pd on alumina) in 25 mL of H20 was bubbled with H2 for 30 min before adding 25 mL of a 50.0 mM solution of substrate in water or in a 1:1 (v:v) mixture of ethanol and water. Much less mass of the commercial catalyst was used for hydrogenation reactions because this material has a 36-fold higher Pd loading than [PAA/PEI-Pd(0)]3PAA films on alumina. (The amount of Pd in 10 mg commercial catalyst is about 1.42 times higher than that in 250 mg of [PAA/PEI-Pd(0)]3PAA on alumina.) Gas chromatography (Shimadzu GC-17A equipped with an RTx- BAClcolumn) was employed to monitor reactions using response factors that were determined with a standard mixture containing 10 mM reactant, product, and isomer (if applicable), except for the case of 2-methyl-2-propen-1-ol, for which we did not have pure isomer. In this case only, the isomer response factor was assumed to be the same as the product, as these compounds have the same number of C and O atoms. IH NMR spectra obtained on 300 MHz or 500 MHz spectrometers and '3 C NMR spectra obtained on 75 MHz or 125 MHz spectrometers were plotted with chemical shift reported relative to the residue peaks of methanol (5 3.30 for 1H NMR and 49.0 for 13C NMR). For molecules containing multiple double bonds, GC-MS (Hewlett Packard G1800B GCD system with a non-polar HP-SMS column) with appropriate response factors was used to determine TOF. TOFs were calculated from the slope of the linear portions of plots of percent hydrogenation versus time (conversion less than 50 %) with the intercept of the 44 line forced through zero. Each experiment was repeated twice and the i values represent the difference between the average and the data points. Representative calculation of response factor and TOF are shown in appendix. 2.2.6. Recycling Experiment Catalytic hydrogenation was performed in 100% water as described above. Most of the solvent was decanted after the reaction, the catalyst was transferred to another flask, and the residual water was removed in vacuum for more than 4 hours prior to determination of the remaining catalyst mass. About 5 % of the catalyst was lost during the recycling process. The hydrogenation reaction was repeated with this catalyst prior to another recycling step and a third hydrogenation reaction. 2.2.7. Hydrogenation Reactions by Wilkinson’s Catalyst Catalytic hydrogenation was performed in a lOO-mL oven dried three-neck, round-bottomed flask containing a condenser and a bubbler. N2 was bubbled through the round-bottom flask for 10 min before the reaction. During the reaction, hydrogen at a gauge pressure of 50 kPa was bubbled through a frit at the bottom of a solution that was vigorously stirred with magnetic stir bar. Prior to the reaction, 23 mg (0.025 mole) of RhCl(PPh3)3, Wilkinson’s catalyst, was stirred in 18 mL of benzene with bubbling of H2 for 20 min before adding 2.5 mmole substrate in 2 mL of benzene using a glass syringe. Sample was collected using a glass syringe at various time intervals that depended on the rate of the reaction. Gas chromatography was employed to monitor the reaction using response factors as described above. 45 2.2.8. Isolation of Products After 97% hydrogenation of compounds 5 (0.140 g, 50 mM solution in 25 mL 1:1 water and ethanol, later mixed with 25 mL water) and 7 (0.193 g, 50 mM solution in 25 mL 1:] water and ethanol, later mixed with 25 mL water), the catalyst was removed from the solution by filtration through cotton, and about lg of NaCl was added to the solution. Products (6 and 8) were extracted with 25 mL of anhydrous diethyl ether. The organic layer was collected and dried over anhydrous sodium sulfate, and solvent was removed under reduced pressure to afford crude compounds 6 and 8. Pure compound 8 was afforded in 81% yield (0.158 g) after column chromatography with 9:1 hexane and ethyl acetate. In the case of hydrogenation of compound 5, 0.120 g of crude product (84% yield) was isolated after extraction. GC/MS analysis showed that the ratio of 6 and 9 in crude 6 is 8:1. 1H NMR indicated the presence of small amounts of impurities in the crude product as shown in the spectra, which are given in the appendix. Compound 6: lH NMR (500 MHz, CD30D): 8 0.90 (t, 3H), 1.27-1.51 (m, 4H), 1.67 (dd, 3H), 3.95 (m, 1H), 5.41 (m, 1H), 5.61 (m, 1H), 13C NMR (125 MHz, CD3OD): 5 14.3, 17.8, 19.7, 40.5, 73.4, 127.1 and 135.4. Spectral data were consistent with commercially available material. Compound 8: 1H NMR (300 MHz, CD30D): 8 0.88 (t, 3H), 1.11 (s, 3H), 1.37-1.51 (m, 4H), 1.61 (s, 3H), 1.66 (s, 3H) 2.00 (m, 2H), 5.11 (m, 1H), 13C NMR (75 MHz, CD3OD): 5 8.5, 17.6, 23.6, 25.9, 26.2, 34.9, 42.2, 73.4, 125.8 and 132.0 46 2.2.9. Transport Studies Diffusion dialysis experiments were performed using a home-built dialysis apparatus (shown in Figure 2.1) in which a porous alumina membrane (Whatman Anodisc, 0.02 um-diameter surface Stirrers Alumina — Membrane 1; Receiving Source Phase O-ring Phase 0 Membrane Figure 2.1. Home-built diffusion dialysis set-up for transport experiments. (Figure courtesy of Dr. Jinhua Dai). pores) coated with a [PAA/PEI]7 film was sandwiched between two glass cells. (More layers were used in these experiments than in preparing catalysts to ensure complete coverage of the pores in the underlying alumina support.) The source cell was filled with 90 mL of a 20 mM aqueous solution of the desired unsaturated alcohol (allyl alcohol, crotyl alcohol, 2-methy1-2-propen—1-ol, or 3-methyl-2-buten-1-ol), while the receiving cell initially contained 90 mL of deionized water. Both the source and the receiving sides were stirred vigorously to minimize concentration polarization at the membrane 47 interface, and samples of the receiving phase were taken every 20 min and analyzed by gas chromatography using methanol as an internal standard. 2.3. Results and Discussions 2.3.1. NP Size Figure 2.2 & 2.3 shows the TEM image of a [PAA/PEI-Pd(0)]3PAA film on a carbon-coated grid. Consistent with prior results,23 the distribution of particle diameters ranged from 1-3 mm as shown in Figure 2.4. The estimated average particles size is 2.2 nm. The relatively impermeable PAA/PEI films likely inhibit ion and NP diffusion to give small particles. Figure 2.2. TEM image of [PAA/PEUPd(0)]3PAA on carbon-coated cupper grid. 48 Figure 2.3. TEM image of a [PAA/PEI-Pd(0)]3PAA film on a carbon-coated copper grid. This image was taken at a lower magnification than Figure 2.2 in to better show the distribution of NPs. 30. % of particles 1.1 1.3 1.5 1.7 1.9 2.1 2.3 2.5 2.7 2.9 3.1 3.3 Diameter (nm) Figure 2.4. Histogram of Pd NP diameters in several TEM images of [PAA/PEI- Pd(0)]3PAA film deposited on a TEM grid. The average particle size is 2.2 nm. 49 2.3.2. Intermolecular Selectivity Using similar films on alumina particles, we initially compared the rates of hydrogenation of unsaturated alcohols with different substituents on the double bond and found very high selectivities (Table 2.1). With a [PAA/PEI-Pd(0)]3PAA-on-alumina catalyst in water, allyl alcohol (1) could be 82% hydrogenated in 15 min (formation of about 18% propionaldehyde isomer prohibits complete hydrogenation) with an initial turnover frequency (TOF) of 3500 moles hydrogenated/mole catalyst/h.45 In contrast, 2- methyl-2-propen-1-ol (2), crotyl alcohol (3), and 3-methyl-2-buten-l-ol (4) exhibited TOFs that ranged from 14 to 100-fold lower (Table 2.1, column 4). By comparison, a commercially available 5% Pd on alumina catalyst facilitated hydrogenation of 1, 2, 3 and 4 with TOFs that varied by at most a factor of 2 when water was the solvent (Table 2.1, column 2). 50 Table 2.1. TOFs for the hydrogenation of a series of unsaturated alcohols by different Pd catalysts. TOF (moles hydrogenated/ mol Pd/ h)a 5%Pd on 5%Pd on [PAA/PEI- [PAA/PEI- Substrates Alumiréa Alumina Pd(O)]3P€xA Pd(O)]3PAA (H2O) (H2O/EtOH)c (H2O) (H2O/EtOH)c 1 =¥ 1680i30 1160a120 3505s175 2700i100 0H (2770120) (1700i170) (53001500) (5590i50) 2 =< 835d:65 415245 120i5 75a5 OH (2275a25) (1270a130) (365:1:35) (230i20) 3 /=\_OH 1705215 1060i50 250115 175.215 4 >=\_ OH 1220a190 ssoaso 35a5 25a5 [a] Values in parentheses represent TOFs where the isomer is included as part of the product, although it is not hydrogenated. Each experiment was repeated twice, and the i values represent the difference between the average and the data points. [b] 25 mM substrate was hydrogenated in water. [c] 25 mM substrate was hydrogenated in a 3:1 (v:v) mixture of water and ethanol. Interestingly, although the TOP for hydrogenation of allyl alcohol using the [PAA/PEI-Pd(0)]3PAA on alumina catalyst is higher than that of the commercially available 5% Pd on alumina, selectivity is still much higher for the Pd NPs in the polyelectrolyte multilayer. However, the commercial catalyst has a 36-fold higher wt % of Pd than [PAA/PEI-Pd-(0)]3PAA on alumina, as determined by atomic emission spectroscopy. To keep the amount of Pd in the different reactions about the same, 250 mg of [PAA/PEI-Pd(0)]3PAA on alumina was added to the reaction mixtures while only 10 mg of 5 wt % Pd on alumina was employed. Overall, the molar ratio of substrate to 51 Pd atoms was 400 for hydrogenation using [PAA/ PEI-Pd(0)]3PAA on alumina and 280 for hydrogenation using the commercial 5% Pd on alumina. Hydrogenation was also performed in 3:1 mixtures of water and ethanol because of the relatively low solubility of the substituted unsaturated alcohols in water. For the [PAA/PEI—Pd(0)]3PAA on alumina catalyst, the TOF for allyl alcohol decreased by 23% in the less polar mixture, possibly because the reduced swelling of the polyelectrolyte films decreases the rate of substrate diffusion to the catalytic sites. However, TOFs also decreased for the commercial catalyst, suggesting that other factors influence the hydrogenation rate. Nevertheless, the observed selectivities are similar in the two solvent systems (compare columns 4 and 5 of Table 2.1), reaching 2 orders of magnitude (1/4) for the [PAA/PEI-Pd(0)]3PAA on alumina catalyst. As a control experiment, hydrogenation of allyl alcohol and its isomer propionaldehyde was attempted in the absence of hydrogen gas using [PAA/PEI- Pd(0)]3PAA as a catalyst. No reaction occurred, which rules out the possibility of other reductants, such as trace amounts of residual NaBH4, driving the reaction. The propionaldehyde and 2-methylpropionaldehyde isomers of 1 and 2, respectively, are not hydrogenated under the reaction conditions, so their formation results in a decreased rate of hydrogenation. Isomer formation was particularly extensive for 2 (ratio of the isomer to product was 1.5-2 with 50% of the reactant remaining) and also occurred for allyl alcohol (1) when using either the commercial 5% Pd on alumina or the [PAA/PEI-Pd(0)]3PAA catalyst. Hence the TOFs of these molecules are somewhat lower than what might be expected, especially in the case of 2 (row 2, Table 2.1). The sum of TOFs for isomerization and hydrogenation are given in parentheses in Table 2.1 for cases where significant isomer formation occurred. 52 To examine the recylability of the [PAA/PEI-Pd(0)]3PAA on alumina, this catalyst was recovered after the hydrogenation of allyl alcohol in water and dried prior to use in a subsequent hydrogenation. Remarkably, the TOF was 3200i100 moles hydrogenated/mol Pd/h for three consecutive hydrogenations of allyl alcohol. Thus, unlike homogeneous catalysts such as tris-(triphenylphosphine)rhodium(I) chloride, [PAA/PEI-Pd(0)]3PAA on alumina can be readily recovered and reused. The selectivity of the embedded Pd NPs compares very well with that of Wilkinson’s catalyst, which previously showed a rate of hydrogenation for l-hexene (monosubstituted unsaturated alcohols) in benzene that is 1.1 and 49 times higher than that for 2-methylpent-1-ene (a disubstituted unsaturated alcohol) and l- methylcyclohexene methylcyclohexene (a trisubstituted unsaturated alcohol),46 respectively. To better compare the catalytic selectivity of [PAA/PEI-Pd(0)]3PAA films with that of Wilkinson’s catalyst, we determined the TOFs for the hydrogenation of compounds 1, 2, and 4 with Wilkinson’s catalyst (Table 2.2). (In this case, the reactions had to be performed in benzene rather than water because Wilkinson’s catalyst is not soluble in aqueous solutions.) The selectivities of Wilkinson’s catalyst for 1/2 and 1/4 were 2.6 and 54, respectively, which is in reasonable agreement with the data mentioned above for l-hexene, 2-methylpent-1-ene, and l-methylcyclohexene. (The GC peak corresponding to l-butanol, the hydrogenated product of 3, overlapped with that of benzene, so hydrogenation of this compound with Wilkinson’s catalyst was not examined.) Wilkinson’s catalyst shows only minor selectivity for mono versus disubstituted unsaturated alcohols, whereas the embedded NPs show selectivities ranging from 14 to 36 for allyl alcohol over 2 and 3. In the case of the trisubstituted unsaturated alcohol 4, the selectivity (1/4) of the embedded NPs is 2-fold higher than that of the 53 Wilkinson’s catalyst. However, unlike the [PAA/PEI-Pd(0)]3PAA films, Wilkinson’s catalyst showed minimal formation of the isomer of 1 and none of the isomer of 2. In contrast, Wilkinson’s catalyst showed formation of the isomer of 4 (the amounts of isomer and hydrogenated product were the same although rates of formation of both were low), while the [PAA/PEI-Pd(0)]3PAA film did not show any readily detectable isomer. Table 2.2. TOFs for the hydrogenation of a series of unsaturated alcohols by Wilkinson’s catalyst. SUbStTateS TOF (moles hydrogenated/ mol of catalyst/ h)al 1. _ 270i10 —\—0H (290i10) 2. : 105i25 OH 4. 5 >=\_ (10) OH [a] Values .in parentheses represent the sum of TOFs for the hydrogenation and isomerlzatlon. 2.3.3. Intramolecular Selectivity We further examined the catalytic behavior of [PAA/PEI-Pd(0)]3PAA films on alumina for intramolecularly selective hydrogenation of molecules containing a combination of mono and di or tli substituted double bond (Table 2.3). In this case, 1,5- heptadien-4-ol (5) and 3,7-dimethyl-1,6-octadien-3-ol (7) were selectively hydrogenated into 2-hepten-4-ol (6) and 3,7-dimethyl-6-octen—3-ol (8), respectively, when hydrogenation was catalyzed by [PAA/PEI-Pd(0)]3PAA. In a control experiment, the TOF for the hydrogenation of 2-hepten—4-ol (6) to 4-heptanol (9) was 3% of the TOF for 54 reduction of 5 to 6, indicating much slower conversion for the disubstituted double bond. This is consistent with the fact that at 97% conversion of 5, the isolated product was an 8:1 mixture of 6 and 9 (84% overall yield). (Reactions are first order in substrate 50 product ratios cannot be calculated as simple ratios of TOFs.) No product with both hydrogenated double bonds was observed by GC-MS and NMR for compound 7 (81% yield of 8 after purification) because trisubstituted double bond are hydrogenated very slowly (Table 2.1). The TOF for the hydrogenation of 5 to 6 is 2.6 times that for hydrogenation of 7 to 8, probably because the more bulky groups 0. to the monosubstituted carbon-carbon double bond in 7 restrict access to catalytic sites and diminish the rate of the reaction. A5 a consequence, compounds 5 and 7 required 15 and 135 min, respectively, to achieve about 97% conversion to hydrogenated products. Control experiments with the 5% Pd on alumina also showed some intramolecular selectivity in the hydrogenation of 5 and 7, but much less than that for the [PAA/PEI-Pd(0)]3PAA catalyst. At about 99% hydrogenation of 5 by the commercial catalyst, the product distribution was 34% 6 and 65% 9. The TOF for the hydrogenation of compound 6 (550i130) was also lO-fold greater with the commercial catalyst than with [PAA/PEI-Pd(0)]3PAA. In the case of hydrogenation of 7 using the commercial catalyst, after 240 min, 95% of the mixture was the doubly hydrogenated product, but 25 min after starting, the mixture contained 94% 8 after only 25 min of reaction. In this case, the commercial catalyst may also be a viable choice for somewhat selective hydrogenation, but one would need to carefully monitor the reaction. 55 Table 2.3. TOFs (moles hydrogenated/ mole Pd/ h) for selective hydrogenation of monosubstituted double bonds in the presence of di and tri substituted double bonds in the same molecules. The reaction was catalyzed by [PAA/PEI—Pd(0)]3PAA on alumina in a 3:1 mixture of water and ethanol. Substrates Major Products TOFsa H H 5 /\j\/\ 6 Ai/\ 1835i” / / / 7 W 3 W 700i140 HO HO H /\j:/\ 6 Ai/\ 9 / SSiS [a] The reaction was catalyzed by [PAA/PEI-Pd(0)]3PAA on alumina in a 3:1 mixture of water and ethanol. TOF is expressed as moles/hydrogenated/mol Pd/h. Our prior studies of diffusion of allyl alcohol derivatives through porous alumina membranes (0.02-um surface pore size) coated with [PAA/PEI]7 films suggested that different rates of diffusion to catalytic sites imparted selectivity.23’24 Figure 2.5 presents analogous dialysis studies with compounds 1-4 and also shows selective transport through [PAA/PEI]7 films on alumina. The ratios of transport rates (slopes in Figure 2.5) for 1/2, 1/3, and 1/4 were 2.2, 2.0 and 6.3, respectively. However, relative rates of hydrogenation in water are 25, 12 and 82 for 1/2, 1/3, and 1/4, respectively, suggesting that differences in transport rates are insufficient to account for the hill hydrogenation selectivity among these compounds. Moreover, selective hydrogenation within molecules 5 and 7 shows that these compounds can diffuse through the film to reach catalytic sites and yet only one double bond is predominantly hydrogenated. Diffusion through the film cannot be limiting hydrogenation of the second double bond. On the 56 other hand, compounds 1-4 are all hydrogenated at relatively similar rates (1/2=2.1, 1/3=0.95, and 1/4=1.2) with the commercial catalyst (5% Pd on Alumina) suggesting that selectivity cannot have a completely electronic basis (Table 2.1, column 2). We think that steric constraints for binding of multiply substituted double bonds to embedded NPs must be significantly greater than similar constraints in binding to the 5% Pd on alumina. I investigate the selectivity in the hydrogenation of unsaturated alcohols as a function of the NP size in chapter 3. 57 0.08 - 0.07 - '8 0.05 i 0.04 - 0.03 - 0.02 - / / 0.01 . / 0 T“ 1 l I I 0 20 40 60 80 Time (min) Receiving-phase concentration (mM) Figure 2.5. Time evolution of the receiving phase concentrations of allylic alcohols during diffusion dialysis of 1 (squares), 2 (triangles), 3 (diamonds), and 4 (circles) through porous alumina membranes coated with a [PAA/PEI]7 film. The source phase initially contained 20 mM of the allylic alcohols, and the receiving phase was initially deionized water. 2.4. Conclusion Deposition of Pd (II) in polyelectrolyte films followed by reduction yields catalytic, NP-containing films that can be used as catalysts in inter and intramolecularly selective hydrogenation of unsaturated alcohols. Remarkably, the selectivity of these heterogeneous catalysts exceeds that of Wilkinson’s homogenous catalyst. Intramolecular selectivity and studies of diffusion through PAA/PEI films suggest that selectivity stems from steric constraints for binding of multiply substituted double bonds to embedded NPs. 58 2.5. Appendix Calculation of TOFs From GC Data With Response Factors Determination of Response Factors sP cp*k, sR = cR SI C|*k2 8.. z C—R Where Sp = Peak area corresponding to product S1 = Peak area corresponding to isomer SR = Peak area corresponding to Reactant Cp = Concentration of product CR = Concentration of reactant C1 = Concentration of isomer k1 = Response factor for product to reactant k2 = Response factor for isomer to reactant Response factors were calculated from the signals due to a solution containing 10 mM product, 10 mM reactant, and 10 mM isomer, and three chromatograms with three standard solutions were averaged to determine the value. 59 Calculation of % Hydrogenation 1 mL samples were removed every few minutes (time intervals depended on the rate) from the hydrogenation mixture and placed in autosampler vials for GC analysis. Percent hydrogenation was determined as shown below. C 11: % hydrogenation = P x 100% = SP k2 x 100% CP+CR+C' SP*k2 + SRikkfl'kz + S1*k1 When the rates of hydrogenation and isomerization were added % reaction = % of hydrogenation + % of isomerization: C + C S *k + S*k %reaction= P ' x100% = P 2 ' ‘ x100% CP+CR+C| Sp*k2 + SR*k1*k2 + S|*k1 Table A-1 shows some sample data. Table 2. A-l Time-dependent percentage of crotyl alcohol hydrogenation catalyzed by [PAA/PEI-Pd(0)]3PAA in water. Time (min) % of hydrogenation 0.0 0 10.0 7.93 20.0 24.3 30.0 33.89 40.0 45.3 6O Crotyl Alcohol (Realetle‘nt) Butanol A (Desired Product) r r 5.5 5.6 5.7 5.3 1 Minutes Figure 2. A-l Example gas chromatograms of the reaction mixture from hydrogenation of crotyl alcohol (3) after 10 min (black line) and 40 min (red line) of reaction time. The reaction was catalyzed by [PAA/PEI-Pd(0)]3PAA in water. Integration was performed using instrument software. Time Vs % of Hydrogenation 50 — 1: y =1.1313x ° 40 — ‘8 0 31 30 ~ 2. 20 .. Q- 0 10 _ .\° 0 ' l l T j 0 10 20 3O 40 50 Time (min) Figure 2. A-2. Plot of time-dependent percentage of crotyl alcohol hydrogenation catalyzed by [PAA/PEI-Pd(0)]3PAA in water. 61 r T—T pen U1 ! + Q C) 1 9° l 31' i 7 1 6 j r— n 5 :1 HO 4 3 2/ *.___1 1 *,____1 —n JMJ Figure 2. A-3 . 1H NMR spectrum of isolated crude 2-heptene-4-ol (compound 6) in CD30D. Impurity peaks are indicated with a *. 62 IIIIII pen 40 IIIIIIIIIIIIIIII‘IIII IIIIIII 60 171 l 80 100 120 III]IIITIFIIIIITTTTFIIIIFIll 1 1‘0 lllrnfifijfilr 160 1111 180 rTI—lfirfTT I: Figure 2. A-4 13C NMR spectrum of isolated crude 2-heptene-4-ol (compound 6) in CD3OD. Impurity peaks are indicated with a *. 63 l i ,,r 1 I ,1 c: L £29“. fl 4 *4; : ‘\\ a; 9 __\_, a; w 10 ‘ j g. \ a," o 6 \ Cited" {‘4 8 9 , i i 1n ‘0 \ 7 e; I O 010’ 1 CD300 Figure 2. A-5. 1H NMR spectrum of purified 3,7-dimethyl-6-octen-3-ol (compound 8) in CD3OD. Impurity peaks are indicated with a * 64 Figure 2. A-6. 13C NMR spectrum of purified 3,7-dimethyl-6-octen-3-ol (compound 8, chapter 2) in CD3OD. Impurity peaks are indicated with a * 65 2.6. References (l) Schmid, G. In Nanoscale Materials in Chemistry; Klabunde, K. J ., Ed., Wiley Interscience: New York, 2001; pp. 15. (2) Chen, S.; Ingram, R. S.; Hostetler, M. J .; Pietron, J. J.; Murray, R. W.; Schaaff, T. G.; Khoury, J. T.; Alvarez, M. M.; Whetten, R. L. Science 1998, 280, 2098. (3) Li, Y.; Boone, E.; El-Sayed, M. A. Langmuir 2002, I8, 4921. (4) Rao, C. N. R.; Kulkami, G. U.; Thomas, P. J.; Edwards, P. P. Chemistry-A European Journal 2002, 8, 28. (5) Fendler J. H. Nanoparticles and Nanostructuredfilms: Preparation, Characterization and Applications; Wiley-VCH: Weinhein, Germany, 1998. (6) Wilson, 0. M.; Knecht, M. R.; Garcia-Martinez, J. C.; Crooks, R. M. Journal of the American Chemical Society 2006, 128, 4510. (7) Liu, Z.; Ling, X. Y.; Su, X.; Lee, J. Y. Journal of Physical Chemistry B 2004, 108, 8234. (8) Narayanan, R.; El-Sayed, M. A. Journal of Catalysis 2005, 234, 348. (9) Pronkin, S. N.; Simonov, P. A.; Zaikovskii, V. I.; Savinova, E. R. Journal of Molecular Catalysis A: Chemical 2007, 265, 141. (10) Park, K. H.; Jang, K.; Kim, H. J .; Son, S. U. Angewandte Chemie International Edition 2007, 46, 1152. (11) Valden, M.; Lai, X.; Goodman, D. W. Science 1998, 281, 1647. (12) Corma, A.; Sema, P. Science 2006, 313, 332. (13) Come, A.; Sema, P.; Garcia, H. Journal of the American Chemical Society 2007, 129, 6358. (14) Zheng, N.; Stucky, G. D. Journal of the American Chemical Society 2006, 128, 14278. (15) Mallick, K.; Scurrell, M. S. Applied Catalysis A: General 2003, 253, 527. (16) Sun, C.; Peltre, M.-J.; Briend, M.; Blanchard, J.; Fajerwerg, K.; Krafft, J.-M.; Breysse, M.; Cattenot, M.; Lacroix, M. Applied Catalysis A: General 2003, 245, 245. (17) Manda], S.; Roy, D.; Chaudhari, R. V.; Sastry, M. Chemistry of Materials 2004, I6, 3714. 66 (18) Chen, J.-H.; Lin, J.-N.; Kang, Y.-M.; Yu, W.-Y.; Kuo, C.-N.; Wan, B.-Z. Applied Catalysis A: General 2005, 291, 162. (19) Pathak, S.; Greci, M. T.; Kwong, R. C.; Mercado, K.; Prakash, G. K. S.; Olah, G. A.; Thompson, M. E. Chemistry of Materials 2000, 12, 1985. (20) Mei. Y.; Sharma, G.; Lu, Y.; Ballauff, M.; Drechsler, M.; lrrgang, T.; Kempe, R. Langmuir 2005, 21 , 12229. (21) Mei. Y.; Lu, Y.; Polzer, F.; Ballauff, M.; Drechsler, M. Chemistry of Materials 2007, 19, 1062. (22) Sharma, G.; Mei, Y.; Lu, Y.; Ballauff, M.; lrrgang, T.; Proch, S.; Kempe, R. Journal of Catalysis 2007, 246, 10. (23) Kidambi, S.; Dai, J.; Li, J.; Bruening, M. L. Journal of the American Chemical Society 2004, 126, 2658. (24) Kidambi, S.; Bruening, M. L. Chemistry of Materials 2005, I 7, 301. (25) Niu, Y.; Yeung, L. K.; Crooks, R. M. Journal of the American Chemical Society 2001, 123, 6840. (26) Oh, S.-K.; Niu, Y.; Crooks, R. M. Langmuir 2005, 21 , 10209. (27) Garcia-Martinez, J. C.; Lezutekong, R.; Crooks, R. M. Journal of the American Chemical Society 2005, 12 7, 5097. (28) Scott, R. W. J.; Wilson, 0. M.; Oh, S.-K.; Kenik, E. A.; Crooks, R. M. Journal of the American Chemical Society 2004, 126, 15583. (29) J iang, Y.; Gao, Q. Journal of the American Chemical Society 2006, 128, 716. (30) Pittelkow, M.; Moth-Poulsen, K.; Boas, U.; Christensen, J. B. Langmuir 2003, I9, 7682. (31) Oyamada, H.; Akiyama, R.; Hagio, H.; Naito, T.; Kobayashi, S. Chemical Communications 2006, 4297. (32) Kralik, M.; Biffis, A. Journal of Molecular Catalysis A: Chemical 2001, 1 7 7, 113. (33) Nakao, R.; Rhee, H.; Uozumi, Y. Organic Letters 2005, 7, 163. (34) Park, C. M.; Kwon, M. S.; Park, J. Synthesis 2006, 3790. (35) Liu, Y.; Khemtong, C.; Hu, J. Chemical Communications 2004, 398. 67 (36) Hou, Z.; Theyssen, N.; Brinkmann, A.; Leitner, W. Angewandte Chemie International Edition 2005, 44, 1346. (37) Yan, W.; Wang, R.; Xu, Z.; Xu, J .; Lin, L.; Shen, Z.; Zhou, Y. Journal of Molecular Catalysis A: Chemical 2006, 255, 81. (38) Adlim, M.; Bakar, M. A.; Liew, K. Y.; Ismail, J. Journal of Molecular Catalysis A: Chemical 2004, 212, 141. (39) Osborn, J. A.; Jardine, F. H.; Young, J. F .; Wilkinson, G. Journal Of Chemical SocietyA 1966, 1711. (40) Schmitt, J .; Decher, G.; Dressick, W. J .; Brandow, S. L.; Geer, R. E.; Shashidhar, R.; Calvert, J. M. Advanced Materials 1997, 9, 61. (41) Mamedov, A. A.; Belov, A.; Giersig, M.; Mamedova, N. N.; Kotov, N. A. Journal of the American Chemical Society 2001, 123, 7738. (42) Wang, T. C.; Rubner, M. F .; Cohen, R. E. Chemistry of Materials 2003, 15, 299. (43) Joly, 8.; Kane, R.; Radzilowski, L.; Wang, T.; Wu, A.; Cohen, R. E.; Thomas, E. L.; Rubner, M. F. Langmuir 2000, 16, 1354. (44) Wang, T. C.; Rubner, M. F .; Cohen, R. E. Langmuir 2002, I8, 3370. (45) The TOF reported for hydrogenation of allyl alcohol by [PAA/PEI-Pd(0)]3PAA on alumina is 3.9-fold higher than what we reported previously. The higher value is due in large part to a 2.5-fold higher concentration of allyl alcohol. (46) Takaya, H.; Noyori, R.; In Comprehensive Organic Synthesis: Selectivity, Strategy and Efliciency in Modern Organic Chemistry; Trost, B. M.; Fleming, 1. Eds, Pergamon, New York, 1991, Vol 8 pp. 443. 68 CHAPTER 3 SELECTIVITY AS A FUNCTION OF NANOPARTICLE SIZE IN THE CATALYTIC HYDROGENATION OF UNSATURATED ALCOHOLS 3.1. Introduction Control over particle size is essential in creating highly active NP catalysts for alkene hydrogenation,"5 alcohol oxidationf’7 Suzuki8 and Heck9 coupling, Norrish type 0 II reactions,l reduction of aromatic nitro compounds,ll CO oxidation,”’13 and dl4,15 I.l6 electrooxidation of formic aci and methane NP size is also a critical parameter for achieving partial hydrogenation in the conversion of alkynes to alkenes,”18 01,13 unsaturated aldehydes to unsaturated alcohols,'9 and conjugated alkenes to monoalkenes.2°’2' Moreover, NP catalysts show high intermolecular selectivities in the hydrogenation of alkenes22 and allylic alcohols,”27 but selectivity has not been studied as a function of particle size. Examination of selectivities is particularly attractive for understanding the effect of particle size on reactivity because rate enhancements due solely to increases in surface area should cancel when calculating the ratios of reaction rates for structurally related compounds. This chapter shows that both the rates and selectivities of hydrogenation reactions vary dramatically with the diameter of catalytic NPs. Remarkably, in the hydrogenation of mono and disubstituted double bonds, NPs with average diameters of 2.2 nm exhibit selectivities as high as 39 for allyl alcohol over crotyl alcohol, whereas NPs with average diameters of 3.4 nm show corresponding selectivities less than 1.3. Such variations in selectivity occur with polymer-encapsulated NPs immobilized on alumina as well as polymer-capped NPs in solution. 69 The first step in examining the effect of NP size on catalytic activity is development of a method to synthesize NP catalysts with several different average particle sizes and narrow size distributions. Among the methods for controlling NP size, a number of studies demonstrate that layer-by-layer deposition of polyelectrolyte films containing metal ions and subsequent reduction of these metal ions provide a straightforward technique for creating encapsulated NP5.23'25’28'32 The amount of metal ion present during film deposition dictates the particle size, and layer-by-layer deposition on micron—sized alumina provides heterogeneous catalysts. Polyelectrolyte-NP catalysts can also form in solution, which allows comparison of particle-size effects in supported films and solution-based catalysts. Palladium NPs in polyelectrolyte films are particularly attractive for investigating relationships between particle size and reactivity/selectivity because they show high selectivity in the hydrogenation of unsaturated alcohols.23'24 In the previous chapter we described the preparation of Pd NPs encapsulated in multilayer polyelectrolyte films through layer-by-layer adsorption of poly(acrylic acid) (PAA) and poly(ethylenimine)- Pd(II) (PEI-Pd(II)) complexes and reduction of Pd(II) with NaBH4. Transmission electron microscopy confirmed the formation of Pd NPs, and 'the selectivities of [PAA/PEI-Pd(0)]3PAA films in the hydrogenation of monosubstituted over multisubstituted double bonds were as high as 100.23 However, we employed only one average particle size. This work examines the hydrogenation of a series of unsaturated alcohols using NP/polyelectrolyte catalysts with several average NP sizes and shows that selectivity is an extremely sensitive function of particle diameter. 70 3.2. Experimental Section 3.2.1. Materials Poly(ethylenimine) (PEI, MW = 25 000 Da), poly(acrylic acid) (PAA, Mw = 90 000 Da), ct-alumina (100 mesh, typical particle size 75-100 um), allyl alcohol (99.0%), crotyl alcohol (97.0%, mixture of isomers), 2-methyl-2-propen—l-ol (98.0%), 3-methyl-2- buten-l-ol (99.0%), propionaldehyde, 3-methyl-1-butanol (98.5+%), isovaleraldehyde (97.0%), isobutyraldehyde (98.0%), 3-methyl-1-penten-3-ol (99.0%), 3-methyl-3- pentanol (99.0%), cis-3-hepten-l-ol (2 95.0%) potassium tetrachloropalladate(11) (99.99%), and sodium borohydride (98%) were purchased from Aldrich. l-Heptanol and 3-buten-l-ol were obtained from TCI, and butyraldehyde (2 99.0%) was purchased from Fluka. l-Butanol (J.T. Baker), n-propyl alcohol (Columbus Chemical), iso-butyl alcohol (Spectrum) and ethanol (100%, Pharmco) were used as received. Aqueous solutions were prepared with deionized water (18.2 MQ cm, Milli-Q purification system), and hydrogen (99.9%) was obtained from AGA gases. 3.2.2. Preparation of Pd NPs Encapsulated in Polyelectrolytes 3.2.2.1. [PAA-Pd(0)/PEI]3 on Alumina Preparation of PAA/PEI-Pd(0) films on alumina was described previously,”24 and formation of PAA-Pd(0)/PEI films occurred similarly. Alumina was alternately suspended for 10 min in a solution that contained 20 mM PAA and 1, 4, 8, or 15 mM K2PdCl4 and for 10 min in a solution that contained PEI (1 mg/mL). (The molarity of PAA is given with respect to the repeating unit.) Prior to their use in multilayer assembly, the PAA solution was adjusted to pH 4.0 with 0.1 M NaOH, and the PEI 71 solution was adjusted to 9.0 with 0.1 M HCl. After the deposition of each polyelectrolyte layer, the alumina was allowed to settle, the supernatant was decanted, and the solid was suspended similarly (vigorous stirring) in three lOO-mL aliquots of deionized water for 5 min to remove excess polyelectrolyte. Overall, three PAA-Pd(II)/PEI bilayers were deposited. Reduction to Pd(0) was performed by stirring the [PAA-Pd(II)/PEI]3 films on alumina in 100 mL of freshly prepared 0.1 mM NaBHa for 30 min, followed by washing with three 100-mL aliquots of deionized water for 5 min each. After decanting the final water rinse, the residual water was removed under reduced pressure overnight to afford a catalytic powder. 3.2.2.2. [Pd(0)/PEI] on Alumina [Pd(0)/PEI]3 films on alumina were prepared as described previously.25 In brief, 15 g of el-alumina powder was stirred for 30 min in 100 mL of a solution that contained K2PdC14 (1 or 15 mM) and 0.1 M KCl. The alumina was washed as described above prior to suspension in PEI (2 mg/mL PEI, pH adjusted to 9.0 with 0.1 M HCl) for 10 min and additional washing. After deposition of three PdC142'/PEI bilayers, the reduction of Pd(II) to Pd (0) was performed as described above. 3.2.2.3. PAA-Pd(0)/PEI in Solution PAA-Pd(II) was prepared by mixing 20 mM PAA and K2PdCl4 (l, 4, 8, or 15 mM) and adjusting the pH of the solution to 4 with 0.1 M NaOH. Five mL of this solution and 5 mL of PEI solution (1 mg/mL PEI, pH adjusted to 9.0 with 0.1 M HCl) were mixed together in a 20 mL glass vial at which point the solution became cloudy. Freshly prepared 0.1 M NaBH4 (5 mL) was added to the cloudy PAA-Pd(II)/PEI, which 72 was then stirred vigorously for 3 h in a loosely capped via]. The resulting clear, brown solution was then immediately used for hydrogenation. Crooks et al. previously reported that dendrimer-encapsulated Pd NPs, a 3 We also homogeneous catalytic system, are unstable in the presence of oxygen.3 observed that the brown color of the PAA-Pd(0)/PEI solution faded to pale yellow over 24 h if the solution was not capped tightly and refiigerated. This was especially apparent when the concentration of Pd(II) in the deposition solution was 1 mM. 3.2.3. Characterization of Pd NPs For TEM imaging, films were deposited on carbon-coated copper grids that were pretreated in a UV/ozone cleaner for 1 min. [PAA-Pd(O)]/PEI]3 and reduced [PdClaz' /PEI]3 films were formed as described above, except the grid was simply immersed for 5 min in deposition solutions without stirring, and rinsing between each step consisted of a 1-min washing with deionized water. To characterize the PAA-Pd(0)/PEI solutions prepared with different initial concentrations of Pd(II) (1, 4, 8, 15 mM), we first diluted the reduced solutions with sufficient deionized water to obtain 0.33 mM Pd. We then spotted a drop of these solutions on separate carbon-coated copper grids that had been pretreated with UV/ozone for 1 min, and allowed the solutions to dry at room temperature overnight. Imaging was performed on a JEOL-2200FS microscope using an accelerating voltage of 200 kV. Diameters of at least 250 particles for each sample were measured manually on digital images, and uncertainties in average particle diameters represent standard deviations. The amounts of palladium in different catalysts were determined by atomic emission spectroscopy (AES). Standard solutions (0.1 to 0.5 mM) were prepared by dissolving 73 K2PdC14 in 0.1 M HNO3, and sample solutions were prepared by stirring the desired amount of alumina-supported catalyst (60 to 500 mg, depending on the Pd content) in 2 mL of aqua regia for 15 min. The aqua regia solutions were diluted to 12 mL with water and centrifuged (the (ll-alumina support does not dissolve in aqua regia), and the supernatant was analyzed using its emission at 340.5 nm. Table 3.1 lists the amounts of Pd in each of the different catalysts. For the PAA-Pd(0)/PEI solutions, the amount of Pd(0) was calculated by assuming total conversion of Pd(Il) to Pd(O). 74 Table 3.]. Amounts and Pd Content of Catalysts Used in Hydrogenation Reactions. Catalyst Amount of Moles of Pd in Catalyst the Catalyst [PAA-Pd(0)/PEI]3 011 A1203 500 mg 1.91 (Pd(II) = 1 mM), Catalyst A [PAA-Pd(0)/PEI]3 on A1203 250 mg 4.15 (Pd(II) = 4 mM), Catalyst B [PAA-Pd(0)/PEI]3 on A1203 125 mg 7.66 (Pd(II)= 8 mM), Catalyst C [PAA-Pd(0)/ PEI]3 on A1203 125 mg 12.8 (Pd(II) = 15 mM), Catalyst D [PAA-Pd(O)/PEI]3(PAA/PEI)3 on A1203 500 mg 1,13 (Pd(II) = 1 mM), Catalyst E [PAA-Pd(O)/PEI]3 (PAA/PEI)3on A1203 125 mg 112 (Pd(II) = 15 mM), Catalyst F Reduced [PdCl42'/PEI]3 on A1203 250 mg 3,71 (Pd(II) = 1 mM), Catalyst G Reduced [PdCl42'/PEI]3 on A1203 100 mg 110 (Pd(II)= 15 mM), Catalyst H PAA-Pd(0)/PEI solution 500 mL 1.67 (Pd(II)= 1 mM), Catalyst I PAA-Pd(0)/PEI solution 5_00 mL 6.65 (Pd(II) = 4 mM), Catalyst J PAA-Pd(0)/PEI solution 300 mL 8.00 (Pd(II)= 8 mM), Catalyst K PAA-Pd(O)/PEI solution 300 mL 1500 (Pd(II) = 15 mM), Catalyst L 75 3.2.4. Hydrogenation Reactions Catalytic hydrogenation was performed in a 100-mL, three—neck, round-bottomed flask. Suspended alumina-supported catalyst in 25 mL of H20 (or 25 mL of diluted catalyst solution in the case of NPs stabilized by polyelectrolytes in solution) was bubbled with H2 for 30 min before adding 25 mL of a 50 mM solution of substrate in water or in a 1:1 (v:v) mixture of ethanol or methanol or THF and water. During the reaction, hydrogen at a gauge pressure of 50 kPa was bubbled through a frit at the bottom of a solution that was vigorously stirred. Different amounts of catalyst (Table 3.1) were used for hydrogenation reactions because each type of material had a different wt% of Pd. 0.5-mL aliquots were removed from the reaction vessel at specific times, filtered with cotton to remove the catalyst, and diluted by addition of l-mL deionized water prior to analysis by gas chromatography (Shimadzu GC-17A equipped with an RTx- BAClcolumn). In the GC analysis of hydrogenation of 3, product and reactant peaks showed a small overlap, and one limit of integration was placed at the minimum between the peaks. Response factors were determined to calculate the turnover frequency (TOF) as described previously.23 TOFs were calculated from the slope of the initial portion (total conversion of reactant to isomer and hydrogenated product <50%) of plots of percent hydrogenation versus time. In fitting the initial data, the intercept of the line was forced through zero. Each experiment was repeated twice and the i values represent the difference between the average and the data points. 76 3.3. Result and Discussion 3.3.1. Catalyst Synthesis and Variation of Particles Size The initial step in studying the effect of NP size on catalytic selectivity is to prepare a series of catalysts with different particle sizes. Previously, we made [PAA/PEI- Pd(II)]3PAA films using 2 mM Pd(II) in the PEI-Pd(II) deposition solution. Subsequent reduction of the Pd(II) gave [PAA/PEI-Pd(0)]3PAA coatings containing Pd NPs with diameters of 1-3 nm.23 We tried to enhance NP size by increasing the concentration of Pd(II) in the PEI deposition solution, but at a Pd(II) concentration of 12 mM, the PEI solution became cloudy, perhaps because the Pd(II) serves as a cross- linking agent for PEI chains. To overcome this challenge, we add Pd(II) to the PAA, rather than PEI, deposition solution. The 20 mM PAA deposition solution (PAA concentration is given with respect to the repeating unit) remains clear at Pd(II) concentration as high as 15 mM. Prior studies demonstrated deposition of metal ions in PAA/protonated poly(allylamine) films and subsequent formation of nanoparticles,3"32 but in the present case we add the metal ions during film formation. Most importantly, reduction of [PAA-Pd(II)/PEI]3 using NaBH4 yields [PAA- Pd(0)/PEI]3 films, and the sizes of the resulting Pd NPs increase with increasing concentrations of Pd(II) in PAA deposition solutions. Figures 3.1, 3.2, 3.3, 3.4, and 3.5 present TEM images of films on carbon-coated grids and histograms of particle sizes. 77 N N O 0| 1 1 1n 2 .2 E 15- 8 g 10- n. 0 .\° 5' o 1.2 1.8 2.4 3 Nanoparticle Diameter (nm) Figure 3.1. (a) TEM image of a [PAA-Pd(0)/PEI]3 film on a carbon-coated copper grid. 1 mM Pd(II) was present in the PAA-K2PdCl4 solution used for film deposition. A few of the NPs are circled. (b) Histogram of Pd NP diameters in several TEM images. The average particle size is 2.2iO.6 nm. The area-average particle size is 2.5:l:0.7 nm. 78 30- 251 20. 15. 1o- 5. o % of Nanoparticles 1.8 2.1 2.4 2.7 3.0 3.3 3.6 3.9 4.2 Nanoparticle Diameter (nm) Figure 3.2. (a) TEM image of a [PAA—Pd(0)/PEI]3 film on a carbon-coated copper grid. 4 mM Pd(II) was present in the PAA-K2PdCl4 solution used for film deposition. A few NPs are circled. (b) Histogram of Pd NP diameters in several TEM images. The average particle size is 3.03:0.8 nm. The area-average particle size is 3.1:k0.5 nm. 79 % of Nanoparticles 2.1 2.7 3.3 3.9 4.5 Nanoparticle Diameter (nm) Figure 3.3. (a) TEM image of a [PAA-Pd(O)/PEI]3 film on a carbon-coated copper grid. 8 mM Pd(II) was present in the PAA-K2PdC14 solution used for film deposition. (b) Histogram of Pd NP diameters in several TEM images. The average particle size is 3.2i0.6 mn. The area-average particle size is 3.5i0.7 nm. 80 M (II I O' .3 A 0 0| 1 1 % of Nanoparticles UI O 2.4 3 3.6 4.2 4.8 5.4 6 Nanoparticle Diameter (nm) .5 . on Figure 3.4. (a) TEM image of a [PAA-Pd(0)/PEI]3 film on a carbon-coated copper grid. 15 mM Pd(II) was present in the PAA-K2PdCl4 solution used for film deposition. (b) Histogram of Pd NP diameters in several TEM images. The average particle size is 3.4:l:1.3 nm. The area-average particle size is 3.8:l:1.1 nm. 81 With Pd(II) concentrations of 1, 4, 8, or 15 mM in the PAA deposition solutions, the average particle diameters in [PAA-Pd(0)/PEI]3 films on TEM grids are 2.2iO.6, 2.9i0.5, 3.2i0.6, and 3.4:t0.9 nm, respectively. Area-average particle sizes might be more representative of catalytic activities because larger particles make up a disproportionate share of the total catalyst surface area. We calculate the area-average diameters, DA, using equation 1, where D, represents the diameters of individual particles. ZD i D i2 i 1' DA (1) Values of DA are 2.5:t0.7, 3.1i0.5, 3.53:0.7, and 3.8i1.1 nm for films prepared using 1, 4, 8, and 15 mM Pd(II), respectively, in the 20 mM PAA deposition solutions. Occasionally, TEM images of [PAA-Pd(0)/PEI]3 films prepared with 15 mM Pd(II) in the PAA deposition solutions exhibit areas containing large particles with average diameters of 160i100 nm (Figures 3.5). We did not include these larger particles in the calculation of average particle diameters because they constitute only about 18 % of the total particle surface area (Table 3.2), and we are not sure whether they are representative of all samples. However, the absolute values of TOFs for catalysis with alumina-supported [PAA-Pd(0)/PEI]3 films that are prepared using 15 mM Pd(II) should be viewed with caution. 82 Figure 3.5. Low-magnification TEM images of large particles and NPs in [PAA- Pd(0)/PEI]3 films on carbon-coated copper grids. The fihns were prepared with 15 mM Pd(II) in the PAA deposition solution. (a) Image showing large particles with diameters of 160i100 nm, as determined from several TEM images. Such large particles appeared in ~30% of similar images. (b) Image of NPs. Larger spots are overlapping multiple particles. The dark spot in the center of low-magnification images is due to a defect in the lens where it was burned by the electron beam. 83 Table 3.2. Calculation of the relative surface area of large (ranging from ~50-300 nm) particles and NPs (< 5 nm) in [PAA-Pd(O)/PEI]3 films and reduced [PdC142'/PEI]3 films on carbon-coated TEM grids. The initial concentration of Pd(II) was 15 mM in the solution containing PAA-PdClaz‘ or PdC142'. Large particles were not prevalent when using lower concentrations of Pd(II). % occupy by big particles and NPs Specific Descriptions [PAA- [Pdle' Pd(0)/PEI]3 rpm], Average percentage of image area covered by large particles ~ 0 ~ 0 (when present) 11 /° 1 1 /o Percentage of low-magnification images containing large particles ~30% ~6O% Percentage of image area covered by large particles in all low-magnification images (including images with and ~3 % ~7 % without large particles) Average percentage of image area covered by NPs ~ 0 ~ 0 (determined from high magnification images) 14 /o 14 A: Fraction of total Pd surface area due to small NPs (determined from the image areas covered by large and small 0.82 0.67 NPs)a [a] Because both the surface area of a particle and the image area it covers are proportional to the square of the particle diameter, the fraction of the total Pd surface area due to NPs is equivalent to the image area covered by NPs divided by the sum of the image areas covered by NPs (< 5 nm) and larger particles (ranging from ~ 50-300 nm). Although we use films deposited on alumina particles for catalysis, we perform the TEM analyses with films on carbon-coated Cu grids because it is very difficult to visualize NPs on the rough alumina microparticles. The NPs likely have similar diameters on the TEM grids and on the alumina, but we cannot verify the diameters of particles on alumina. To address this problem, we also prepare polyelectrolyte/NP composites in water. These materials serve as catalysts and are amenable to TEM 84 imaging, which allows a direct correlation between catalytic activity and particle size. A recent study showed that in many cases, NPs in water do not aggregate in the presence of polyelectrolytes.34 The preparation of the suspended polyelectrolyte/NP composites includes the addition of an aqueous solution of PEI (1 mg/ml, pH adjusted to 9) to a solution of PAA (20 mM, pH adjusted to 4) that contains various concentrations of K2PdC14 (1, 4, 8, or 15 mM). The yellow-colored PAA-Pd(II) solution becomes cloudy after addition of PEI and then turns to a clear brown color after addition of 0.1 M aqueous NaBHa and vigorous stirring. The pH of the PAA-Pd(II)/PEI mixture is about 5, but the addition of 0.1 M NaBHa gives a clear solution with a pH of ~9. Raising the pH of PAA-Pd(II)/PEI solutions by addition of 0.1 M NaOH also yields a clear solution, presumably because less protonation of PEI results in weaker PAA/PEI complexes. However, both polyelectrolytes are important for stabilizing the NPs. Addition of 0.1 M NaBHa to solutions containing either PAA and 8 mM Pd(lI) or PEI and 8 mM Pd(II) produces a black precipitate. To determine the sizes of NPs suspended in these solutions, we deposit a drop of solution on a carbon-coated copper grid and then image the NPs with TEM. These TEM images along with histograms of particle sizes (Figures 3.6, 3.7, 3.8, 3.9) indicate the formation of NPs whose size is a function of the concentration of Pd(II) in solution prior to reduction. In this case, the average particle diameters that result from the use of 1, 4, 8, and 15 mM Pd(II) to form NPs are 2.6i0.8, 32:20.7, 3.5i0.7 and 5.1i1.3 nm, respectively. Corresponding area-average diameters are 3.1:t0.8, 3.4i0.7, 3.7:l:0.7 and 5.8i1.4 nm. We do not observe any large (>9 nm) particles in PAA-Pd(0)/PEI solutions 85 prepared with Pd(II) concentrations of 15 mM in the PAA deposition solutions (Figure 3.10). % of Nanoparticles 1.2 1.8 2.4 3 3.6 4.2 4.8 Nanoparticle Diameter (nm) Figure 3.6. (a) TEM image of a dried PAA-Pd(0)/PEI solution on a carbon-coated copper grid. 1 mM Pd(II) was initially present in the PAA solution. (b) Histogram of Pd NP diameters in several TEM images. The average particle size is 2.6108 nm. The area- average particle size is 3.1i0.8 nm. 86 N 0| ;1 % of Nanoparticles ..n .1. N 0 0| 0 (II D 1 r l 1 A . on 2.4 3 3.6 4.2 4.8 Nanoparticle Diameter (nm) Figure 3.7. (a) TEM image of a [PAA-Pd(0)/PEI] solution on a carbon-coated copper grid. 4 mM Pd(II) was initially present in the PAA-PdCl42' solution. (b) Histogram of Pd NP diameters in several TEM images. The average particle size is 3.2iO.7 nm. The area- average particle size is 3.4i0.7 um. 87 N O a 0| 0| % of Nanoparticles 3 O 1.8 2.4 3 3.6 4.2 4.8 Nanoparticle Diameter (nm) Figure 3.8. (a) TEM image of a dried PAA-Pd(0)/PEI solution on a carbon-coated copper grid. 8 mM Pd(II) was initially present in the PAA solution. (b) Histogram of Pd NP diameters in several TEM images. The average particle size is 3.5:l:0.7 nm. The area- average particle size is 3.7:l:0.7 um. 88 14 - 12 - 10 - 3 . 5 . 4 . 2 . o . 2.4 3 3.6 4.2 4.8 5.4 6 6.6 7.2 7.8 Nanoparticle Diameter (nm) %of Nanoparticles Figure 3.9. (a) TEM image of a [PAA-Pd(0)/PEI] solution on a carbon-coated copper grid. 15 mM Pd(II) was initially present in the PAA-PdClaz' solution. (b) Histogram of Pd NP diameters in several TEM images. The average particle size is 5.1:l:1.3 nm. The area-average particle size is 5.8:l:1.4 nm. 89 Figure.3.10. (a) TEM image of a dried PAA-Pd(0)/PEI solution on a carbon-coated copper grid. 15 mM Pd(II) was initially present in the PAA solution. This image was taken at a relatively low magnification to show the distribution of NPs and the absence of large particles on the TEM grid. The dark spot in the center of low-magnification images is due to a defect in the lens where it was burned by the electron beam. 3.3.2. Selective Hydrogenation with Films on Alumina Particles To investigate catalytic selectivity as a fimction of NP diameter, we first investigated hydrogenation of allyl alcohol (1, a monosubstituted unsaturated alcohol), 2- methyl-Z-propen-l-ol (2, a disubstituted unsaturated alcohol), and crotyl alcohol (3, a disubstituted unsaturated alcohol). TOFs (moles hydrogenated/mol Pd/h) at low substrate conversion provide a quantitative comparison of the reactivities and selectivities of the different catalysts. Unforttmately, interpretation of the TOFs is complicated by isomerization of the unsaturated alcohols to the corresponding aldehydes, which are not readily hydrogenated. Thus, Table 3.3 contains TOFs for initial hydrogenation along 90 with values in parentheses that reflect the sum of the initial hydrogenation and initial isomerization TOFs. Isomerization is most extensive for 2, presumably because its isomer is a trisubstituted unsaturated alcohol that is more stable than 2 and readily tautomerizes to the corresponding aldehyde. Table 3.3 shows that regardless of NP size, both the TOF for hydrogenation and the combined hydrogenation/isomerization TOF are higher for 1 than for 2 and 3, which contain disubstituted double bonds. In addition, selectivity varies dramatically with the particle size. The general trend in Table 3.3 is that the rate of hydrogenation of 1 (column 3) decreases with increasing NP size, while the corresponding TOFs for 2 (column 4) and 3 (column 5) increase with increasing NP size, even though the surface area to volume ratio for the catalyst is smaller for larger particles. Most remarkably, this trend yields selectivities (ratios of TOFs for different compounds, columns 6 and 7 of Table 3.3) that are as much as 90—fold larger with the smallest NPs than with the largest. These trends also hold when considering the sum of hydrogenation and isomerization. The selectivities for 1/2 and 1/3 with the smaller NPs are especially notable because Wilkinson’s catalyst, the prototypical homogeneous catalyst for selective hydrogenation, shows essentially no selectivity in the hydrogenation of mono- and disubstituted double bonds.23’35 91 Table 3.3. TOFs and TOF ratios for unsaturated-alcohol hydrogenation catalyzed by [PAA-Pd(O)/PEI]3 films that contain Pd NPs with different diameters. The catalytic films were deposited on micron-sized alumina. Values in parentheses are the sums of TOFs for hydrogenation and isomerization. Average TOFs (moles hydrogenated/ mol Pd/ h)a Ratio 0f TOPS nano- particle diameter _ Catalyst (nm) _\_OH OH /=\_OH 1/2 1/3 1 2 3 CatalystA b 2 2 2900i570 12d:1 753:15 240 39 [pd(11)]=1 mM ' (6200i2400) (85i5) (85i15) (73) (73) Catalyst B b 2.9 1700i830 160i10 450i80 11 3.8 [Pd(11)]=4 mM (4100a 800) (730i20) (6202.430) (5.6) (6.6) CatalystC b 3 2 1100.600 165:1:5 610i60 6.7 1.8 [pd(11)]=3 mM ' (2400:}:800) (895i15) (760i80) (2.7) (3.2) Catalyst D b 3 4 760i40 290i70 5801100 2.6 1.3 [pd(11)]=15 mM ' (2050i80) (1400i100) (790:1:160) (1.5) (2.6) [a] 50 mL of 25 mM substrate was hydrogenated in water. TOFs are initial values determined at conversions less than 50%. [b] The initial concentration of KdeCl4 in the PAA deposition solutions. Several studies show that polyelectrolyte multilayers provide a significant barrier to the diffusion of small compounds, and even molecules as small as glycerol and 36 Thus, one methanol show much slower transport than water in polyelectrolyte films. possible explanation for the high selectivities of NPs in polyelectrolyte films is that the polyelectrolytes decrease the rate at which bulkier molecules reach catalytic sites. To examine whether simple diffirsion through polyelectrolyte films enhances selectivity, we coated films containing the smallest and largest NPs with additional layers of polyelectrolytes. Pd was reduced prior to the deposition of additional layers. As comparison of Table 3.3 and Table 3.4 shows, addition of three PAA/PEI bilayers on the 92 Table 3.4. TOFs and TOF ratios for unsaturated-alcohol hydrogenation catalyzed by [PAA—Pd(O)/PEI]3/(PAA/PEI)3 films that contain Pd NPs with different diameters. The films were deposited on micron-sized alumina. Values in parentheses are the sums of TOFs for hydrogenation and isomerization. Average TOFs (moles hydrogenated/ mol Pd/ h)a Ratio 0f TOFS nano- Catalyst particle _ — diameter —\—OH 0H /—\—0H (m) 1 2 3 1/2 1/3 Catalyst E 2.2 1500i250 10+.1 35:5 150 43 [Pd(11)]=1 me (3000i1000) (40i5) (404.5) (75) (75) Catalyst F 3.4 1200i90 85i5 40015 14 3.0 [pd(11)]=15 me (2600i600) (3702215) (540i35) (7.0) (4.8) [a] 25 mM substrate was hydrogenated in water. TOFs are initial values determined at conversions less than 50%. [b] The initial concentration of KdeCl4 in the PAA deposition solutions. top of a film containing 2.2 nm-diameter particles does not have a large effect on the relative TOFs of 1 and 2 or 1 and 3, although the magnitudes of TOF values decrease by a factor of ~2. Addition of three PAA/PEI bilayers to films containing NPs with a diameter of 3.4 nm increases selectivity as much as 6-fold, but the selectivity of these catalysts is still ll-(1/2) and lS-fold (1/3) lower than that with the 2.2-um NPs (Table 3.4). These data suggest that the most important factor behind selectivity is NP size. To further demonstrate that NP size is the primary factor behind selectivity, we studied catalysis with reduced [PdCl42’/PEI]3 coatings on alumina. (The nomenclature PdCl42' reflects deposition conditions and not the state of Pd in the films). These NP- containing films should be more permeable than [PAA-Pd(O)/PEI] coatings because they contain only one polyelectrolyte.25 After reduction of the Pd(II) in these films, the average particle size is 3.43:0.8 nm when using 15 mM PdCl42' to prepare the films and 93 % of Nanoparticles 1.5 2.1 2.7 3.3 3.9 4.5 Nanoparticle Diameter (nm) Figure 3.11. (a) TEM image of a [Pd(0)/PEI]3 film on a carbon-coated copper grid. The concentration of PdCl42' used in depositing the precursor [PdC142'lPEI]3 film was 1 mM. (b) Histogram of Pd NP diameters in several TEM images. The average particle size is 3.0:t0.9 nm. The area-average particle size is 3.5i1.0 nm. 94 25- 20- b 15- 10- 5.. % of Nanoparticles 0 1 .8 2.4 3 3.6 4.2 4.8 5.4 Nanoparticle Diameter (nm) Figure 3.12. (a) TEM image of a [Pd(0)/PEI]3 film on a carbon-coated copper grid. The concentration of PdCl42' used in depositing the precursor [PdCl42‘/PEI]3 films was 15 mM. (b) Histogram of Pd NP diameters in several TEM images. The average particle size is 3.4i0.8 nm. The area-average particle size is 3.8i0.9 nm. 95 Figure 3.13. (a) Low-magnification TEM image of a reduced [PdCl42’lPEI]3 film on a carbon-coated copper grid. 15 mM PdCl42' was initially present in the deposition solution. The average size of these large particles in several TEM images is 170i110 nm. Large particles are visible in about 60% of these low-magnification images, but we estimate that the surface area of the large particles constitutes ~33% of the surface area in the film (Table 3.2). 3.0:t0.9 run when using a 1 mM PdCl42' solution (See Figures 3.11 and 3.12). Corresponding area-average diameters are 3.8i0.9 and 3.5:hl.0 nm respectively. In this case, NP sizes are not very different because PdCl42‘ does not bind to a polymer in solution before adsorbing to the film, but rather adsorbs directly to the polyelectrolyte film. Because the amount of adsorbed PdCl42' in Pdle/PEI films is somewhat self- limiting as the film becomes negatively charged, PdChz’ concentration should not have a dramatic effect on the amount of Pd in the film. However, in some areas of reduced 96 [PdCl42'/PEI]3 films prepared with 15 mM Pd(II), TEM images show the presence of large particles (average particle size 170i110 nm) as well as NPs (Figure 3.13). We estimate that the large particles represent ~33 % of the total Pd surface area. We did not include these larger particles in the calculation of average particle diameters because we are not sure whether they are representative of all samples. The absolute values of TOFs for catalysis with reduced [PdCl42'/PEI]3 films that are prepared using 15 mM Pd(II) should, thus, be viewed with caution. Table 3.5 summarizes the TOFs for the hydrogenation of 1, 2 and 3 using the two different reduced [PdCl42'/PEI] films on alumina. The smaller NPs in catalyst G give selectivities that are an order of magnitude higher than hydrogenation using catalyst H, even though the average NP diameters in the two catalysts differ by only 0.40 nm. Table 3.5. TOFs and TOF ratios for unsaturated-alcohol hydrogenation catalyzed by reduced [PdCl42'/PEI]3 films that contain Pd NPs with different diameters. The films were deposited on micron-sized alumina. Values in parentheses are the sums of TOFs for hydrogenation and isomerization. TOFs (moles hydrogenated! mol Pd/ h)a Ratio 0f TOPS Avarage nano- _ _ Catalyst particle —\—OH {OH /—\—OH diameter 1 2 3 1/ 2 1/3 (11m) Catalyst G 3.0 2230i70 25d:5 220i40 89 10 [pd(11)]=1 me (4600i38 (150:1:40) (250i40) (30) (19) 0) Catalyst H 3.4 880:1:110 380i110 580i80 2.3 1.5 [pd(11)]=15 me (2300i10 (1700d:350) (7501100) (1.4) (3.1) 0) [a] 25 mM substrate was hydrogenated in water. TOFs are initial values determined at conversions less than 50%. [b] The initial concentration of KdeCl4 in the deposition solutions. 9 97 3.3.3. Hydrogenation with NP/Polyelectrolyte Complexes in Solution Because we cannot effectively image NP catalysts on alumina, we also examined hydrogenation with suspended PAA-Pd(0)/PEI catalysts prepared by addition of a PEI (1 mg/mL) solution to 20 mM PAA solutions containing different concentrations of Pd(II). Reduction of the Pd(II) with NaBH.; yields suspended PAA-Pd(0)/PEI catalysts with a variety of Pd NP sizes. One concern when using these NPs is that excess NaBH4, which is not removed from the NP solution, might promote side reactions. As a control experiment, we attempted reduction of allyl alcohol in the absence of hydrogen gas using the PAA-Pd(0)/PEI solution containing excess Nth. No reaction occurs, which rules out the possibility of NaBH4 participating in side reactions. Similar to catalysis by NP films on alumina, with the PAA-Pd(0)/PEI solution, the TOF for catalytic hydrogenation of 1 decreases with increasing particle size, whereas the TOFs for 2 and 3 either increase or remain approximately constant as particle size increases (Table 3.6). Thus, overall selectivity shows an order of magnitude decrease on going from an average NP diameter of 2.6 nm to 5.1 nm. Comparison of Tables 3.3, 3.4, 3.5 and 3.6 shows that selectivities are frequently higher for these suspended NPs than they are for [PAA-Pd(0)/PEI]3 and reduced [PdCl42'/PEI]3 films on alumina. The suspended NPs in solution may present a more uniform set of catalytic sites than the films on alumina powder, which might favor high selectivity. In general, supported catalysts are complex materials due to the presence of many types of potential active sites. 98 Table 3.6. TOFs and TOF ratios for unsaturated-alcohol hydrogenation catalyzed by solutions of PAA-Pd(0)/PEI that contain Pd NPs with different diameters. Values in parentheses are the sum of TOFs for hydrogenation and isomerization. TOFs (moles hydrogenated/ mol Pd/ h)a Ratio 0f TOFS Average nano- — _ Catalys‘ particle —1¥°” 0” /_3\—°” 1/2 1/3 diameter (nm) Catalyst 1 37001860 12 30110 310 120 [Pd(II)]=l me 2.6 (580011900) (7515) (35115) (73) (170) Catalyst J 16001190 12 4515 130 35 [Pd(II)]=4 me 3.2 (29001600) (5015) (50110) (59) (59) Catalyst K 11001560 1214 3515 95 33 [Pd(II)]=8 me 3.5 (24001950) (6015) (4515) (40) (54) Catalyst L 8901210 30110 80115 30 10 [Pd(II)]=15 me 5-1 (17001350) (100115) (100115) (16) (18) [a] 25 mM substrate was hydrogenated in water. TOFs are initial values determined at conversions less than 50%. [b] The initial concentration of KdeCl4 in the PAA deposition solutions. 3.3.4. Solvent Effects on Hydrogenation Previous studies of catalytic hydrogenation showed that variations in solvent can lead to changes in reaction rates, differences in the extent of double-bond isomerization, or changes in selectivity.23’24’27’37'38 Presumably, molecules of solvents such as MeOH, EtOH and THF adsorb to the catalyst to alter reactivities. Additionally, swelling of polyelectrolyte films in water may increase reaction rates.39 We examined hydrogenation with catalysts A and D using solvents consisting of 3 parts water and 1 part of either MeOH or EtOH or THF. These two catalysts represent the widest range of particle diameters among the NPs deposited on alumina. Table 3.7 summarizes the TOFs for hydrogenation of l, 2, and 3 in different mixed solvents. For catalyst D (average particle diameter of 3.4 nm), the rate of hydrogenation for all of the compounds decreases in the 99 Table 3.7. TOFs and TOF ratios for unsaturated-alcohol hydrogenation catalyzed by [PAA-Pd(0)/PEI]3 films in different solvents. The films contain Pd NPs with different diameters and are deposited on micron-sized alumina. Values in parentheses are the sum of TOFs for hydrogenation and isomerization. TOFs (moles hydrogenated/ mol Pd/ h)a Ratio 0f TOPS =\_ =< fl Catalysts Solvent 1 OH 2 OH 3 OH “2 1/3 Mlxtures Water & 14001100 4 1515 350 93 MeOH (25001250) (30110) (2015) (83) (125) (3: 1) b Catalyst A Water & EtOH 16001400 611 4015 267 40 [Pd(II)]=l mMc (3:1) (29001650) (4515) (4515) (64) (64) Water & THF 6001200 - 25115 - 24 (3:1) (1050150) (-) (-) Water & 12001350 28015 7001100 4.3 1.7 MeOH (390011150) (14001200) (1050120 (2.8) (3.7) (3: 1) 0) Catalyst Db Water & EtOH 8101320 220110 4001100 3.7 2.0 [P d(II)]=15 (3:1) (290011 100) (15001220) (6601160) (1.9) (4.4) mM° Water & THF 300160 - 250120 - 1.2 (3: 1) (130013 90) (-) (-) [a] Several TOFs could not be estimated because GC signals of the substrates, products, or isomers overlapped with the very large peaks corresponding to solvents. [b] Average nanoparticle sizes in catalysts A and D are 2.2 and 3.4 nm respectively. [c] The initial concentration of KdeCl4 in the PAA deposition solutions. order water > water/methanol, water/ethanol > water/THF, which may reflect the degree of swelling of the polyelectrolyte. Higher swelling in more polar solvents should lead to higher reaction rates. With catalyst A, TOFs are higher in water (data in Table 3.3) than the mixed solvent systems (Table 3.7), but there is no clear trend in the different solvents. Table 6 shows that in most cases, selectivities are higher in 3:1 water/methanol than any 100 of the other solvents. The small methanol molecules might more effectively compete for binding sites with 2 and 3 than with 1. 3.3.5. Hydrogenation of Other Unsaturated Alcohols To examine the generality of these results, we studied the hydrogenation of several additional molecules using catalysts A and D, which have approximate NP diameters of 2.2 and 3.4 nm, respectively. As Table 3.8 shows, the rate of hydrogenation of compounds 4-7 decreases in the order monosubstituted double bonds>disubstituted double bond> trisubstituted double bonds. (These reactions occur in 3:1 water/ethanol to ensure substrate solubility.) Compounds 4 and 1 both contain a monosubstituted double bond, but in 3:1 waterzethanol, hydrogenation of 4 by catalyst A occurs at a slower rate than hydrogenation of 1 (compare Tables 3.7 and 3.8). Nevertheless, the TOF for the hydrogenation of 4 is still 11 times higher than that for 6, which contains a disubstituted double bond. Interestingly, using the smaller NPs, the TOF for hydrogenation of 5 is 11- fold lower than that for hydrogenation of 1 in 3:1 waterzethanol. This suggests that when using the smaller NPs, rates are affected by the addition of a bulky substituent at to the double bond, which is consistent with previous studies.24 However, even with 5, hydrogenation using catalyst A is much faster than with the di and trisubstituted compounds 2, 3, 6 and 7. In contrast, catalyst D, which has much larger NPs, shows little difference in the TOFs among the molecules containing monosubstituted double bonds (1, 4, and 5). Furthermore, selectivities for hydrogenation of 4 and 5 vs di and trisubstituted double bonds (6 and 7, respectively) are much smaller when using the larger NPs in catalyst D than when using catalyst A. 101 Table 3.8. TOFs and TOP ratios for unsaturated-alcohol hydrogenation catalyzed by [PAA-Pd(0)/PEI]3 films that contain Pd NPs with different diameters. The catalytic films were deposited on micron-sized alumina. Values in parentheses are the sum of TOFs for hydrogenation and isomerization. TOFs (moles hydrogenated/ mol Pd/ h)a’b OH _ OH M / / Catalysts / OH /\|/\ /'_—/_\__\OH v 4 5 6 7 Catalyst Ac 290:th 150130 2713 812 [Pd(II)]-l de (“03:10) H H (') Catalyst 11" 895145 760110 23312 130110 d (2175165) () (..) (15015) [Pd(II)]=15 mM [a] 25 mM substrate was hydrogenated in 3:1 waterzethanol. TOFs are initial values determined at conversions less than 50%. [b] In some cases, the TOP was not estimated because the amount of isomer detected was negligible. [c] Average nanoparticle sizes for catalysts A and D are 2.2 and 3.4 nm respectively. [d] The initial concentration of KdeCl4 in the PAA deposition solutions. 3.4. Possible Explanation of Selectivity The above data demonstrate that selectivity in hydrogenation varies greatly with particle size, and related studies of NP reactivity imply that such trends arise from geometrical effects.40 Doyle, et al. suggest that the sensitivity of catalytic hydrogenation rates to NP size occurs because the reaction takes place on specific types of atoms that are more (or less) prevalent on small particles.2 Large particles mainly contain terraces with atoms of high coordination numbers, whereas small particles contain comparatively high numbers of surface atoms with low coordination numbers (edge and corner atoms). A recent study based on STM images shows that Pd NPs with diameters <4 nm are highly defective, whereas NPs with diameters >4 nm start to develop large and well defined facets?"41 Additionally, in some cases, if a TOF is calculated with respect to a 102 Face Atoms Defect Atoms Scheme 3.1. Cuboctahedra Pd NP with different atoms. defined type of atom (edge atoms, for example), the dependence of TOF on particle size disappears, and the reaction is deemed structure-sensitive but size-independents'lu' Using the Van Hardeveld and Hartog statistics, we calculated the number of different types of surface atoms on cuboctahedra Pd NPs (Scheme 3.1) containing the same number of atoms as average NPs in catalysts I, J, K and L (NPs suspended in polyelectrolyte solutions) (Table 3.9).42 Figure 3.14 presents the calculated percentage of surface, defect (edge and comer), and face atoms with respect to the total atoms for each average particle size and shows that the percentage of face atoms changes minimally (from 26% to 20%) as the average particle size increases from 2.6 to 5.1 nm. However, the percentage of defect atoms decreases from 18% to 5% as the average particles size increases from 2.6 to 5.1 nm. In general, the TOFs for hydrogenation of monosubstituted compounds decrease with increasing NP diameter, while the TOFs for di- and trisubstituted compounds increase slightly or are unaffected as NP size increases. One 103 Table 3.9. Formula used to calculate m, the number of shells for NPs having cuboctahedron eometry, and the number of Pd atoms at the various positions available on NP surfaces. 2 Total no. of atoms (l/3)(2m—1) (5m2-5m+3) No. of surface atoms 10m2-20m+12 Vertex atoms (5-coordinate) 12 Edge atoms (7-coordinate) 24(m-2) Sq. face (8-coordinate) 6(m-2)2 Triangular face (9-coordinate) 4(m-3)(m—2) The total number of atoms (N) was calculated by the following equation1 N = 471(r)3/(3Vg), where Vg = the volume of one Pd atom =15 A3 explanation for this observation is that in the hydrogenation of monosubstituted double bonds, defect atoms are the primary active centers for catalysis, whereas hydrogenation of multisubstituted double bonds occurs on terraces. Hence, TOFs for the monosubstituted compounds increase as NP size decreases because of a larger fraction of defect atoms in the small NPs. Several recent studies show that interactions of alkenes with catalysts change with alkene size, and hydrogenation of higher alkenes such as trans-2-pentene takes place on face atoms.2’5'43 Rates of hydrogenation on terrace sites should be much less dependent on particle size because the fraction of total atoms that are on terraces is relatively independent of the NP diameter. Thus, it is possible that 104 preferential hydrogenation of monosubstituted double bonds on defect sites leads to high selectivities with small NPs. 50. 45- 40* 35- 30- 251‘ ‘ A +%ofeurfaceatome 20'. A I %ofdefectatoms 15- A %offaceatome 10‘ ' 5-1 l 0 t t . 2.5 3.5 4.5 5.5 Average Nanoparticle Diameter (nm) % of Different Atoms Figure 3.14. Plot of the calculated percentage of surface, defect and face atoms with respect to the total atoms for each particle size. 3.5. Conclusion . Selectivity in the hydrogenation of monosubstituted over disubstituted double bonds increases dramatically as the diameter of catalytic NPs decreases, and this size- dependent selectivity occurs with both suspended NPs and NPs deposited on micron- sized alumina. High selectivities stem from the fact that TOFs for the hydrogenation of monosubstituted double bonds decrease as particle size increases, whereas the TOFs for the hydrogenation of multisubstituted double bonds show the opposite trend. When using the smallest NPs (average diameters 2.2 run), these trends result in selectivities >100 for the hydrogenation of monosubstituted over disubstituted double bonds. Selectivities 105 might occur because monosubstituted double bonds are hydrogenated on defect sites, while hydrogenation of multisubstituted double bonds takes place on terraces. 106 3.6. References ( 1) Wilson, 0. M.; Knecht, M. R.; Garcia-Martinez, J. C.; Crooks, R. M. Journal of the American Chemical Society 2006, 128, 4510. (2) Doyle, A. M.; Shaikhutdinov, S. K.; Freund, H.-J. Angewandte Chemie International Edition 2005, 44, 629. (3) Doyle, A. M.; Shaikhutdinov, S. K.; Jackson, S. D.; Freund, H.-J. Angewandte Chemie International Edition 2003, 42, 5240. (4) Xu, 8.; Liew, K. Y.; Li, J. Journal of American Oil Chemists' Society 2007, 84, 117. (5) Shaikhutdinov, S.; Heemeier, M.; Baumer, M.; Lear, T.; Lennon, D.; Oldman, R. J .; Jackson, S. D.; Freund, H. J. Journal of Catalysis 2001, 200, 330. (6) Chen, J.; Zhang, Q.; Wang, Y.; Wan, H. Advanced Synthesis & Catalysis 2008, 350, 453. (7) Li, F .; Zhang, Q.; Wang, Y. Applied Catalysis A: General 2008, 334, 217. (8) Li, Y.; Boone, E.; El-Sayed, M. A. Langmuir 2002, 18, 4921. (9) Le Bars, J.; Specht, U.; Bradley, J. S.; Blackmond, D. G. Langmuir 1999, 15, 7621. (10) Kell, A. J .; Donkers, R. L.; Workentin, M. S. Langmuir 2005, 21 , 735. (11) Panigrahi, S.; Basu, S.; Praharaj, S.; Pande, S.; Jana, S.; Pal, A.; Ghosh, S. K.; Pal, T. Journal of Physical Chemistry C 2007, 111, 4596. (12) Valden, M.; Lai, X.; Goodman, D. W Science 1998, 281 , 4. (13) Overbury, S. H.; Schwartz, V.; Mullins, D. R.; Yan, W.; Dai, S. Journal of Catalysis 2006, 241, 56. (14) Zhou, W. P.; Lewera, A.; Larsen, R.; Masel, R. 1.; Bagus, P. S.; Wieckowski, A. Journal of Physical Chemistry B 2006, 110, 13393. (15) Zhou, W.; Lee, J. Y. Journal of Physical Chemistry C 2008, 112, 3789. (16) Bergamaski, K.; Pinheiro, A. L. N.; Teixeira-Neto, E.; Nart, F. C. Journal of Physical Chemistry B 2006, 110, 19271. (17) Semagina, N.; Renken, A.; Laub, D.; Kiwi-Minsker, L. Journal of Catalysis 2007, 246, 308. 107 (18) Semagina, N.; Renken, A.; Kiwi-Minsker, L. Journal of Physical Chemistry C 2007, 111, 13933. (19) Claus, P.; Brueckner, A.; Mohr, C.; Hofmeister, H. Journal of the American Chemical Society 2000, 122, 11430. (20) Tardy, B.; Noupa, C.; Leclercq, C.; Bertolini, J. C.; Hoareau, A.; Treilleux, M.; Faure, J. P.; Nihoul, G. Journal of Catalysis 1991, 129, 1. (21) Silvestre-Albero, J.; Rupprechter, G.; Freund, H.-J. Journal of Catalysis 2006, 240, 58. (22) Vasylyev, M. V.; Maayan, G.; Hovav, Y.; Haimov, A.; Neumann, R. Organic Letters 2006, 8, 5445. (23) Bhattacharjee, S.; Bruening, M. L. Langmuir 2008, 24, 2916. (24) Kidambi, S.; Dai, J.; Li, J.; Bruening, M. L. Journal of the American Chemical Society 2004, 126, 2658. (25) Kidambi, S.; Bruening, M. L. Chemistry of Materials 2005, I 7, 301. (26) Oh, S.-K.; Niu, Y.; Crooks, R. M. Langmuir 2005, 21, 10209. (27) Niu, Y.; Yeung, L. K.; Crooks, R. M. Journal of the American Chemical Society 2001, 123, 6840. (28) Schmitt, J.; Decher, G.; Dressick, W. J.; Brandow, S. L.; Geer, R. E.; Shashidhar, R.; Calvert, J. M. Advanced Materials 1997, 9, 61. (29) Mamedov, A. A.; Belov, A.; Giersig, M.; Mamedova, N. N.; Kotov, N. A. Journal of the American Chemical Society 2001, 123, 7738. (30) Wang, T. C.; Rubner, M. F .; Cohen, R. E. Chemistry of Materials 2003, 15, 299. (31) Joly, S.; Kane, R.; Radzilowski, L.; Wang, T.; Wu, A.; Cohen, R. E.; Thomas, E. L.; Rubner, M. F. Langmuir 2000, I 6, 1354. (32) Wang, T. C.; Rubner, M. F .; Cohen, R. E. Langmuir 2002, 18, 3370. (33) Scott, R. W. J.; Ye, H.; Henriquez, R. R.; Crooks, R. M. Chemistry of Materials 2003, 15, 3873. (34) Rouhana, L. L.; Jaber, J. A.; Schlenoff, J. B. Langmuir 2007, 23, 12799. 108 (35) Takaya, H.; Noyori, R. In Comprehensive Organic Synthesis: Selectivity, Strategy and Efliciency in Modern Organic Chemistry; Trost, B. M.; Fleming, 1., Eds., Pergamon: New York, 1991; Vol 8, pp. 443. (36) Liu, X.; Bruening, M. L. Chemistry of Materials 2004, I6, 351.. (37) Sulman, E.; Bodrova, Y.; Matveeva, V.; Semagina, N.; Cerveny, L.; Kurtc, V.; Bronstein, L.; Platonova, O.; Valetsky, P. Applied Catalysis A: General 1999, I 76, 75. (38) Augustine, R. L.; Warner, R. W.; Melnick, M. J. Journal of Organic Chemistry 1984, 49, 4853. (39) Miller, M. D.; Bruening, M. L. Chemistry of Materials 2005, 17, 5375. (40) Bond, G. C. Surface Science 1985, 1 5 6, 966. (41) Freund, H. J.; Baumer, M.; Libuda, J .; Risse, T.; Rupprechter, G.; Shaikhutdinov, S. Journal of Catalysis 2003, 216, 223. (42) Van Hardeveld, R.; Hartog, F. Surface Science 1969, 15, 189. (43) Stacchiola, D.; Burkholder, L.; Tysoe, W. T. Surface Science 2003, 542, 129. 109 CHAPTER 4 CHEMOSELECTIVE NITROARENE HYDROGENATION CATALYZED BY PALLADIUM NPs EMBEDDED IN POLYELECTROLYTE MULTILAYERS 4.1. Introduction Functionalized aniline derivatives are important synthetic intermediates in the production of dyes, herbicides, pesticides, and pharmaceuticals, and aromatic nitro compounds routinely serve as precursors to their corresponding anilines."7 The transformation of nitrobenzene to aniline proceeds through intermediate nitroso and hydroxylamine compounds,8 but slow conversion of hydroxylamine to aniline sometimes 9-12 results in the accumulation of hydroxylamine, as well as the formation of hydrazines,'3"4 azoarenes'5 or azoxyarenes.”21 Reduction of other functional groups in nitroaromatic compounds can also lead to undesired side products.”23 Transformation of nitro compounds to aromatic amines is generally accomplished 22,24-26 through catalytic reduction 112337-33 or hydrogenation, although electrochemical34 or enzymatic reactions have also been used.”36 Stoichiometric reagent-based processes are selective towards the reduction of the nitro functional group, but such processes are frequently expensive and not environmentally friendly. The catalytic hydrogenation of nitroarenes using molecular hydrogen is both a clean and convenient method, but the chemoselective hydrogenation of a nitro group in the presence of other reducible groups is sometimes challenging. Catalytic hydrogenation by H2 is typically less chemoselective than stoichiometric methods.2 Thus, a number of recent studies aimed at developing new catalysts and protocols for chemoselective hydrogenation of nitroarenes with H2. 110 Several methods for chemoselective hydrogenation of nitroarenes catalyzed by 4 supported Pd catalysts were reported.27’37‘39 Holy, et a1. use polystyrene-supported anthranilic acid to anchor a Pd(II) complex for the hydrogenation of aromatic nitro compounds at temperatures ranging from 90-100°C and pressures ranging from 400-1000 psi.38 The reaction times vary from 1 h to 15 h, but hydrogenation of nitrobenzene for 48 h affords only 60% aniline at room temperature and 60 psi pressure. Nitro compounds with different reducible functional groups were not tested in this study. F réchet, et al.. described a polybenzimidazole-supported Pd(O) catalyst for hydrogenation of nitrobenzene at room temperature and under 45 psi H2 pressure.39 However, this catalyst was also not used for hydrogenation of various aromatic nitro compounds containing other reducible functional group. Choudary, et al. showed the use of a montmorillonitesilylpalladium (II) complex for chemoselective hydrogenation of nitro aromatic compounds at room temperature in EtOH, which is illustrated in Scheme 4.1. However, the preparation of the montrnorillonitesilylpalladium (II) complex requires a long procedure as shown in scheme 4.2.37 MW N02 I I . NH2 O-Sl—(CH2)"H2N’PdC'z W H2, EtOH, 5-10 mins ‘ V R= -CHO, -COCH3 Scheme 4.1. Chemoselective hydrogenation of aromatic nitro compounds by a montrnorillonitesilylpalladium (II) complex.37 111 M NaCl W, (EtO)3Si(CH2)3NH£ Montmorillonite ———> 1 . 0.1 N HCI Toluene,reflux 48h 0’?"’(CH2)3NH2 H-Montmorillonite m 1. Pd(PhCHZCNhCIZ I 1 Benzene, rt Q’~?'i"((3Hz)3'\'HzF’dC'2 f 2. Soxhlet extraction 8h Scheme 4.2. Preparation of a montmorillonitesi1ylpalladium (1.1) complex. As mentioned in previous chapters, metal NPs are frequently better catalytic materials than bulk metals because of their high surface area to volume ratio and their size-dependent electronic properties.‘“"12 The use of metal NPs for the chemoselective reduction of nitroarene derivatives recently attracted tremendous interest.22'24’28’30’3l’43’44 Recent work by Sema, et al. published in Science shows that titania-supported gold NPs provide selectivity for nitro group hydrogenation in the presence of C=C, C=O, and CEN functional groups. However, these reactions occur under high temperature (1 10-140°C) and pressure (9-25 bar) with relatively long reaction times (0.03 to 9.5 h).3| Ranu, et al. used Cu NPs and ammonium formate in ethylene glycol for reduction of aromatic nitro compounds with high chemoselectivity. However, these reactions take place under high temperature (120°C) and very long reaction times (8-12 h).24 Pd NPs, in particular, generated in situ22 or supported on nanofibers23 are used for reduction of nitro compounds. Maleczka, et al. prepared Pd NPs in situ using polymethylhydrosiloxane (PMHS), palladium (II) acetate, and aqueous potassium fluoride. Pd NPs generated in situ chemoselectively reduce aromatic nitro compounds in 112 a high yield with short reaction times (30 min). However, in some cases either reduction of aromatic nitro compounds is not highly chemoselective (see scheme 4.3) or the reduction stops at an intermediate stage. Motoyama, et al. prepared Pd and Pt NPs supported on carbon nanofiber catalyst and used them for chemoselctive hydrogenation of several aromatic nitro compounds containing other reducible functional groups. However, the reactions took place under high pressure (10 atrn) and required relatively long times (4 h) to complete. Additionally, in this case, hydrogenation of aromatic nitro compounds is not always highly chemoselective as shown in scheme 4.4. Therefore, an alternative efficient yet simple chemoselective procedure is highly desirable. 4 equiv PMHS NO NC)2 2 equiv KF aq NH2 2 5 mol% Pd(OAc)2, THF, rt, 30 min 1 + CHO CHO CH20H 73% 27% Scheme 4.3. Silicon hydride reduction of nitro benzaldehyde catalyzed by Pd NPs generated in-situ.22 N02 H2 (3-10 atm) PtIPd supported on carbon nanofiber (catalyst): AcOEt, rt-5 50°C CN CH2NH2 64% 32% Scheme 4. 4. Chemosezl3ective hydrogenation of p-nitrobenzonitrile catalyzed by Pd NPs on a carbon nanofiber.23 113 This chapter describes a very simple, room-temperature protocol for highly chemoselective hydrogenation of aromatic nitro compounds containing other readily reducible functional groups. The procedure uses the catalytic [PAA-Pd(0)/PEI]3 films described in previous chapters and a new catalyst prepared by reducing PdC142' impregnated in alumina. These catalysts show nearly complete selectivity for hydrogenation of nitro groups in the presence of aldehyde, nitrile, amide, ketone, and ester groups. 4.2. Experimental 4.2.1. Materials Poly(ethylenimine) (PEI, Mw = 25 000 Da), poly(acrylic acid) (PAA, Mw = 90 000 Da), a-alumina (100 mesh, typical particle size 75-100 um), potassium tetrachloropalladate(ll) (99.99%), sodium borohydride (98%) and all nitro compounds were purchased from Aldrich and used as received. Ethanol (100%) was obtained from Pharmco. Aqueous solutions were prepared with deionized water (18.2 M!) cm, Milli-Q purification system), and hydrogen (99.9%) was obtained from AGA gases. Flash chromatography was performed with silica gel 60, particle size 0.040-0.063 mm (230— 400 mesh), which was purchased from EM science. 1H NMR and '3C NMR spectra were recorded on a Varian spectrometer (300 MHz for 1H and 75 MHz for 13C, respectively), with chemical shifis reported relative to the residue peaks of solvent chloroform (6 7.24 for ]H and 77.0 for 13C). Synthetic yields refer to chromatographically and spectroscopically pure compounds unless otherwise stated. GC-MS (Hewlett-Packard G18OOB GCD system with a nonpolar HP-5 MS column) was used to analyze hydrogenation reactions. All reactions were carried out in clean glassware with magnetic 114 stirring and were monitored by thin-layer chromatography with precoated silica gel plates, unless otherwise noted. 4.2.2. Preparation of Pd N Ps on Alumina 4.2.2.1. [PAA-Pd(0)/PEI]3 [PAA-Pd(0)/PEI]3 films on alumina were prepared following the procedure in Chapter 3. To form an initial PAA layer, alumina was vigorously stirred for 10 min in a pH 4 solution containing 20 mM PAA (molarity of PAA is given with respect to the repeating unit) and 15 mM K2PdC14. After rinsing, the alumina coated with PAA was stirred in a pH 9 solution containing PEI (1 mg/mL) to form a PAA-Pd2+/PEI film. After the deposition of each layer, the coated alumina was allowed to settle, the supernatant was decanted, and the solid was washed through vigorous stirring in three 100-mL aliquots of deionized water (3 x 5 min) to remove excess polyelectrolyte. In total, three PAA-Pd(II)/PE1 bilayers were deposited on the alumina powder. Pd(0) nanoparticles were formed by stirring the [PAA-Pd(II)/PEI]3 on alumina in 100 mL of fieshly prepared 0.1 mM NaBH4 for 30 min, and the reduced films were washed with three lOO-mL aliquots of deionized water (3 x 5 min). After decanting the final water rinse, the residual water was removed under reduced pressure overnight to afford a catalytic powder, which was stored in a clean dry vial until use. 4.2.2.2. Reduced Pdle' on alumina Fifteen mg of a-alumina was stirred in a 100-mL solution containing 0.5 mM PdC142' and 0.1 M KCl for 30 min. The powder was then washed with three lOO-mL 115 aliquots of deionized water (3 x 5 min), and the Pd(II) was reduced to Pd(O) as described above to afford a catalytic powder. 4.2.3. Characterization of Pd NPs To obtain TEM images, [PAA-Pd(0)]/PEI]3 films were deposited on carbon- coated copper grids that were pretreated in a UV/ozone cleaner for 1 min. The deposition procedure on the TEM grid is similar to that described above, except that the grid was simply immersed for 5 min in solutions without stirring, and rinsing between each step consisted of a l-min washing with deionized water. Imaging was performed on a JEOL 2200F S microscope using an accelerating voltage of 200 kV. 4.2.4. Determination of the Pd Content of Catalysts Used in Hydrogenation The amounts of palladium in different catalysts were determined by atomic emission spectrophotometry (AES). Standard solutions (0.1 to 0.5 mM) were prepared by dissolving K2PdCl4 in 0.1 M HNO3, and sample solutions were prepared by stirring 60 mg of catalyst in 2 mL of aqua regia for 15 min. The sample solutions were diluted to 12 mL with water and centrifuged (the a—alumina support does not dissolve in aqua regia) to collect supernatant to analyze using its emission at 340.5 nm. 4.2.5. General Procedure for the Hydrogenation of N itroaromatic Compounds To determine catalyst turnover frequencies (TOFs), catalytic hydrogenation was performed in a 100-mL, three-neck, round-bottomed flask. 30 mg of catalyst was suspended in a 25-mL aliquot 'of deionized water and bubbled with H2 for 30 min. 116 Subsequently, a solution of nitrobenzene (2.5, 7.5, 15, 22.5, or 30 mM) in 25 mL of a 1:1 (v:v) mixture of ethanol or methanol or THF and water was prepared and added to the reaction vessels. During the reaction, hydrogen at a gauge pressure of 50 kPa was bubbled through a frit at the bottom of the solution, which was vigorously stirred. Sample aliquots (0.5 mL) were removed from the reaction vessel at specific times and diluted by addition of 1 mL of THF prior to analysis by GC-MS with appropriate response factors. The response factors were determined to calculate the TOF as described in chapter 2.45 TOFs were calculated from the slope of the initial portion (total conversion of reactant to isomer and hydrogenated product <50%) of plots of percent hydrogenation versus time. In fitting the initial data, the intercept of the line was forced to be zero. In reactions used for product isolation, 125 mg of catalyst (more catalyst was used to speed up the reaction) was suspended in 25 mL of EtOH or a mixture of 20 mL water and 5 mL of Ethanol and bubbled with H2 for 30 min. A 30 mM solution of substrate in 25 mL EtOH was added to the reaction vessels and the mixture was stirred until the hydrogenation was complete as judged by TLC. The reaction mixture was filtered through cotton with a top layer of Celite to remove the catalyst, and the solvent was evaporated under reduced pressure, dissolved in a minimum volume of solvent, and subjected to flash chromatography using hexanes/EtOAc. 4-Aminobenzaldehyde (2): Hydrogenation of 4-nitrobenza1dehyde (0.75 mmol, 0.113 g) in 3:2 EtOH and water for 15 min afforded >99% conversion. After collection of the crude product and flash chromatography (hexanes/EtOAc: 5:1), nitroarenes were 117 collected as a yellow liquid in the presence of solvent. Upon complete removal of the solvent, a yellow solid formed, which was insoluble in all solvents tested (presumably because of polymerization). Therefore, NMR in CDC13 was performed with a sample that contained residual EtOAc. '11 NMR (300 MHz, CDC13): 8 9.64 (s, 1H), 7.60 (d, r = 6.59 Hz, 2H), 6.61 (d, J = 6.59 Hz, 2H), 4.50 (bs, 2H); 13C NMR (75 MHz, CDC13): 5 190.3, 152.8, 132.2, 126.9, 113.8.22 4-Aminobenzonitrile (3): 4-nitrobenzonitrile (0.75 mmol, 0.111 g) was subjected to the general procedure for reducing nitroarenes with stirring for 90 min followed by flash chromatography (hexanes/EtOAc: 5:1) to afford an inseparable mixture (0.084 g) of 4- aminobenzonitrile (95%) and 4-hydroxyaminobenzonitrile (5%) as a bright yellow solid. For 4-aminobenzonitrile: ‘H NMR (500 MHz, CDC13): 8 7.36 (d, J = 8.79 Hz, 2H), 6.61 (d, J = 8.79 Hz, 2H), 4.23 (bs, 2H); 13C NMR (125 MHz, CDC13)I 8 150.6, 133.6, 120.2, 114.3, 99.6.22 Methyl 4-aminobenzoate (4): Methyl 4-nitrobenzoate (0.75 mmol, 0.136 g) was subjected to the general procedure for reducing nitroarenes with stirring for 15 min followed by flash chromatography (hexanes/EtOAc: 5:1) to afford 0.11 g (97% yield) of methyl 4-aminobenzoate (14) as a yellow solid. 1H NMR (300 MHz, CDCl3): 8 7.83 (d, J = 8.14 Hz, 2H), 6.61 (d, J = 8.70 Hz, 2H), 4.06 (bs, 2H), 3.79 (s, 3H); 13C NMR (75 MHz, CDC13): 8 167.1, 150.8, 131.6, 119.7, 113.7, 51.6.22 4-Aminoacetophenone (5): 4-nitroacetophenone (0.75 mmol, 0.124 g) was subjected to the general procedure for reducing nitroarenes with stirring for 15 min followed by flash chromatography (hexanes/EtOAc: 5:1) to afford 4-aminoacetophenone (5) 0.094 g (96% yield) as a yellow solid. 1H NMR (300 MHz, CDC13): 6 7.78 (d, J = 6.15 Hz, 2H), 6.62 118 (d, J = 6.08 Hz, 2H), 4.06 (bs, 2H), 2.48 (s, 3H); 13C NMR (75 MHz, CDC13): 8 196.5, 151.3, 130.7, 127.6, 113.6, 26.0.22 4-Aminobenzamide (6): 4-nitrobenzamide (0.75 mmol, 0.125 g) was subjected to the general procedure for reducing nitroarenes with stirring for 15 min followed by flash chromatography (EtOAc/MeOH, 100/0 then 9:1) to afford 0.096 g (94% yield) of 3- aminobenzamide (6) as a yellow solid. 1H NMR (300 MHz, CD30D): 8 7.63 (d, J = 7.58 Hz, 2H), 6.65 (d, J = 7.49 Hz, 2H), 4.87 (s, 4H); 13C NMR (75 MHz, CD3OD): 8 172.8, 153.5, 130.5, 122.2, 114.6.22 2-Aminotoluene (7): 2-nitrotoluene (0.75 mmol, 0.088 mL) was subjected to the general procedure for reducing nitroarenes with 60 min of stirring followed by flash chromatography (hexanes/EtOAc: 5:1) to afford 0.77 g (95% yield) of 2-aminotoluene as a yellow oil. 1H NMR (300 MHz, CDCl3)! 6 7.04 (m, 2H), 6.74-6.66 (m, 2H), 3.55 (bs, 2H), 2.17 (s, 3H); 13C NMR (75 MHz, CDC13)I 8 144.5, 130.4, 126.9, 122.3, 118.6, 114.9, 17.3.22 4.3. Result & Discussion 4.3.]. Catalyst Preparation and Characterization We prepared [PAA-Pd(0)/PEI]3 films on micron-size alumina particles using 15 mM Pd(II) in the PAA-Pd(II) deposition solution and reduction of Pd(II) to Pd(O) by NaBI-I4. Pd NPs embedded in these polyelectrolyte multilayers are stable and can be stored for months without losing their catalytic activity. Moreover, as shown in chapter 2, the catalytic activity of this material is not altered by recycling.45 The average diameters of small NPs in these [PAA-Pd(0)/PEI]3 films on TEM grids are 3.4109 nm 119 (see chapter 3). Additionally, some big particles with average diameters of 1601100 nrn appear in some images, but we roughly estimate that these particles constitute only 18% of the total particle surface area (see chapter 3). Figure 4.1 shows the TEM image of small particles in a [PAA-Pd(0)/PEI]3 film on a TEM grid and as well as a histogram of the diameters of the small NPs. Figure 4.1. (a) TEM image of a [PAA-Pd(0)/PEI]3 film on a carbon-coated copper grid. 15 mM Pd(II) was present in the PAA-K2PdCl4 solution used for film deposition. 120 '0 0| 1 tn .2 20- b .2 E Q15- 2 «110- 2 h 0 5. ,,\° 0 .3 on 2.4 3 3.6 4.2 4.8 5.4 6 Nanoparticle Diameter (nm) Figure 4.1. (b) Histogram of Pd NP diameters in several TEM image based on their area. 4.3.2. Hydrogenation of nitrobenzene as a function of solvent and nitrobenzene concentration To maximize TOFs, we initially evaluated the catalytic activity of [PAA- Pd(0)/PEI]3 films on alumina for hydrogenation of nitrobenzene in several solvent systems. As Table 1 shows, hydrogenation in THF and a mixture of water and THF gives TOFs that are significantly lower than hydrogenation in pure EtOH or in a mixture of EtOH and water. Previous reports also showed that solvents affect the rate of the hydrogenation of aromatic nitro compounds,46 and EtOH is commonly used to perform these transformations.27’30’37'39 We further examined hydrogenation using different initial nitrobenzene concentrations. The data in Table 4.2 show that TOFs do not vary significantly with nitrobenzene concentrations ranging from 2.5 mM to 30 mM, which suggests that even at 2.5 mM nitrobenzene, the catalyst is essentially saturated with substrate. The calculated TOF for the hydrogenation of 15 mM nitrobenzene by Pd NPs 121 in [PAA-Pd(O)/PEI]3 films is about the same as the TOF with commercially available 5% Pd on alumina (Table 4.2), and films prepared by impregnation of the alumina with [PdC142'] and subsequent reduction of Pd(II) with NaBI-I4 also show similar TOFs. Table 4.1. TOFs for nitrobenzene hydrogenation catalyzed by alumina-supported [PAA- Pd(II)/PEI]3 films in different solvents. no2 NH2 30 mg of [PAA-Pd(0)/PEI]3 © on alumina _ H2 (50 kpa). rt. 1 solvent 1 a TOFs (moles hydrogenated/ mol Pd/ h)a (3:1) (3:1) 100% 100% Water & EtOH Water & THF EtOH THF 1 320 460 1460 450 [a] 50 mL of 7.5 mM substrate was hydrogenated in different solvents. TOFs are initial values determined at conversions less than 50%. Table 4.2. TOFs (moles hydrogenated/moles Pd/h) for nitrobenzene hydrogenation by different catalysts in EtOH. TOFs (moles hydrogenated/ mol Pd/ h)a [PAA-Pd(II)/PEI]3 film 5% Pd Reduced PdCl42' on alumina on on alumina alumina 'tr b 2.5b 7.5 15 22.5 30 b b [n1 0 enzene] mM me me me me 15 mM 15 mM TOF 1350 1460 1950 1690 1540 1430 1950 [a] TOFs are initial values determined at conversions less than 50%. [b] Substrate’s initial concentration in the reaction mixture. 122 4.3.3. Hydrogenation of nitrobenzaldehyde In many synthetic reactions, aromatic nitro compounds must be selectively hydrogenated in the presence of other functional groups such as aldehydes and nitriles. Recent report indicate that aminals can be easily accessed from aminobenzaldehyde,‘17 so we initially investigated whether [PAA-Pd(0)/PEI]3 catalysts can facilitate chemoselective hydrogenation of 4-nitrobenzaldehyde, 2, to 4-aminobenzaldehyde. Aminals are present in many natural product48'49 and are also important building blocks in the preparation of many biological active compounds.”55 The chemoselective reduction of nitrobenzaldehyde to amino benzaldehyde is challenging due to the presence of the easily reducible carbonyl group (Scheme 4.5). Using EtOI-l as a solvent, we carried out the hydrogenation of 2 in the presence of [PAA-Pd(0)/PEI]3 on alumina, reduced [PdCl42'] on alumina, 5% Pd on alumina, or 10% Pd on carbon. Table 4.3 shows that in EtOH, [PAA-Pd(0)/PEI]3 films and reduced [PdClaz'] on alumina are the most selective catalysts for the hydrogenation of 2 to 4- aminobenzaldehyde (2a). The use of [PAA-Pd(0)/PEI]3 films and reduced [PdClaz'] on alumina gives only 2a and 4-aminobenzyl alcohol (2b) as hydrogenation products, and the ratio of these compounds in the product is about 17:1. The chemoselectivity with the commercial catalysts in EtOH is much lower than with the [PAA-Pd(0)/PEI]3 films, and the Pd on C shows much more reduction of the aldehyde than 5% Pd on alumina. The relatively small size of NPs in [PAA-Pd(0)/PEI]3 films may account for their high chemoselectivity in the hydrogenation of 2, but firrther studies are needed to confirm that this is the case. In contrast to a previous report (Scheme 4.3), we do not observe the formation of 4-nitrobenzyl alcohol (2d) with any catalyst.22 123 N02 NHZ Catalyst H2 (50 kpa), rt solvent CHO CH20H CH2 OH 2 Scheme 4.5. Hydrogenation of 4-nitrobenzaldehyde to several possible products. Table 4.3. Product distributions from 4-nitrobenzaldehyde hydrogenation (>99% conversion) using several catalysts and two solvent systems. % of Different Products Catalyst Solvent 2a 2b 2c 2d 100% EtOH [PAA-Pd(O)/PEI]3 (15 min) 94 6 - - A1 ' on umrna (32) EtOH and water 99.3 0.40 - - (20 min) 100% EtOH 5% Pd (15 min) 79.2 20.8 - - . a on Alumrna (3:2) _ _ EtOH and water 98.7 0.40 (30 min) 100% EtOH Reduced [PdC142'] (15 min) 95 5 - - on Alumrna (7: 5) EtOH and water 100 - - - (20 min) 10% Pd 100% EtOH on Carbona (10 min) 39 50 10.5 - (15 rmn) (3:2) EtOH and water 94.5 3.5 2 - (15 min) [a] Substrate concentration was initially 15 mM in the reaction mixture. [b]Substrate concentration was initially 10 mM in the reaction mixture. 124 We also performed the hydrogenation reaction in a 3:2 (v:v) EtOH and water mixture. The limited solubility of 2 in water restricts the total water content for this reaction to about 40%. Water is a byproduct of the hydrogenation of aromatic nitro compounds. However, interestingly, we observe improved chemoselectivity in the presence of water for the hydrogenation of 2 to 2a with all of the catalysts. Further studies needs to be performed to understand the mechanistic reason why chemoselectivity increases in the presence of water. 4.3.4. Hydrogenation of Other Nitroarenes To further explore the scope of hydrogenation with [PAA-Pd(0)/PEI]3 films and reduced [PdC142'] on alumina, we examined the hydrogenation of the nitroarenes shown in Table 4.4. 4-nitrobenzonitrile (3) is an interesting example among them. Maleczka, et al. reported that the reduction of 3 by a hydride reducing agent PMHS, produced the corresponding N-hydroxylamine (3b), even when the reaction was performed for 12 h instead of 30 min22 Cu and Au nanoparticles promoted the catalytic transformation of 3 to 4-aminobenzonitrile (3a) chemoselectively, but these reactions were run at 120°C and 110-140°C, respectively.24’3°’3 l Corma, et al. reported that Pd/C and PVC catalysts are not chemoselective for the hydrogenation of 4-nitrobenzonitrile at high temperature and pressure (140°C, 25 bar).31 Motoyama, et al. recently showed that the hydrogenation of 3 catalyzed by nanofiber-supported Pd NPs produces 64 % 3a and 32% di(p-aminobenzyl) amine (see Scheme 4.4).23 In our case, the hydrogenation of 4-nitrobenzonitrile by [PAA-Pd(0)/PEI]3 films in pure EtOH produces 95% of the desired compound, 3a, and only 5% 3b. The ratio of 3a and 3b in the product did not change when the reaction was performed for 90 min instead of 15 mi 125 Table 4.4. Catalytic hydrogenation of various aromatic nitro compound by [PAA- Pd(0)/PEI]3 films and reduced [PdClaz'] on alumina. [PAA- Reduced Pd(O)/PEI]3 [PdClaz‘] Product Substratesa . b . b Yleld (%) Yleld (%) (Time, min) (Time, min) NC—QNOz NC‘Q'NW 88c 90d 90 4O 3 31 < ) ( ) H3COZC-_©-N02 H30020‘Q-NH2 97 98 (15) (10) 4 4a 92 98 Hecoc—Q—Noe “acOCQNHZ (15) (10) 5 5a (15) (10) 6 6a 95 98 QNOZ Q'NHZ (60) (120) 7 7a [a] 15 mM substrate was subjected to hydrogenation conditions in the presence of 125 mg of [PAA- Pd(0)/PEI]3 on alumina in 50 mL of EtOH at room temperature. [b] >99% consumption (calculated by GC/MS) of the starting materials takes place. [0] The overall isolated yield is 93%, which consists of 95% 3a and 5% 3b. [(1] The overall isolated yield is 96%, which consists of 94% 3a and 6% 3b. A similar ratio of 3a and 3b (17:1) was obtained in the hydrogenation of 3 catalyzed by reduced PdCl42' on alumina. With both [PAA-Pd(0)/PEI]3 and reduced PdCl42' on alumina catalysts, the hydrogenation of nitroarenes containing electron withdrawing substituents such as esters 126 (4), ketones (5) and amides (6) is 100% chemoselective and allow isolation of very high yields of the desired products 4a, 5a, and 6a after flash chromatography. Long reaction times are required for hydrogenation of 7 due to the steric hindrance caused by the methyl group at the ortho position, but >99% conversion can be achieved within a few hours. Taken together, these studies demonstrate the high chemoselectivity of [PAA- Pd(0)/PEI]3 and reduced PdCl42' on alumina catalysts. The high selectivity of these catalysts relative to commercial Pd/C and 5% Pd on alumina may relate to the size of the Pd particles in the catalysts. Unfortunately, it is very difficult to image the Pd nanoparticles on the very rough supports, so we do not know their sizes at this time. Future studies of selectivity as a firnction of NP size, similar to those in chapter 3, may shed light on the mechanism of chemoselectivity. 4.4. Conclusion In conclusion, NPs embedded in polyelectrolyte multilayer films on micron-size alumina powder provide highly active and chemoselective catalysts for the hydrogenation of nitroarenes to their corresponding amines. Reduced PdCl42’ on alumina is similarly active and selective. Reactions in ethanol proceed more rapidly than reactions in THF, and mixtures of water and ethanol give higher chemoselectivity than that pure ethanol in the case of hydrogenation of 4-nitrobenza1dehyde. Chemoselective hydrogenation of aromatic nitroamides, nitroesters, nitroaldehydes, and nitronitriles is also possible with these catalysts. 127 4.5. Appendix ~ 11 r- H r- N ’— M NH2 6 r— ' ). ‘— II Figure 4. A-l. 1H NMR spectrum of purified 2—Aminotoluene (compound 7) in CDCl3. 128 20 40 NH2 100 120 140 160 II II JITJIJJ 'rrrrrrrrrrrrr IIITTT II]UYJYrIJlT—rlefIIIJIIT‘rrIIYTITIIIIIIIVIT'IWTfTIJIIIIIIITTT I 180 I Figure 4. A-2. '3 C NMR spectrum of purified 2-Aminotoluene (compound 7) in CDCl3. 129 NH2 1 cozcn3 JL Figure 4. A-3. 1H NMR spectrum of purified Methyl-4-aminobenzoate (compound 4) in CDC13. 130 711111I’111] 20 NH2 40 CO2CH3 WWW [FJIIIIJIIIIIIIIIIWII11111111II 100 FITTrTIITIWIIIIlTIIITI 120 140 160 ITI'I #TTTTllllllljl'rlllllIITTIIT 180 l r v Figure 4. A-4. 13’C NMR spectrum of purified Methyl-4-aminobenzoate (compound 4) in CDC13. 131 raj pn- NH2 I I I I CONH2 _ _J 1 i —‘ _ .1 fl ___J ‘l—n —to 1.1— it 1—b _J j Figure 4. A-S. 1H NMR spectrum of purified 4-Aminobenzamide (compound 6) in CD30D. 132 .i .llillllI “.111“ “11.1 1 Jill iu.11||t.t‘I l"| l NH2 CONH2 100 120 1‘0 160 i ‘ I 1‘ I, . l‘w ‘ “ ‘ 1 ‘ . ‘ l ‘ y‘ . ‘ l 1‘. I 10111 I 1111....“Js111.11.41] lulillltlui II I nil-1.111 1‘11”le l .. ‘Il 1L.1. t‘dl .I It . .1 1 ‘ 180 Jldil I‘lli‘di. [III 1h .5 Figure 4. A-6. '3 C NMR spectrum of purified 4-Aminobenzamide (compound 6) in CD30D. 133 r- N NH2 i i + EtOAc CHO 11 . 1111 J I - "_ Figure 4. A-7. 1H NMR spectrum of purified 4-Aminobenzaldehyde and residual AcOEt (compound 2) in CDC13. 134 NH2 + EtOAc CH0 E 1 100 l 120 140 160 ITIIIIITIIIIIIII III 1.0 Figure 4. A-8. '3 C NMR spectrum of purified 4-Aminobenzaldehyde and residual AcOEt (compound 2) in CDC13. 135 NH2 1 CN I j )— Figure 4. A-9. ]H NMR spectrum of purified 4-Aminobenzonitrile (compound 3) in CDC13. 136 I E .. _ O __ N )— NH ~ 2 a _ O _ ‘ __ O 2 90 d r— b ' o )— D E +- O' —_—_—_4 - Q ___4 ~ H L r— b __‘_________-_ .. t r l r O _: r— N >- H fl l- -—-4 I— :l P r O r— ' - H ——1 l- “ O r‘ In - H r ‘ O — O t- r. Figure 4. A-l0. '3 C NMR spectrum of purified 4-Aminobenzonitri1e (compound 3) in CDC13. 137 .1. . NH2 1 COCH3 JL Figure 4.A-11. 1H NMR spectrum of purified 4-Aminoacetophenone (compound 5) in CDCl3. 138 NH2 COCH3 IT'IIIIIIIIIIITTIIITIITTIIIIIITIIIIII'Ij—r 100 120 1‘0 160 180 IIIIITIIIIIWIII‘IrTTIIIIIIFITIthIITrIIIIrITJIII'IHWI]II Figure 4. A-12. ‘3 C NMR spectrum of purified 4-Aminoacetophenone (compound 5) in CDC13. 139 4.6. References (l) Tafesh, A. M.; Weiguny, J. Chemical Reviews 1996, 96, 2035. (2) Downing, R. S.; Kunkeler, P. J .; van Bekkum, H. Catalysis Today 1997, 37, 121. (3) Adams, J. P. Journal of the Chemical Society, Perkin Transactions 1 2002, 2586. (4) Ono, N. The Nitro Group in Organic Synthesis; Wiley-VCH: New York 2001. (5) Dale, D. J.; Dunn, P. J.; Golightly, C.; Hughes, M. L.; Levett, P. C.; Pearce, A. K.; Searle, P. M.; Ward, G.; Wood, A. S. Organic Process Research & Development 2000, 4, l7. (6) Brickner, S. J.; Hutchinson, D. K.; Barbachyn, M. R.; Manninen, P. R.; Ulanowicz, D. A.; Garmon, S. A.; Grega, K. C.; Hendges, S. K.; Toops, D. S.; Ford,C. W.; Zurenko, G. B. Journal of Medicinal Chemistry 1996, 39, 673. (7) Jacobsen, M. F .; Moses, J. E.; Adlington, R. M.; Baldwin, J. E. Organic Letters 2005, 7, 641. (8) Haber, F. Elektrochemistry 1898, 22, 506. (9) Rondestvedt, C. S., Jr.; Johnson, T. A. Synthesis 1977, 850. (10) Entwistle, I. D.; Gilkerson, T.; Johnstone, R. A. W.; Telford, R. P. Tetrahedron 1978, 34, 213. (11) Feuer, H.; Bartlett, R. 8.; Vincent, B. F ., Jr.; Anderson, R. S. Journal of Organic Chemistry 1965, 30, 2880. (12) Yanada, K.; Yamaguchi, H.; Meguri, H.; Uchida, S. Journal of the Chemical Society, Chemical Communications 1986, 1655. (13) Furst, A.; Moore, R. B. Journal of the American Chemical Society 1957, 79, 5492. (14) Olah, G. A. Journal of the American Chemical Society 1959, 81, 3165. (15) Hutchins, R. 0.; Lamson, D. W.; Rua, L.; Milewski, C.; Maryanoff, B. Journal of Organic Chemistry 1971-, 36, 803. (16) Osuka, A.; Shimizu, H.; Suzuki, H. Chemistry Letters 1983, 1373. (17) A200, J. A.; Grimshaw, J. Journal of the Chemical Society C: Organic 1968, 2403. 140 (18) Ohe, K.; Uemura, S.; Sugita, N.; Masuda, H.; Taga, T. Journal of Organic Chemistry 1989, 54, 4169. (19) Ogata, Y.; Mibaye, J. Journal of Organic Chemistry 1962, 27, 2048. (20) Hutton, J .; Waters, W. A. Journal of the Chemical Society B: Physical Organic 1968, 191. (21) Porta, F .; Pizzotti, M.; Cenini, S. Journal of Organometallic Chemistry 1981, 222, 279. (22) Rahaim, R. J ., Jr.; Maleczka, R. E., Jr. Organic Letters 2005, 7, 5087. (23) Takasaki, M.; Motoyama, Y.; Higashi, K.; Yoon, S.-H.; Mochida, 1.; Nagashima, H. Organic Letters 2008, 10, 1601. (24) Saha, A.; Ranu, B. Journal of Organic Chemistry 2008, 73, 6867. (25) Adger, B. M.; Young, R. G. Tetrahedron Letters 1984, 25, 5219. (26) Yuste, F.; Saldana, M.; Walls, F. Tetrahedron Letters 1982, 23, 147. (27) Xi, X.; Liu, Y.; Shi, J .; Cao, S. Journal of Molecular Catalysis A: Chemical 2003, I92, 1. (28) Zuo, B.; Wang, Y.; Wang, Q.; Zhang, J.; Wu, N.; Peng, L.; Gui, L.; Wang, X.; Wang, R.; Yu, D. Journal of Catalysis 2004, 222, 493. (29) Magdalene, R. M.; Leelamani, E. G.; Nanje Gowda, N. M. Journal of Molecular Catalysis A: Chemical 2004, 223, 17. (30) Chen, Y.; Qiu, J .; Wang, X.; Xiu, J. Journal of Catalysis 2006, 242, 227. (31) Corma, A.; Sema, P. Science 2006, 313, 332. (32) Deshpande, R. M.; Mahajan, A. N.; Diwakar, M. M.; Ozarde, P. S.; Chaudhari, R. V. Journal of Organic Chemistry 2004, 69, 4835. (33) Maity, P.; Basu, S.; Bhaduri, S.; Lahiri, G. K. Advanced Synthesis & Catalysis 2007, 349, 1955. (34) Noel, M.; Ravichandran, C.; Anantharaman, P. N. Journal of Applied Electrochemistry 1995, 25, 690. (35) Blackie, J. A.; Turner, N.; Wells, A. Tetrahedron Letters 1997, 38, 3043. 141 (36) Li, F .; Cui, J .; Qian, X.; Zhang, R. Chemical Communications 2004, 2338. (3 7) Mukkanti, K.; Rao, Y. V. S.; Choudary, B. M. Tetrahedron Letters 1989, 30, 251. (3 8) Baralt, E.; Holy, N. Journal of Organic Chemistry 1984, 49, 2626. (39) Li, N. H.; Frechet, J. M. J. Journal of the Chemical Society, Chemical Communications 1985, 1100. (40) Chen, S.; Ingrrna, R. S.; Hostetler, M. J .; Pietron, J. J .; Murray, R. W.; Schaaff, T. G.; Khoury, J. T.; Alvarez, M. M.; Whetten, R. L. Science 1998, 280, 2098. (41) Li, Y.; Boone, E.; El-Sayed, M. A. Langmuir 2002, 18, 4921. (42) Rao, C. N. R.; Kulkami, G. U.; Thomas, P. J.; Edwards, P. P. Chemistry-A European Journal 2002, 8, 28. (43) Blaser, H.-U. Science 2006, 313, 312. (44) Chen, Y.; Wang, C.; Liu, H.; Qiu, J.; Bao, X. Chemical Communications 2005, 5298. (45) Bhattacharjee, S.; Bruening, M. L. Langmuir 2008, 24, 2916. (46) Aramendia, M. A.; Borau, V.; Gomez, J.; Jimenez, C.; Marinas, J. M. Applied Catalysis 1984, 10, 347. (47) Zhang, C.; De, C. K.; Mal, R.; Seidel, D. Journal of the American Chemical Society 2008, 130, 416. (48) Crich, D.; Banerjee, A. Accounts of Chemical Research 2007, 40, 151. (49) Hajicek, J .; Taimr, J .; Budesinsky, M. Tetrahedron Letters 1998, 39, 505. (50) Mhaske, S. B.; Argade, N. P. Tetrahedron 2006, 62, 9787. (51) Mhaske, S. B.; Argade, N. P. Journal of Organic Chemistry 2001, 66, 9038. (52) Liu, J.-F.; Ye, P.; Sprague, K.; Sargent, K.; Yohannes, D.; Baldino, C. M.; Wilson, C. J .; Ng, S.-C. Organic Letters 2005, 7, 3363. (53) Lee, E. S.; Park, J.-G.; Jahng, Y. Tetrahedron Letters 2003, 44, 1883. (54) Hamid, A.; Elomri, A.; Daich, A. Tetrahedron Letters 2006, 47, 1777. 142 (55) Bowman, W. R.; Elsegood, M. R. J.; Stein, T.; Weaver, G. W. Organic & Biomolecular Chemistry 2007, 5, 103. 143 CHAPTER 5 CONCLUSIONS AND FUTURE WORK This dissertation demonstrates that metal ions can be incorporated inside polyelectrolyte films and subsequently reduced by NaBH4 to give polymer-encapsulated NPs. Most importantly, these NPs are highly selective in the hydrogenation of unsaturated alcohols and nitroarenes. However, this work has only begun to probe the potential of NPs to serve as highly selective catalysts. There are likely a number of reactions in which NP size plays a vital role in effecting catalytic selectivity. Below I briefly summarize the most important points of my research and suggest a number of directions in which this research could expand in scope and importance. 5.1. Selective Catalysis Chapter 2 shows that the rate of hydrogenation of allyl alcohols containing methyl substituents on the double bond is order(s) of magnitude lower than the rate of hydrogenation of allyl alcohol. Such selectivity depends on both the number of substituents on the double bond and the size of these substituents. Polyelectrolyte- embedded Pd NPs also show intramolecular selectivity for the hydrogenation of molecules which contain mono and multisubstituted carbon-carbon double bonds. Additionally, chapter 2 indicates that the presence of the polyelectrolyte matrix restricts the access of larger substrates to the catalyst. However, size-based selective diffusion of substrates through the polyelectrolyte matrix only accounts for a small fraction of the 144 total inter and intramolecular selectivity in the hydrogenation of unsaturated alcohols catalyzed by polyelectrolyte-encapsulated Pd NPs. 5.2. Selective Hydrogenation of Unsaturated Alcohols as a Function of NP Diameter Chapter 3 demonstrates that selectivity is a very strong function of Pd NP size in the hydrogenation of unsaturated alcohols. The rate of hydrogenation of allyl alcohol decreases as the NP size increases, but hydrogenation rates for crotyl alcohol and. 2- methyl-2-propen-l-ol increase as NP size increases. Chapter 3 further suggests that these different trends in reactivity for the hydrogenation of unsaturated alcohols are largely due to interactions with specific types of surface atoms on the catalytic NPs. Presumably the hydrogenation of allyl alcohol occurs primarily on defect atoms. Crotyl alcohol and 2- methyl-2-propen—l-ol, on the other hand, might bind to face atoms during the hydrogenation reaction. 5.3. Chemoselective Hydrogenation of Aromatic Nitro Compounds by Pd NPs encapsulated in polyelectrolyte films Chapter 4 extends the catalytic scope of polyelectrolyte-encapsulated Pd NPs to the chemoselective hydrogenation of aromatic nitro compounds that contain other readily reducible functional groups. The nanoparticles show higher selectivities than commercial catalysts such as Pd/C and Pd/alumina. This chapter also demonstrates that the addition of water improves chemoselectivity, even in case of commercially available Pd catalysts. 145 5.4. Future Work 5.4.1. Effect of NP Size on the Chemoselective Hydrogenation of 4- nitrobenzaldehyde In chapter 4, I showed that [PAA-Pd(0)/PEI]3 films are more chemoselective catalysts than commercially available Pd/C or 5% Pd on alumina for the hydrogenation of 4 nitrobenzaldehyde in ethanol. The [PAA-Pd(0)/PEI]3 films in this case were prepared using 15 mM Pd(II) in the PAA deposition solution. Chapter 3 indicates that the average Pd NP diameter in this catalyst is 3.4 nm. It will be interesting to investigate the chemoslectivity for the hydrogenation of 4-nitrobenza1dehyde as a function of NP diameter. This may help to explain why the NP catalysts are more selective than Pd/C or Pd on alumina. 5.4.2. The Contribution of Size-Dependent Electronic Properties to Selective Hydrogenation Chapter 3 suggests that the NP-size-dependent selectivity for hydrogenation of monosubstituted double bonds is primarily due to geometrical effects. However, several studies indicate that in some cases, differences in reaction rates as a function of NP size arise from either electronic or geometrical effects or both."3 As the diameters of NPs decrease, in addition to changes in the fraction of different atom types (e.g., face versus edge), electronic properties of the NPs change due to the transition from metal to insulator to molecule.4'5 This change might contribute to the selective catalytic properties of NPs because interactions of electron rich and poor substrates with NP surfaces may vary. Hydrogenation of nitrobenzene containing electron donating (-CHO, -COCH3), 146 withdrawing (-OH, -NH2), and sterically hindered (-CH3) groups would be an ideal set of reactions for investigating the contribution of changing NP electronic properties to selective hydrogenation 5.4.3. Synergistic Effects Using Bimetallic NPs Bimetallic NPs manifest electronic properties that are uncommon in single-metal NPs, therefore these materials are attractive for developing new catalysts with unique reactivities and selectivities.6'8 A recent report indicates that the rate of hydrogenation of allyl alcohol is enhanced in the presence of the bimetallic dendrimer-encapsulated PdPt NPs compared to only Pt or only Pd NPs supported by the dendrimer.6 Additionally, the enhanced catalytic activity of dendrimer-supported PdAu catalysts in CO oxidation is indicative of a synergistic bimetallic interaction.7 Chemoselective hydrogenations of nitro compounds containing other reducible groups such as olefinic bonds are industrially important reactions. During the reduction of nitostyrene, transition-metal catalysts usually show poor chemoselectivity due to the hydrogenation of the olefinic bond to produce unwanted side products. Halogenated anilines are another industrially important class of compounds due to their presence in pharmaceuticals.9'1 ' Halogenated aniline is often synthesized from halogenated nitro benzene, but dehydrohalogenation is a very common, unwanted side reaction. Sema, et al. showed that titania—supported gold NPs provide selectivity for nitro group reduction in the presence of olefinic bonds or halogens. However, these reactions occur under high temperature and pressure with relatively long reaction times as 147 discussed in chapter 4. Pd NPs show very poor chemoselectivity for the hydrogenation of these nitro compounds, but reactions are very fast at room temperature. Preparation of bimetallic NPs encapsulated in a polyelectrolyte matrix can be easily performed by complexing two types of metal ions with polyelectrolytes during the deposition. Utilization of PdAu bimetallic NPs for chemoselective hydrogenation of nitro groups in the presence of olefinic bonds or other halogen groups under mild condition at short reaction times should again expand the scope of reactions available with NP catalysts. 5.5. Summary Statement This dissertation showed that NPs encapsulated in a polyelectrolyte matrix can serve as highly selective catalysts. I envision that there is enormous potential for significant expansion to a wide range of catalytic reactions. 148 5.6. References (1) Wilson, 0. M.; Knecht, M. R.; Garcia-Martinez, J. C.; Crooks, R. M. Journal of the American Chemical Society 2006, 128, 4510. (2) Bratlie, K. M.; Lee, H.; Komvopoulos, K.; Yang, P.; Somorjai, G. A. Nano Letters 2007, 7, 3097. (3) Zhou Wei, P.; Lewera, A.; Larsen, R.; Masel Rich, 1.; Bagus Paul, S.; Wieckowski, A. Journal of Physical Chemistry. B 2006, 110, 13393. (4) Chen, S.; Murray, R. W.; Feldberg, S. W. Journal of Physical Chemistry B 1998, 102, 9898. (5) Guo, R.; Murray, R. W. Journal of the American Chemical Society 2005, 127, 12140. (6) Scott Robert, W. J.; Datye Abhaya, K.; Crooks, R. M. Journal of the American Chemical Society 2003, 125, 3708. (7) Scott Robert, W. J.; Sivadinarayana, C.; Wilson Orla, M.; Yan, Z.; Goodman, D. W.; Crooks Richard, M. Journal of the American Chemical Society 2005, 12 7, 1380. (8) Wilson, 0. M.; Scott, R. W. J.; Garcia-Martinez, J. C.; Crooks, R. M. Journal of the American Chemical Society 2005, 127, 1015. (9) Jacobsen, M. F.; Moses, J. E.; Adlington, R. M.; Baldwin, J. E. Organic Letters 2005, 7, 641. (10) Brickner, S. J.; Hutchinson, D. K.; Barbachyn, M. R.; Manninen, P. R.; Ulanowicz, D. A.; Garmon, S. A.; Grega, K. C.; Hendges, S. K.; Toops, D. 8.; Ford, C. W.; Zurenko, G. B. Journal of Medicinal Chemistry 1996, 39, 673. (11) Dale, D. J.; Dunn, P. J.; Golightly, C.; Hughes, M. L.; Levett, P. C.; Pearce, A. K.; Searle, P. M.; Ward, G.; Wood, A. S. Organic Process Research & Development 2000, 4, 17. 149 11111111111llllrll ll 11 It 02956 9310