_ as Rea L 1.5 . 5.. A3 i: .‘ xucx, .w. 13...“.“13 5B,. .35. I I :. z? I. x . : i217“. LIBRARY Michigan State University This is to certify that the dissertation entitled HYPERBRANCHED POLYMER FILMS AS ULTRATHIN MEMBRANE SYSTEMS presented by 80 Young Kim has been accepted towards fulfillment of the requirements for Ph.D. degree in Chemistry J" 7,4 ( fl g/km Major professU * Date 0 O a" MS U is an Affirmative Action/Equal Opportunity Institution 0-12771 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 6/01 c:/ClRC/DateDue.p65-p.15 HYPERBRANCHED POLYMER FILMS AS ULTRATHIN MEMBRANE SKINS AND ADHESIVES By Bo Young Kim A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 2002 ABSTRACT HYPERBRANCHED POLYMER FILMS AS ULTRATHIN MEMBRANE SKINS AND ADHESIVES By Bo Young Kim Hyperbranched polymers are attractive materials for forming membranes and thin films because their tree-like structure allows them to cover underlying substrates. Field- emission scanning electron microscopy and atomic force microscopy images clearly show that hyperbranched poly(acry1ic acid) (PAA) films can completely cover permeable substrates without filling underlying pores, thus creating ultrathin membranes on porous supports. The hyperbranched PAA provides an ultrathin, discriminating membrane, while the porous support gives mechanical strength. Gas-transport studies indicate that 3-layer PAA films do not show a high selectivity by themselves, but selectivity improves significantly after covalent derivatization of PAA with H2NCH2(CF2)6CF3. Hyperbranched poly(amidoamine) (PAMAM) dendrimers are another material used for synthesis of hyperbranched membranes. These polymers are well-defined polycations that contain a high charge density at their surface due to multiple amine groups. Thus, dendrimers provide a unique material for depositing films by alternating electrostatic adsorption of polycations and polyanions. The thickness of PAMAM dendrimer/linear poly(acry1ic acid) films varies dramatically with deposition pH, e.g., the thickness of an adsorbed generation 4 (G4) PAMAM dendrimer/PAA “bilayer” can be tuned from <10 A to >2,000 A. Thicknesses of GS PAMAM dendrimer/PAA bilayers can be as high as 4,000 A. The thickest films result when the dendrimer is deposited at pH 8 and PAA is deposited at pH 4 because both polymers are only partially ionized at these pH values. The relatively low charge density on both PAA and dendrimers requires more adsorption to compensate the surface charge. Alternating deposition of 5-bilayer G4 PAMAM dendrimer/PAA films on porous alumina supports yields highly gas permeable membranes, presumably due to the open interior of these molecules. Dendrimer/PAA films can also form microporous polyelectrolyte films through structural changes induced by a reduction in ambient pH. The high density of functional groups at the exterior of dendrimers also makes these materials attractive as possible adhesives. This dissertation shows that a dendrimer interlayer is capable of bonding strongly two PAA-coated surfaces. Specifically, two gold-coated Si wafers were coated with PAA-terminated dendrimer films, and a drop of 10'5 M dendrimer solution was placed between the wafers. Pressing the wafers together then yielded a bond strength of >1.6 MPa. Cross-sectional field-emission scanning electron microscope (FESEM) images show that the films on the two wafers are in intimate contact, suggesting that the thin dendrimer interlayer electrostatically binds the surfaces together. This room-temperature, aqueous-based adhesion may prove useful in applications such as bonding of wafers to form micro-electro-mechanical devices or lab- on-a-chip systems. Overall, these studies demonstrate the versatility and unique character of hyperbranched polymer films. The dense functionality and unusual structure of these systems provide unique opportunities for surface modification, separation, and adhesion. T 0 My Family iv ACKNOWLEDGEMENTS I would like to express my sincere appreciation to Prof. Merlin L. Bruening for his guidance, support, encouragement and mentoring throughout the work presented here. It has been a real pleasure to work with him, and I have certainly learned a great deal during the research in the Bruening lab. I also wish to thank my guidance committee members - Dr. Blanchard, Dr. Borhan, Dr. Geiger for their guidance and constructive comments. I am thankful to my former advisor Dr. Eick for his encouragement and help throughout graduate study. Also a sincere thank-you to all the people in the Bruening group for being “Bruening group family” at some point: Dr. Nagale, Dr. Xiao, Dr. Huang, Jeremy, Anika, Skanth, Sandra, Brian, Dan, Matt, Anagi, Jinhua, Shaoyun, Yingda and Dr. Hong. In the absence of Dr. Nagale’s help for the gas permeability measurements, I would probably still have crosstalk on the ultrathin, hyperbranched Poly(acrylic acid) (PAA) on porous alumina supports. Without Dr. Bi’s wonderfully managed Field-Emission Electron Microscope (FESEM) and Atomic Force Microscope (AF M), none of the micrographs presented here would have been possible. I give thanks to my father, mother, sisters and brother’s family for listening, giving advice, and providing me with perspective. They always encouraged me to continue in the pursuit of my scientific endeavors. I also give sincere thanks to J ineun, Yuna, and Dale for their support and understanding me. TABLE OF CONTENTS List of Tables List of Figures CHAPTER 1. Introduction 1.1 Synthetic Separation Membranes 1.2 Gas Separation Membranes 1.3 Hyperbranched Polymers 1.4 Structure of this Dissertation 1 .5 References CHAPTER 2. Ultrathin Hyperbranched Poly(acrylic acid) Membranes on Porous Alumina Supports 2. 1 Introduction 2.2 Experimental 2.3 Results and Discussion 2.4 Conclusions 2.5 References CHAPTER 3. pH-Dependent Growth and Morphology of Multilayer Dendrimer/Poly(acrylic acid) Films 3. 1 Introduction 3.2 Experimental 3.3 Results and Discussion 3.4 Conclusions 3.5 References CHAPTER 4. The Use of Dendrimer/Poly(acrylic acid) Films as Adhesives 4. 1 Introduction 4.2 Experimental 4.3 Results and Discussion 4.4 Conclusions 4.5 References CHAPTER 5. Conclusions and Future Work vi vii viii l 1 20 22 27 28 32 35 70 71 74 75 76 80 93 94 97 98 99 101 110 111 112 LIST OF TABLES Table 2.1 Ratios of the fluxes of different gases through 3- layer PAA films and fluorinated 3-1ayer PAA films. 68 Table 3.1 Ellipsometric thicknesses of G4 and G8 PAMAM dendrimer (pH=8)/PAA (pH=4) films on gold as a function of the number of bilayers. 84 vii Figure 1.1 Figure 1.2 Figure 1.3 Figure 1.4 Figure 1.5 Figure 1.6 Figure 1.7 Figure 2.1 Figure 2.2 Figure 2.3 Figure 2.4 Figure 2.5(a) LIST OF FIGURES Membrane-based separation. Schematic representation of various membranes. Bisphenol A polysulfone. Polycarbonate. Gas transport through a dense membrane according to the solution-diffusion model. Synthesis of hyperbranched PAA films. Synthesis of PAMAM Dendrimers. Idealized cartoon of a hyperbranched PAA film grafied onto a porous alumina substrate. Substrate thickness is not drawn to scale. Schematic diagram of the synthesis of 1 layer of PAA on a surface. Additional layers were prepared by grafting onto previously deposited PAA. External reflection FTIR spectra of 1, 2, 3, 4, 5, and 6 layers of PAA on porous alumina (0.02 mm pore diameter) that was sputter coated with 63 nm thick gold. Transmission F TIR spectra of a 3-layer PAA film before and afier fluorination. The substrate was gold-coated (5 nm) porous alumina (0.02 um pore diameter). Field-emission scanning electron micrograph (cross- section) of porous alumina (0.02 um pore diameter) before deposition of PAA films. In this image, 5 nm of gold was deposited before PAA growth. viii 13 15 30 31 36 38 42 Figure 2.5(b) Figure 2.5(e) Figure 2.5(d) Figure 2.5(e) Figure 2.5(f) Figure 2.6 Figure 2.7 Field-emission scanning electron micrograph (cross- section) of porous alumina (0.02 pm pore diameter) after deposition of 4-layer PAA films. In this image, 5 nm of gold was deposited before PAA growth. After cleaving the membrane, sample was sputter-coated with 5 nm of gold imaging. F ield-emission scanning electron micrograph (cross- section) of porous alumina (0.02 pm pore diameter) after deposition of 6-layer PAA films. In this image, 5 nm of gold was deposited before PAA growth. After cleaving the membrane, sample was sputter-coated with 5 nm of gold imaging. F ield-emission scanning electron micrograph (cross- section) of porous alumina (0.02 um pore diameter) before deposition of PAA fihns. In this image, 63 nm of gold was deposited before PAA growth. F ield-emission scanning electron micrograph (cross- section) of porous alumina (0.02 pm pore diameter) after deposition of 4-layer PAA films. In this image, 63 nm of gold was deposited before PAA growth. After cleaving the membrane, sample was sputter-coated with 5 nm of gold imaging. Field-emission scanning electron micrograph (cross- section) of porous alumina (0.02 um pore diameter) after deposition of 6-layer PAA fihns. In this image, 63 nm of gold was deposited before PAA growth. After cleaving the membrane, sample was sputter-coated with 5 nm of gold imaging. Field-emission scanning electron micrographs (cross- section) of porous alumina (0.2 pm pore diameter) before and after deposition of 6—PAA layers. 5-nm of gold was deposited before PAA growth, and samples were sputter- coated with 5 nm of gold for imaging after cleaving. Field-emission scanning electron micrograph (cross- section) of a 4-layer PAA films grown of a gold-coated (63 nm) porous alumina membrane (0.02 pm pore diameter). The membrane was imaged without coating with gold after cleavage. ix 43 44 45 46 47 48 49 Figure 2.8(a) Figure 2.8(b) Figure 2.8(c) Figure 2.8(d) Figure 2.8(e) Figure 2.8(f) Figure 2.9(a) Figure 2.9(b) Tapping mode atomic force micrograph of gold-coated porous alumina before grafting of PAA films. The gold coating on 0.02 pm surface pore-diameter- Anopore membranes was 5 nm thick in this image. Tapping mode atomic force micrograph of gold-coated porous alumina after grafting of 4-layer PAA films. The gold coating on 0.02 pm surface pore- diameter-Anopore membranes was 5 nm thick in this image. Tapping mode atomic force micrograph of gold-coated porous alumina after grafting of 6-layer PAA films. The gold coating on 0.02 pm surface pore- diameter—Anopore membranes was 5 nm thick in this image. Tapping mode atomic force micrograph of gold-coated porous alumina before grafting of PAA films. The gold coating on 0.02 um surface pore-diameter- Anopore membranes was 63 nm thick in this image. Tapping mode atomic force micrograph of gold-coated porous alumina after grafting of 4-layer PAA films. The gold coating on 0.02 pm surface pore- diameter-Anopore membranes was 63 nm thick in this image. Tapping mode atomic force micrograph of gold-coated porous alumina after grafting of 6-layer PAA films. The gold coating on 0.02 pm surface pore- diameter-Anopore membranes was 63 nm thick in this image. Tapping mode atomic force micrographs of gold-coated porous alumina before grafting of hyperbranched PAA. The gold coating on 0.2 pm pore-diameter-Anopore membranes was 5 nm thick in this image. Tapping mode atomic force micrographs of gold-coated porous alumina after grafting of 6-layers hyperbranched PAA. The gold coating on 0.2 um pore-diameter-Anopore membranes was 5 nm thick in this image. 52 53 54 55 56 57 58 59 Figure 2.9(c) Figure 2.9(d) Figure 2.10 Figure 2.11 Figure 2.12 Figure 3.1 Figure 3.2 Figure 3.3 Figure 3.4 Tapping mode atomic force micrographs of gold-coated porous alumina before grafting of hyperbranched PAA. The gold coating on 0.2 pm pore-diameter-Anopore membranes was 63 nm thick in this image. Tapping mode atomic force micrographs of gold-coated porous alumina after grafting of 6-1ayers of hyperbranched PAA. The gold coating on 0.2 um pore-diameter-Anopore membranes was 63 nm thick in this image. Optical micrograph of a defect on a porous alumina substrate coated with 63 nm of gold. Flow rates of several gases through a 3-layer PAA film on gold-coated (5 nm) porous alumina (0.02 um pore diameter) as a function of inlet pressure. The membrane area was 2.8 cm2. Permeability values with standard deviations are listed in the text. Flow rates of several gases through a fluorinated 3-layer film on gold—coated (5 nm) porous alumina (0.02 um pore diameter) as a function of inlet pressure. The membrane area was 2.8 cm2. Permeability values with standard deviations are listed in the text. Schematic diagram of the deposition of a dendrimer/PAA bilayer on a gold surface. Repetition of steps 2 and 3 yields multiplayer films. Reflection FTIR spectra of a MFA-modified gold substrate coated with (a) 0.5 (b) 1.5 (c) 2 and (d) 2.5 G4 dendrimer/PAA bilayers. The half bilayer results in the dendrimer being the top layer in the film. Deposition pH values were 8 for G4 dendrimers and 4 for PAA. Ellipsometric thicknesses of 2.5-bilayer G4 dendrimer/PAA films as a function of the pH used for dendrimer and PAA deposition. Standard deviations of thicknesses measured on three different samples ranged from 3 — 35 % of the measured values. Ellipsometric thicknesses of 2.5—bilayer G8 dendrimer/PAA films as a fimction of the pH used for dendrimer and PAA deposition. Standard deviations of thicknesses measured on three different samples ranged from 2-29 % of the measured values. xi 60 61 63 65 66 79 81 82 86 Figure 3.5 Figure 3.6 Figure 3.7 Figure 4.1 Figure 4.2 Figure 4.3 Figure 4.4 Figure 4.5 Optical microscope images of (a) 2.5-bilayer G8 dendrimer (pH=4)/PAA (pH=4) and (b) G8 dendrimer (pH=8)/ PAA (pH=4) films. The ellipsometric thicknesses of the films were 80 A and 4,000 A, respectively. AF M images of 5-bilayer G4 dendrimer (pH=8)/PAA (pH=4) films before (a) and alter (b, c) immersion in a pH 2.5 aqueous solution. For images (a) and (b, polyelectrolytes were deposited in the presence of 0.2 M NaCl, and for image (c), no salt was used during film deposition. Cross-sectional F ESEM images of porous alumina before (a) and after (b) coating with a 4.5-bilayer G4 dendrimer (pH=8)/PAA (pH=4) film. Strategy for wafer bonding using dendrimer/PAA fihns. The inset on the right shows intertwining of dendrimer/PAA bilayers that should occur throughout the film. Ellipsometric thicknesses of dendrimer/PAA films on a base composed of 5 bilayers of PAH/PAA. Deposition pH values were 8 for G4 dendrimers and 4 for PAA. The PAH/PAA base was deposited at pH 4. Cross-sectional field-emission scanning electron microscopy images of (a) 5 dendrimer/PAA bilayers on 5 PAH/PAA bilayers on a gold coated wafer and (b) two wafers coated with similar films that were pressed together with a drop of dendrimer solution between them. Cross—sectional field-emission scanning electron microscopy image of two film-coated wafers that were pressed together with a drop of water between them. The film on the top gold-coated wafer was composed of 5.5 dendrimer/PAA bilayers on 5 PAH/PAA bilayers, while the film on the bottom gold-coated contained 5 dendrimer/PAA bilayers on 5 PAH/PAA bilayers. The tensile strength measurement configuration. A tensile tester is used to pull the bonded wafers apart. xii 88 90 92 103 104 107 108 109 Chapter 1 INTRODUCTION 1.1 Synthetic Separation Membranes Membrane—based separations are important in a wide range of industrial processes including microfiltration,"2 ultrafiltrationf‘4 reverse osmosis,5’6 gas separation7'9 and pervaporationm'l3 Microfiltration and ultrafiltration rely on porous membranes for ”’13 while reverse separation of species such as proteins, viruses, and colloidal particles, osmosis, gas separation and pervaporation utilize dense, nonporous membranes. Perhaps the most well-known application of dense membrane separation is desalination of seawater.14 Despite such successful applications of membrane technology, many membrane-based separations are limited by insufficient selectivity and/or permeability.15 ' '8 Because the permeability of many membrane materials is inversely related to selectivity,7’8 a system with acceptable selectivity often exhibits low flux and vice versa. Thus, synthesis of new membranes with both increased selectivity and high permeability presents a challenge in materials research.16 In its operation, a membrane acts as a selective barrier, or interphase, between two phases that are often homogeneous. Separation occurs when some of the species from a multi-component mixture are transported across the membrane via driving forces such as gradients in concentration (AC), temperature (AT), electric potential (AE), and pressure (AP), while other components are retained by the membrane (Figure 1.1). The success of the separation depends on the selectivity and the flux of the transported species, and these quantities are, in turn, dependent on the physical and chemical properties of the interphase.l4 This dissertation focuses on developing new polymeric materials that are capable of forming ultrathin membranes, because decreasing membrane thickness provides one means of allowing high flux with a high-selectivity material.'9 1 H l“ 1 AC. AP. AT. AE Figure 1.1 Membrane-based separation.l4 Membrane systems are often classified into two categories: symmetric and asymmetric, as shown in Figure 1.2. The symmetric membranes may be further subdivided into three groups: 1) cylindrical porous membranes which are often used for enzyme and DNA separations from dilute solutions,20 2) porous membranes that can be used for microfiltration,"2 and 3) nonporous, homogeneous dense membranes that are normally used for separation of small molecules such as gases.7'8 However, the flux through nonporous, symmetric systems is usually very low because the membrane must be relatively thick to be mechanically stable. Asymmetric membranes can also be subdivided into three groups: 1) porous membranes that have a pore size gradient over the membrane thickness, 2) porous membranes with a dense top layer, and 3) composite membranes that consist of a dense layer or skin on a support prepared from a different material that contains 50 to 150 um- / _ ,I'V“ a"! W. . ,’,' _ .i‘, u '1 a". ‘9" " .‘ ~‘er . . A. '.' .4" q . Porous 5' '~ .- : ' 2. I... 'P I I.‘:__ -»:.o.fi:“‘ .o :L ‘~ “->‘ . ’ Symmetric / \. Homogeneous (nonporous) Membranes Porous \ /' Asymmetric _, Porous with tOPIayer (Integrally skinned) Dense +toplayer <— Porous suppon - , . . Composite Figure 1.2 Schematic representation of various membrane types.l4 diameter pores. In the second category of asymmetric membranes, the system is usually composed of only one material that has an open porous bottom layer, a second layer with smaller pores in the range of 5 - 500 nm and a dense, continuous phase as a top layer. These integrally-skinned, asymmetric membranes are generally prepared by a phase- inversion method that occurs when casting the membrane.21 In both integrally skinned and composite systems, the porous underlying material stabilizes the weak, ultrathin surface films without presenting a substantial barrier to transport. The thin, dense top layer, in contrast, offers high selectivity and some mass transfer resistance, but this resistance is limited due to the minimal thickness of the film. In order to obtain high performance separation membranes with the desired high selectivity and high permeability, the top layer in the composite membranes must be extremely thin because flux is inversely proportional to thickness. Additionally, this layer must be free of defects that decrease membrane selectivity. In chapter 2, we present the development of composite membranes through deposition of hyperbranched poly(acry1ic acid)(PAA) on porous alumina supports. The hyperbranched structure of these fihns allows an ultrathin (< 50 nm) film to cover underlying pores without filling them and thus, provides 3 hi gh-flux membrane. Additionally, these films can be derivatized widely to modulate selectivity. To provide background for that work, we first discuss research and commercial applications of gas separation membranes, then review the solution-diffusion mechanism of gas transport through dense membranes. Subsequently, we review previous work with hyperbranched PAA films as well as dendrimers. The background information on dendrimers is relevant to chapters 3 and 4, which discuss formation of dendrimer/PAA films that are highly gas permeable and the possible use of these films for promoting adhesion. 1.2 Gas-Separation Membranes In 1866, Graham conducted the first studies of membrane-based gas separations and showed that porous membranes can partially separate gas mixtures according to their molecular masses.22 More recently, development of synthetic membranes resulted in industrial gas separations such as removal of C02 from natural gas,23 separation of 02 and N; from air,24 and hydrogen recovery from purge gas in ammonia synthesis.25 One of the most important breakthroughs that allowed practical gas separation actually occurred in the area of reverse osmosis membranes. In the early 1960’s Loeb and Sourirajan developed a process to produce cellulose acetate such that it contained a thin, selective skin on a porous base of the same material.26 The ultrathin skin was absolutely necessary for achieving a reasonable gas flux, and similar membranes were eventually utilized in gas separation.27 Many other membranes based on phase inversion have been developed since the seminal work of Loeb and Souriraj an.”30 Pinnau and Koros developed a dry-wet phase- inversion method for formation of integrally skinned asymmetric membranes that consist of ultrathin skin layers supported by microporous substrates.29 Studies with these membranes along with calculations showed that the gas flux through the substrate (without the skin layer) should be ~10 times higher than that after addition of the skin layer to achieve ~90 % of the gas selectivity of the skin-layer material. In many cases, the resistance of the skin layer can predominantly determine gas transport properties of . 9, asymmetric membranes.2 3‘ Many types of asymmetric membranes have been prepared with skin layers ranging from molecular sieves to polyimides.“32 Suda et al. reported fabrication of asymmetric molecular sieve membranes prepared from polyamic acid membranes via imidization and pyrolysis. The pyrolysis at high temperatures and low heating rates produced the promising gas-selective molecular sieve membranes.” A large number of studies also demonstrate the formation of composite membranes with ultrathin, selective skins.”35 These systems are attractive because only small amounts of the skin material are employed, and thus relatively expensive skin materials can be used. Several studies show that skins with thicknesses of about 50 nm can indeed form the selective skin of a composite membrane. For example, Martin and coworkers employed interfacial polymerization to form poly(aniline) films on the surface of microporous supports. Even though the fihns were only 50 nm thick, they exhibited an Oz/Nz selectivity coefficient of 8.34 Regen et al. even utilized a single monolayer to form the selective skin of a composite membrane. These membranes were fabricated using poly[1-(trimethylsilyl)-1~propyne] (PTMSP) as a support material, a- cyclodextrin/sodium pheonoxide as a gutter layer, and a single monolayer of a modified calix[6]-arene as the primary barrier for transport. This preparation resulted in He/Nz selectivity values as high as 70.35 Industrial applications of polymer membranes in gas separation began in the 1970s at Monsanto. Researchers at Monsanto overcoated asymmetric polysulfone membranes with silicon rubber and showed that the coating could plug defects in the ultrathin, selective layer without decreasing the permeability or selectivity of the skin. This discovery enabled installation of the first large scale gas-separation membrane (Prism®) for the recovery of H2 from ammonia streams. In the 19805, Perrnea introduced their Prism Alpha® membrane for air separation, which was based on asymmetric membranes of bisphenol-A polysulfone (Figure 1.3). General Electric later used ultrathin (< 500 A) silicon/polycarbonate (Figure 1.4) membrane skins for treatment of natural gas to remove impurities such as carbon dioxide, ............ Figure 1.3 Bisphenol A polysulfone 1 I hydrogen sulfide and water. Carbon dioxide and hydrogen sulfide are both corrosive to pipelines, and hydrogen sulfide is very toxic. Moisture must be removed to prevent corrosion and line blackage due to its freezing. At about the same time, Generon used poly(4-methyl-1-pentene) membranes with an Oz/Nz selectivity of about 4. In the 19903, Praxair O l | and Medal produced polyimide OQQ C H 3)2 «QC ‘ C } n membranes with Oz/Nz selectivities of 6 to 8. These Figure 1'4 Polycarbonate membranes could produce 99 % pure nitrogen and gave a cost-competitive alternative to delivered liquid nitrogen for many small users.36 As the above applications show, many successes have been achieved in commercial gas separations. However, on both the commercial and research scale, synthesis of practical membrane skins with thicknesses less than 50 nm is still a challenge. Development of such membranes should enhance flux and increase the versatility of membrane-based separations. We felt that hyperbranched polymers might address this challenge and provide a versatile, derivatizable material for specialized gas or liquid separations. 1.2.1 The Solution-Diffusion Mechanism of Gas Transport Gas separation using nonporous, dense polymeric membranes depends on differences in permeation among the gases of interest, and the mechanism of permeation through such membranes is generally described by the solution-diffusion model.37“40 According to this mechanism (Figure 1.5), permeation of gases proceeds in a three-stage sequence: 1) sorption of the gas at the feed-side membrane surface, 2) diffusion through the membrane, and 3) desorption at the permeate-side membrane surface. Separation of gases is thus based on both solubility and diffusivity factors. The gas-transport properties of a membrane are expressed by two characteristic parameters: 1) the permeability coefficient, P, which is equal to the product of the solubility coefficient (S) and the diffusivity (D) for a specific gas, and 2) the separation factor, a, which is a measure of selectivity of the membrane for one gas over another.14 An understanding of these parameters begins with Fick’s first law, which describes gas . . . . AC . . 41 diffusron due to a concentration gradient, A—X_ , as shown 1n equation 1.1. J=—Dég 1.1 AX In this equation, J is the flux of a specific molecule through the membrane, D is the diffusion coefficient of that molecule in the membrane, AC is the concentration Feed phase Dense Permeate phase membrane 1 Desorption Diffusion 4—-> AX Figure 1.6 Gas transport through a dense membrane according to the solution- diffusion model difference across the membrane, and AX is the membrane thickness. This equation can often be restated as equation 1.2, where C, is the concentration just inside the membrane at the feed side and C, is the concentration in the membrane at the permeate side. The equilibrium concentration C of a specific component at the feed or permeate side of the polymer membrane is often related to its partial pressure, p, and solubility coefficient, S, according to Henry’s Law as shown in equation 1.3. J:D(Co—C,) 1.2 AX Substitution of equation 1.3 into equation 1.2 gives equation 1.4, C=Sxp 13 J=DS(p°_”’) 1.4 AX where p0 is the partial pressure of the gas on the feed side of the membrane and p, is the partial pressure of the gas on the permeate side of the membrane. The product DS is called the permeability, P, and can be expressed for a specific gas according to equation 1.5. JAX po_pl P=DxS= 1.5 The ideal selectivity or of the membrane is defined as the ratio of the individual perrneabilities of a gas pair (a and b) as shown in equation 1.6. P D S crab: “z ”x “ 1.6 Pb Db Sb 10 (I (I and b 5. The ratios are known as the “diffusivity selectivity” and the “solubility selectivity”, respectively. “Diffusivity selectivity” favors transport of the smallest molecule while “solubility selectivity” generally favors the most condensable gas. 1.3 Hyperbranched Polymers Hyperbranched polymers are highly branched macromolecules that provide new 3-dimensional molecular structures that may yield interesting properties for applications 4 ,4 - 4 2 3 enhancement of adhesron, 4 entrapment of such as surface modification, enzymes,45 and chemical sensing.46 These polymers contain an architecture where each subsequent layer contains more polymer chains than the preceding layers, and thus they generally contain an open interior and a relatively crowded periphery. The unique tree- like structure of these molecules could provide at least two potential advantages in membrane formation. First, films may be more capable of covering a porous structure than traditional linear polymers. Second, the open interior of hyperbranched molecules such as dendrimers should provide unique adsorption and diffiision properties. Below, We describe below the specific properties of the hyperbranched polymers employed in this thesis, hyperbranched poly(acry1ic acid) and poly(amidoamine) dendrimers. 1.3.1 Hyperbranched Poly(acrylic acid) Films Hyperbranched PAA fihns are unique candidate for forming ultrathin membranes because their thickness can be controlled over a wide range (2 to 100 nm) by varying the number of surface-grafted PAA layers, and because the carboxylate groups of PAA can be readily derivatized. Using the grafi-on-graft method, Zhou and co-workers first synthesized hyperbranched PAA films on a gold surface.42 In this synthetic procedure, 11 gold slides are first immersed in 0.001 M mercaptoundecanoic acid (MUA) in ethanol (EtOH) to form a self-assembled monolayer (SAM). Activation of the carboxylic acid groups to a mixed anhydride followed by reaction with an u,m-diamino poly(tert- butylacrylate) (HzNR(PTBA)RNH2, R = ((CH2)2NHCO(CH2)2C(CN)(CH3)) yields a grafted layer of PTBA that is specifically synthesized to produce hyperbranched PAA. Hydrolysis of the tert—butyl ester groups of PTBA then yields one grafted PAA layer as shown in Figure 1.6.47 Repeated grafting at multiple —COOH sites on each previous graft leads to hyperbranched PAA films that possess a high density of reactive -COOH functional groups. These functional groups can be derivatized subsequently to introduce specific functionalities into the film. The thickness of these films grows nonlinearly with the number of layers, presumably because each subsequent layer contains more grafts. This procedure allows a 50 mm film to be synthesized with about 5 grafting steps. Although this dissertation focuses on the use of hyperbranched PAA for forming ultrathin, selective membranes, several studies show a variety of potential applications for these films. Lackowski and Crooks reported micron-scale patterning of surfaces using a combination of micro-contact printing of self-assembled monolayers and polymer grafting.48 The three-dimensional nature of the pattern and the fimctionality of the PAA films yielded segregated, chemically sensitive interfaces and corrals for isolating cells.48 Further demonstrating the versatility of hyperbranched PAA, Dermody and co- workers46 prepared gold electrodes coated with hyperbranched PAA that was derivatized with B-cyclodextn'n receptors, and capped with chemically grafted, ultrathin polyamine filter layers. Such films can serve as unique, selective electrochemical sensors. Specifically, a thin, grafted poly(amidoamine) or poly-D-lysine surface layer served as a 12 Number of PAA Layers 1) 000,3 a—co, H N methy' m°m“°""e *Z—CONHR-PTBA-RNHR) p’TSOH'HZO> 2) HZNR-PTBA-RNH2 1 j»CONHR-PAA-RNH, Step 1), then steps 2) and 3) 2 fi—CONHR—{CH2CI3HJn—ECH2CfH)m—zRNHCOZCHZCHg 2 //c\ co,H Nl-lR-PAA-RNH, Step 1), then steps 2) and 3) n2 R2 "‘2' 2 co,H o/ \ NHR PAA-RNH, O\//C \,NHR-[CH ,fiJH]n—[CH, ZCfH}—RNHCOZ CH,CH3 co, H O/ C\NHR-PAA-RNH, Figure 1.6 Synthesis of hyperbranched PAA films.47 13 pH-sensitive molecular filter that allowed selective passage of charged analytes. These studies demonstrate the versatility of hyperbranched PAA, and this versatility could certainly prove useful in membrane development. For example, deposition of specific surface layers could decrease fouling in reverse osmosis or nanofiltration membranes. 1.3.2 Hyperbranched Poly(amidoamine) Dendrimers In contrast to hyperbranched PAA, dendrimers are nearly monodisperse polymers that posess a relatively well—defined molecular architecture.43’49’55 Due to this unique geometry and functionality, many studies have investigated the synthesis, structure, and physical properties of dendrimers. Generally, the synthesis of dendrimers occurs in a step-wise fashion from simple, branched monomer units. There are two synthetic strategies: divergent and convergent. In divergent approaches, the dendritic molecule is built up layer-by-layer from the core to the periphery, while in convergent syntheses, dendrimer components are synthesized initially and then coupled to the dendrimer core. Poly(amidoamine) (PAMAM) dendrimers are perhaps the most common dendritic family, and these molecules are prepared by a divergent scheme as shown in Figure 1.7. In this synthetic approach, each cycle of Michael addition of methyl acrylate (MA) to amines and amidation of the resulting methyl esters with ethylenediamine (EDA) leads to the addition of one more layer of branches, called a generation. Therefore, the generation number of the dendrimer is equal to the number of cycles of Michael addition and amidation performed.49'51’53 One of the unique properties of dendrimers is their ability to serve as molecular “boxes”. Hydrophobic binding,56 hydrogen bonding,5 7metal-1i gand coordination,58 and phys1cal encapsualtron5 can occur in the dendrimer interior, whereas electrostatic l4 +EDA CH3OH HZNTQN O o HNHz «A fj—NH NXN HN-(J ‘—>_H H NH O O NH 2 NH2 1- CH30H +EDA CH3OH NH2 H2N HN / <0 0 NH HszNm 3 H_/-NH2 N NM 0 ‘fi 0 N 0 ”MK 0 N’i-T NXNJ AM: Wit Nit/N ° “ ° Hsz D flxNHz H? o 9 NH NH, tun. Figure 1.7 Synthesis of PAMAM Dendrimers 77 15 52'60'63 and multidentate coordination 6" can take place at the dendrimer interactions surfaces. These properties may dramatically affect the performance of membranes containing these materials. Formation of membranes with dendrimers will require deposition of dendrimer- containing films, and several methods exist for doing this. Wells and Crooks modified MUA-SAMs with amine-terminated poly(amidoamine) PAMAM dendrimers through amide linkages."5 The thickness of these dendrimer monolayers was considerably smaller than the diameter of the dendritic macromolecules in solution, suggesting that dendrimers are compressed along the surface normal and flattened."“”52 Tsukruk reported fabrication of self-assembled dendrimer monolayers by electrostatic deposition, and the thicknesses of these monolayers were also much smaller than the diameter of dendrimers in solution.52 Regen and Watanabe employed amine-terminated PAMAM dendrimers and KthCl4 to construct multilayers by repeated coordination bonding."6 1.3.3 Dendrimers in Alternating Polyelectrolyte Deposition The technique for formation of dendrimer-containing fihns of most relevance to this study is alternating polyelectroyte deposition. This method is a layer-by-layer process that begins with dipping of a charged substrate into a solution containing an oppositely charged polyelectrolyte. After rinsing with deionized water to remove polyelectroyte that is not strongly bound to the substrate, immersion in a solution containing an oppositely charged polyelectrolyte results in adsorption of a second layer on the surface. Upon adsorption of each polyelectroyte, there is charge overcompensation on the surface so that subsequent polymer layers of opposite charge 16 will electrostatically bind to the surface."7 Repetition of the cycle of polyanion and polycation adsorption can continue until the desired numbers of layers are deposited. The structure and composition of multilayered polyelectrolyte films will have a large effect on their properties. Interestingly, although alternating polyelectrolyte deposition occurs via sequential adsorption, multilayer films are not stratified into well- defined layers. Except at the surface of the film, polycations and polyanions are interdigitated to maximize electrostatic interactions,"8 and charges are nearly completely compensated by neighboring polyelctrolytes in these polycation/polyanion complexes. Recent experiments on polyelectrolyte pairs, such as poly(diallydimethylammonium chloride) (PDADMA)/poly(styrene sulfonate) (PSS) films showed essentially complete charge compensation within the bulk of the fihns, i.e., no external ions were present in the films to maintain electrical neutrality. However, excess charge is present at the film surface, and this is essential for film formation.69 In chapter 3 of this dissertation, we examine the pH-dependent grth and morphology of multilayer dendrimer/PAA films prepared by alternating polyelectrolyte deposition. Because dendrimers provide polycations with a unique geometry, films formed from these macromolecules have unusual properties and thicknesses. Dendrimer- containing films can be significantly thicker than films prepared from linear polymers, and the permeability of dendrimer/PAA membranes to gases is extremely high. Additionally, film thickness can be controlled by varying the number of generations in the constituent dendrimers. l7 1.3.4 Dendrimers in Adhesion While working with dendrimer films and membranes, we began to consider their potential for use as adhesives. The multidentate structure of these films may make them particularly well-suited for binding two substrates together. Below, we give some background on organic adhesives and the use of dendrimers in this area. The most commonly used organic adhesives include cyanoacrylates, anaerobic adhesives such as methacrylate esters, and epoxies. a-cyanoacrylate esters polymerize rapidly at room temperature in the presence of surface moisture and then cure to form strong bonds with the substrate.70 Anaerobic adhesives are based on methacrylate esters that are part of a curing system that remains liquid in a plastic bottle that transmits oxygen. The oxygen acts as a radical scavenger and prevents polymerization. In the assembly of metal parts, oxygen is eliminated from the joint, and polymerization occurs.70 Epoxy resins are formed from low molar mass prepolymers containing epoxide end-groups. The diglycidyl ether prepolymers, prepared by reaction of excess epicholrohydrin with bisphenol-A in the presence of a base, usually cured through the use of multi-functional amines that undergo a polyaddition reaction with the terminal epoxide groups.71 The possible mechanisms72 for adhesion include mechanical bonding, chemical bonding and thermodynamic bonding. In mechanical bonding, the two adhering surfaces have surface irregularities that can act as mechanical anchors. Most substrates are rough on the microscopic scale, and the extent of mechanical interlocking is dependent on the rheological properties of the adhesive. Strong mechanical bonding can be obtained on substrates such as wood, paper, and textiles. In chemical bonding, covalent bonds are the 18 primary form of chemical interactions between adhesives and substrates. Secondary interactions involve nonpolar dispersion forces W an der Waals forces), polar dipole interactions, and polar Lewis acid/base interactions. For thermodynamic bonding interaction of the adhesive and substrate should result in a low-energy interface. Generally, a low surface energy adhesive will spread on a higher surface energy substrate to reduce the total surface energy of the system. In addition to the commercially available organic adhesives mentioned above, amine-terminated dendrimer monolayers are useful as adhesion promoters between vapor—deposited Au films and Si-based substrates.44 A recent peel-test study evaluated the strength of adhesion between gold and glass using G2, G4, G6, and G8 PAMAM dendrimer adhesive layers. The G8 adhesion layer gave the strongest connection between the substrate and the Au film because of its large diameter and many surface functionalities. Scanning tunneling microscope images showed that the presence of the dendrimer adhesion layer had no significant impact on the nanostructure of the Au film. In our work, we bond two gold-coated silicon wafers using a G4 dendrimer adhesion layer. We pursued these studies because such films may eventually provide a gentle method of wafer bonding, which is important in the fabrication of micro-electro- mechanical system and lab-on-a-chip devices.”74 There are currently a number of different methods available for silicon wafer-to- wafer bonding, but most of them require relatively harsh conditions.”75 Bonding techniques performed in microstructure fabrication can be divided into three categories. 1) direct bonding: the silicon wafers are directly contacted without an interrnediating layer. The contacted wafers are heated to 800 °C, and Si-O-Si bridging bonds are 19 formed between the surfaces as water molecules are liberated. To obtain bond strengths similar to those in silicon crystals, the heating temperature is increased to 1,000 °C. This method relies on forces that naturally attract surfaces together when they are very smooth and flat. Application of this method is often challenging because of voids between wafers that are created by particles, trapped air or contamination”75 2) Anodic bonding: the bond is formed by applying an electric potential of ZOO-1,000 V at a temperature of 300 - 450 °C. The bonding is typically performed between a sodium-bearing glass wafer and a silicon wafer. One has to ensure that the glass has a good thermal coefficient of expansion fit to silicon.76 3) Intermediate-layer bonding: eutectic bonding involves the deposition of intermediate metallic fihns prior to formation of the bond. An alloy is formed between the wafers by solid-liquid inter-diffusion at the contact interface. Upon cooling the alloy solidifies. In chapter 4 of this dissertation, we present preliminary results that show very strong (>1.6 MPa) adhesion between wafers bonded with dendrimer/PAA layers. This binding method occurs at room temperature and does not require the use of any organic solvents. 1.4 Structure of this Dissertation This chapter of the dissertation describes relevant previous research with membranes and hyperbranched polymeric materials. In the first part, various membranes are presented and compared on the basis of synthetic types. Asymmetric or composite membranes for gas separation, and the mechanism of gas permeation are described. In the second part, the chemical and physical properties of hyperbranched PAA and 20 poly(amidoamine) dendrimers are discussed. Additionally, adhesion is discussed with respect to dendrimer layers. In Chapter 2, work with composite membranes composed of hyperbranched poly(acry1ic acid) grafted on porous alumina substrates is presented. Using F ESEM and AFM, we found that hyperbranched PAA forms defect-free membranes and effectively covers alumina supports without filling underlying pores. Gas-transport measurements with PAA before and after derivatization with H2NCH2(CF2)6CF3 are described and discussed. Chapter 3 details our results on the pH-dependent growth of multilayer films prepared from PAMAM dendrimers and PAA by alternating polyelectrolyte deposition. We show that by systematically varying the pH of deposition solutions and the dendrimer generation, film thickness can be varied over several orders of magnitude. The structures of generation 4 and generation 8 dendrimers as well as solution pH have a profound effect on electrostatic adsorption. Gas permeability measurements with these dendrimer films show that they have a very open structure. Chapter 4 discusses the use of these dendrimer/PAA films in wafer-to-wafer bonding. We inserted a positively charged G4 dendrimer layer between negatively charged PAA layers on each of the two wafers. Mechanical testing revealed that the tensile strength of wafers bonded using the dendrimer/PAA system was >1.6 MPa. Finally, chapter 5 presents general conclusions and future directions in using hyperbranched PAA and dendrimers as membranes, as well as the continued characterization of adhesion resulting from dendrimer/PAA films. 21 1.5 References (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11) (12) (13) (14) (15) (16) (17) (18) Lee, J. D.; Lee, S. H.; Jo, H.; Park, P. K.; Lee, C. H.; Kwak, J. W. Environ. Sci. Techno]. 2000, 34, 3780-3788. Zhang, M.; Song, L. Environ. Sci. T echnol. 2000, 34, 3767-3773. Merchese, J.; Ochoa, N. A.; Pagliero, C.; Almandoz, C. Environ. Sci. T echnol. 2000, 34, 2990-2996. Schuster, B.; Pum, D.; Sara, M.; Braha, O.; Baylay, H.; Sleytr, U. B. Langmuir 2001, 34, 499-503. Alvarez, S.; Riera, F. A.; Alvarez, R.; Coca, J. Ind. Eng. Chem. Res. 2001, 40, 4925-4934. Kwak, S. Y.; Kim, S. S.; Kim, S. H. Environ. Sci. Technol. 2001, 35, 2388-2394. Robeson, L. M.; Burgoyne, W. F.; Langsam, M.; SaVoca, A. C.; Tien, C. F. Polymer 1994, 35, 4970-4978. Robeson, L. M. J. Membr. Sci. 1991, 62, 165-185. Freeman, B. D. Macromolecules 1999, 32, 375-380. Neel, J.; Aptel, P.; Clement, R. Desalination 1985, 53, 297-326. Itoh, T.; Toys, H.; Ishihara, K.; Shimihara, I. J. Appl. Polym. Sci., 1985, 30, 179- 188. Brun, J. P.; Larchet, C.; Melet, M.; Bulvestre, G. J. Membr. Sci. 1985, 23, 257- 283. Krasemann, L.; Tieke, B. J. Membr. Sci. 1998, 150, 23-30. Mulder, M. Basic Principles of Membrane Technology; lst ed.; Kluwer Academic Publishers: Dordrecht, 1996, pp 7-21. Henis, J. M. S.; Tripodi, M. K. Science 1983, 220, 11-17. Haggin, J. Chem. Eng. News 1990, 68, 22-23. Abelson, P. H. Science 1989, 244, 1421. Spillman, R. W. Chem. Eng. News 1989, 68, 11-17. 22 (19) (20) (21) (22) (23) (24) (25) (26) (27) (28) (29) (30) (31) (32) (33) (34) (35) (36) Nagale, M.; Kim, B. Y.; Bruening, M. L. J. Am. Chem. Soc. 2000, 122, 11670- 11678. Lvov, Y.; Decher, G.; Sukhorukov, G. Macromolecules 1993, 26, 5396-5399. Suda, H.; Haraya, K. In Membrane Formation and Modification; Pinnau, I., Freeman, B. D., Eds; ACS Symposium Series 744, American Chemical Society: Washington, DC, 2000, pp 295-313. Graham, T. Philos. Mag. 1866, 32, 401-420. Mazur, W. H.; Chan, M. C. Chem. Eng. Prog. 1982, 78, 38-40. Matson, S. L.; Ward, W. J .; Kimura, S. G.; Browall, W. R. J. Membr. Sci. 1986, 29, 79-96. Spillman, R. W. Chem. Eng. Prog. 1989, 85, 41-62. Loeb, S.; Sourirajan, S. Adv. Chem. Ser. 1962, 38, 117. Gantzel, P. K.; Merten, U. 1&EC. Proc. Des. Dev. 1970, 9, 331. Reuvers, A. J .; van den Berg, J. W. A.; Smith, C. A. J. Membr. Sci. 1987, 34, 45- 86. Pinnau, I.; Koros, W. J. Ind. Eng. Chem. Res. 1991, 30, 1837-1840. Kesting, R. E. In Materials Science of Synthetic Membranes; Lloyd, D. R., Eds.; ACS Symposium Series 269, American Chemical Society: Washington, DC, 1985, pp 131-163. Strathmann, H.; Scheible, R; Baker, R. W. J. Appl. Polym. Sci, 1971, 15, 811- 828. Kim, T. H.; Koros, W. J .; Husk, G. R.; O'Brien, K. C. J. Membr. Sci. 1988, 37, 45-62. Henis, J. M. S.; Tripodi, M. K. J. Membr. Sci. 1981, 8, 233-246. Liu, C.; Martin, C. R. Nature 1991, 352, 50-52. Zhang, L. H.; Hendel, R. A.; Cozzi, P. G.; Regen, S. L. J. Am. Chem. Soc. 1999, 121,1621-1622. Baker, R. W. Ind. Eng. Chem. Res. 2002, 41, 1393-1411. 23 (37) (38) (39) (40) (41) (42) (43) (44) (45) (46) (47) (48) (49) (50) (51) (52) (53) (54) Frisch, H. L.; Stem, S. A. CRC Crit. Rev. Solid State Mater. Sci. 1983, 11, 123- 187. Stem, S. A.; Frisch, H. L. Annu. Rev. Mater. Sci. 1981, 11, 523-550. Crank, J. The Mathematics of Diffusion; 2nd ed.; Clarendon: Oxford, 1975. Matson, S. L.; Lopez, J .; Quin, J. A. Chem. Eng. Sci. 1983, 38, 503-524. Pick, A. Ann. Phys. Chem. 1955, 94, 59. Bruening, M. L.; Zhou, Y.; Aguilar, G.; Agee, R.; Bergbreiter, D. E.; Crooks, R. M. Langmuir 1997, 13, 770-778. Fréchet, J. M. J. Science 1994, 263, 1710-1715. Baker, L. A.; Zamborini, F. P.; Sun, L.; Crooks, R. M. Anal. Chem. 1999, 71, 4403-4406. F ranchina, J. G.; Lackowski, W. M.; Dermody, D. L.; Crooks, R. M.; Bergbreiter, D. E. Anal. Chem. 1999, 71, 3133-3139. Dermody, D. L.; Peez, R. F.; Bergbreiter, D. E.; Crooks, R. M. Langmuir 1999, 15, 885-890. Zhou, Y.; Bruening, M. L.; Bergbreiter, D. E.; Crooks, R. M.; Wells, M. J. Am. Chem. Soc. 1996, 118, 3773-3774. Lackowski, W.; Ghosh, P.; Crooks, R. M. J. Am. Chem. Soc. 1999, 121 , 1419- 1420. Tomalia, D. A.; Naylor, A. M.; Goddard, III. W. A. Angew. Chem. Int. Ed. Engl 1990, 29, 138-140. Newkome, G. R.; Moorfield, C. N.; Baker, G. R.; Saunders. M. J .; Grossman, S. H. Angew. Chem, Int. Ed. Eng. 1991, 30, 1178-1180. Zeng, F.; Zimmerman, S. C. Chem. Rev. 1997, 97, 1681-1712. Tsukruk, V. V. Adv. Mater 1998, 10, 253-257. Tomalia, D. A.; Hedstrand, D. M.; Ferritte, M. S. Macromolecules 1991, 24, 1435-1438. Saville, P. M.; Reynolds, P. A.; White, W. W.; Hawker, C. J .; Fréchet, J. M. J .; Wooley, K. L.; Penfold, J.; Webster, J. R. P. J. Phys. Chem. 1995, 99, 8283-8289. 24 (55) (56) (57) (58) (59) (60) (61) (62) (63) (64) (65) (66) (67) (68) (69) (70) (71) (72) Alonso, B.; Moran, M.; Casado, C. M.; Lobete, F.; Losada, J.; Cuadurado, I. Chem. Mater. 1995, 7, 1440-1442. Naylor, A. M.; Goddard, 111. W. A.; Kiefer, G. E.; Tomalia, D. A. J. Am. Chem. Soc. 1989, 111, 2339-2341. Newkome, G. R.; Guther, R.; Moorefield, C. N.; Cardullo, F .; Echegoyen, L.; Perezcordero, E.; Luftmann, H. Angew. Chem, Int. Ed. Eng. 1995, 34, 2023- 2026. Sttodart, J. F.; Philip, D. Angew. Chem, Int. Ed. Engl. 1996, 35, 1154-1156. Jansen, J. F . G. A.; de Brabander-Van Den Berg, E. M. M.; Meijer, E. W. Science 1994, 266, 1226-1229. He, J. A.; Valluzzi, R.; Yang, K.; Dolukhanyan, T.; Sung, C. M.; Kumar, J.; Tripathy, S. K.; Samuelson, L.; Balogh, L.; Tomalia, D. A. Chem. Mater. 1999, 11, 3268-3274. Khopade, A. J.; Caruso, F. Nano. Lett. 2002, 2, 415-418. Cheng, L.; Cox, J. A. Electrochem. Commun. 2001, 3, 285-289. Yoon, H. C.; Kim, H. S. Anal. Chem. 2000, 72, 922-926. Posnett, D. N.; Mcgrath, H.; Tam, J. P. J. Biol. Chem. 1988, 263, 1719-1725. Wells, M.; Crooks, R. M. J. Am. Chem. Soc. 1996, 118, 3998-3999. Watanabe, S.; Regen, S. L. J. Am. Chem. Soc. 1994, 116, 8855-8856. Schlenoff, J. B.; Dubas, S. T. Macromolecules 2001, 34, 592-598. Dubas, S. T.; Schlenoff, J. B. Macromolecules 1999, 32, 8153-8160. Schlenoff, J. B.; Ly, H.; Li, M. J. Am. Chem. Soc. 1998, 120, 7626-7634. Lee, L. H. Adhesion Science and Technology; Plenum Press: New York and London, 1975; Vol. 98; pp 467-475. Young, R. J.; Lovell, P. A. Introduction to Polymers; 2nd Edition ed.; Chapman & Hall: London, 1992. Kalantar, J .; MS Thesis; Michigan State University: East Lansing, MI, 1988. 25 (73) (74) (75) (76) (77) Baldwin, R. P.; Roussel, T. J .; Crain, M. M.; Bathlagunda, V.; Jackson, D. J .; Gullapalli, J.; Conklin, J. A.; Rekha, P.; Naber, J. F.; Walsh, K. M.; Keynton, R. S. Anal. Chem. 2002, 74, 3690-3697. Schmidt, M. A. Proceedings of the IEEE. 1998, 86, 1575-1585. Bengtsson, S.; Amirfeiz, P. J. Electronic Mater. 2000, 29, 909-915. Choi, W. B.; Ju, B. K.; Lee, Y. H.; Haskard, M. R.; Oh, M. H. J. Vac. Sci. Techno]. B 1997, 15, 477-481. Uppuluri, S.; Keinath, S. E.; Tomalia, D. A.; Dvomic, P. R. Macromolecules 1998, 31, 4498-4510 26 Chapter 2 ULTRATHIN, HYPERBRANCHED POLY(ACRYLIC ACID) MEMBRANES ON POROUS ALUMINA SUPPORTS“ Summary The synthesis of hi gh-flux composite membranes requires deposition of an ultrathin, discriminating layer on a highly permeable support. This chapter describes synthesis, derivatization, and characterization of hyperbranched poly(acry1ic acid) (PAA) membranes on porous alumina supports. PAA films as thin as ~ 40 nm effectively cover underlying pores without filling them, presumably because of their hyperbranched structure. Synthesis of these films begins by sputtering a thin gold layer on the alumina support and then grafting a layer of PAA to a self-assembled monolayer of mercaptoundecanoic acid on the gold. Graft-on-graft deposition of PAA yields the hyperbranched membrane. FESEM (field-emission scanning electron microscopy) and AFM (atomic force microscopy) images clearly show that hyperbranched PAA films can cover the substrate surface completely without filling underlying pores, thus creating an ultrathin membrane on a porous support. PAA membranes are especially attractive because derivatization permits control over transport properties. Gas-transport studies indicate that 3-layer PAA films do not show a high selectivity by themselves, but selectivity improves significantly after covalent derivatization of PAA with HzNCH2(CFz)6CF3. ” This chapter is taken from a publication by Milind Nagale, Bo Young Kim, and Merlin L. Bruening (J. Am. Chem. Soc. 2000, 122, 11670-11678), and is a joint work of the three authors. 27 2.1 Introduction Although membrane separations are attractive because of low energy costs and simple operation, low perrneabilities and/or selectivities often limit membrane applicationsl'4 Successful commercial applications of membrane separations already include production of nitrogen from air and recovery of hydrogen from mixtures with other larger components like nitrogen, methane, and carbon monoxide.5 Improvements in throughput and selectivity, however, could greatly increase the impact of membrane separations.6 One major limitation of gas-separation membranes is that selective materials are generally not highly permeable.7'9 This tradeoff between selectivity and permeability makes decreasing membrane thickness vital for increasing flux without sacrificing selectivity-1,10,“ As shown in Equation 2.1, gas flux through a membrane (F) is inversely proportional to membrane thickness (1) and directly proportional to the pressure gradient across the membrane (Ap) and the permeability coefficient (P) for a specific I? a. gas. F = —— 2.1 The mechanical weakness of ultrathin films necessitates their deposition on highly permeable supports that provide strength. Unfortunately, synthesis of defect-free, ultrathin (<50 nm) membrane “skins” is challenging.“ 1'13 There are several approaches to preparing membranes with ultrathin skins. One strategy involves phase-inversion to prepare an asymmetric membrane from a single material. The first asymmetric gas separation membrane was a Loeb-Sourirajan-type 28 cellulose acetate (CA) material formed by drying CA membranes using quick-freezing . . l4 . . . and vacuum sublrmatron at —10 °C. More recent efforts include synthesrs of integrally . . . . r3 , skinned membranes usmg a dry/wet-phase 1nversron process. Asymmetric carbon molecular sieve membranes can even be prepared by formation of a capillary type polyamic acid membrane followed by imidization and pyrolysis to form a dense selective 15 layer supported by a porous layer. Composite membranes provide a versatile alternative to asymmetric skinned systems. In this case a polymer “skin” is deposited on a separate, highly permeable support. Composite membranes have the advantage that only a small amount of the possibly expensive “skin” material needs to be used. Strategies for forming the “skins” . . 16.17 . . . 18 . . . of these membranes include casting, 1n srtu castmg, 1n Situ condensatron of polymers and/or monomers at the porous support,19 and plasma polymerization.20 In all of these methods, synthesis of defect-free “skins” with thicknesses < 50 nm is difficult. A few recent reports demonstrate that composite membranes with ultrathin “skins” are capable of selective separations. With the thinnest synthetic gas-separation membranes to date, Regen and co-workers showed that Langmuir-Blodgett fihns as thin as a single monolayer can serve as selective membranes when deposited on a continuous 21-23 , . . . . . glassy polymer support. L1u and Martin reported fabrication of ultrathin, conducting r polymer membranes (~50 nm thick) with an Oz/Nz selectivity of 8.l In this case interfacial photopolymerization at the alumina surface resulted in an ultrathin conducting polymer membrane. Similar selectivities can be achieved by electrochemical synthesis of . . 24 . . ultrathin films on porous alumina membranes. Thin film composrte membranes can also be formed on hollow fibers. Paul et a]. formed multilayered composite hollow-fiber 29 membranes by dip-coating hollow-fiber polysulfone supports with a selective poly(4- vinylpyridine) layer (50-150 mm thick) and sealing with a silicone rubber layer. This system has a selectivity of 7 for Oz/Nz.25 We report fabrication of surface-grafted, hyperbranched poly(acry1ic acid) (PAA) membranes. These films are unique in that their highly branched structure allows them to cover underlying pores without filling them, as shown schematically in Figure 2.1.215 <— 3rd layer <— 2nd layer <— 1st layer 0 is an amide bond Solid Substrate Figure 2.1 Idealized cartoon of a hyperbranched PAA film grafted onto a porous alumina substrate. Substrate thickness is not drawn to scale. Additionally, in situ derivatization of these materials allows synthesis of a variety of membranes without the need to prepare new polymers. The synthesis (Figure 2.2) involves first forming a self-assembled monolayer of a carboxylic acid-terminated thiol on a gold-coated porous substrate followed by the conversion of carboxylates to mixed anhydrides. These mixed anhydrides are then used as attachment points for the grafting of amine-terminated poly(tert-butyl acrylate) (PTBA) onto the surface. The hydrolysis of tert-butyl groups to carboxylic acid groups forms a PAA layer, and the whole process is 30 ~2‘~Z‘~Z‘~Z‘ O O O O CICOZCHZCH3 N-methyl morpholine NH2R-PTBA-RNH2 w w is» Irv s s e a r t s 5 <5 .3 r 5 Q“ Q” 0 $0 5 050 gr p-TsOH-HZO, 55°C g or 5 gr PAA= tQH-CHzi-n ‘ PTBA: -[-|CH-CH2-]—n (3:0 C|=O OH 0+ Figure 2.2 Schematic diagram of the synthesis of 1 layer of PAA on a surface. Additional layers were prepared by grafting onto previously deposited PAA. 31 then repeated to graft additional layers to the immobilized PAA. The layer-by-layer synthesis of hyperbranched PAA affords control over film thickness, while derivatization of the films provides a way to control transport properties chemically.26 Derivatization occurs through amidation or esterification of the —COOH groups of PAA with functional amines or alcohols, and a prior study shows that PAA films can be modified to give fluorescent, hydrophobic, ion-binding, biocompatible, and electroactive films.27 Thus, derivatization can provide a wide variety of hyperbranched membranes for investigating relationships between transport and membrane chemistry. This paper focuses on derivatization of hyperbranched PAA with HzNCH2(CF2)6CF3 because this procedure produces selective gas-separation membranes. The occurrence of gas selectivity shows that these membranes are free of defects. 2.2 Experimental Materials. Porous alumina (0.02 and 0.2 um-pore diameter AnoporeTM filters) substrates were purchased from Thomas Scientific. Ethyl chlorofonnate (97%), N- methyl morpholine (99%), p-toluene sulfonic acid monohydrate (98%), 4,4’-Azobis(4- cyanovaleric acid) (75%+), 1,4-dioxane (99%), and tert-butyl acrylate (98%) were purchased from Aldrich. H2NCH2(CF2)6CF3 (97%) was purchased from Lancaster. Acetonitrile (Spectrum, 99.5%), and ethanol (Pharmco, 100%) were used as received. DMF (Aldrich, 99.9%) was dried with molecular sieves for 24 h before use, and deionized water (18.2 MQmm) was obtained using a Milli-Q system. Mercaptoundecanoic acid (MUA) was initially synthesized with a modified literature procedure,28 but most of the MUA used in this research was purchased from Aldrich. 32 a,u)-diamino-terminated poly(tert-butyl acrylate) (H2NR-PTBA-RNH2), R = (CH2)2NHCO(CH2)2C(CN)(CH3), was prepared according to a literature procedure?“27 He (99.995%), N2 (99.99%), 02 (99.99%), CO2 (99.8%), CH, (99.3%), and SF6 (99.99%) were purchased from AGA. Synthesis of PAA films. Alumina substrates were cleaned for 10 min in a UV/ozone cleaner (Boekel Industries, model 135500) after a 10 min immersion in boiling methanol. Subsequently, gold (63 nm or 5 nm) was sputtered (Ted Pella, Pelco SC-7) onto the substrates. Sputtering was performed with a current of 20 mA and a pressure of ~ 0.08 mbar (Ar plasma) at a rate of 1 A/s. Gold-coated porous alumina substrates were cleaned in the UV/ozone cleaner for 12 min before beginning monolayer deposition. To synthesize PAA films, we grafted H2NR-PTBA-RNH2 via amide formation onto a mercaptoundecanoic acid (MUA) monolayer attached to gold-coated porous alumina. Hydrolysis of tert-butyl ester groups with p-toluenesulfonic acid results in a grafted PAA film. Repeating these steps using additional grafting at multiple sites on each prior graft produced layered, hyperbranched polymer membranes. This procedure was described . 26 prevrously. Derivatization of 3-layer PAA films. After synthesis of 3-layer PAA membranes and measurement of their gas permeability, these films were derivatized with H2NCH2(CF2)6CF3. To derivatize PAA membranes, the film was first activated with ethyl chloroforrnate in the presence of N—methylmorpholine to form mixed anhydrides. The film was rinsed with ethyl acetate, dried with N2, soaked in a 0.1 H2NCH2(CF2)6CF3 solution in DMF for 1 hr, rinsed with ethanol, and dried with nitrogen.29 33 Fourier Transform Infrared Spectroscopy. External reflection F TIR spectra of PAA films on substrates coated with 63 nm of gold were acquired using a Nicolet FTIR spectrometer (MAGNA 560) equipped with a MCT detector and a PIKE grazing angle accessory (incident angle of 80°, 256 scans at 4 cm'1 resolution). Transmission FTIR spectroscopy (Mattson Instruments, Infinity Gold) was employed to monitor the growth of hyperbranched PAA on alumina supports coated with only 5 nm of gold because these substrates are not sufficiently reflective for external reflection F TIR. Field-Emission Scanning Electron Microscopy (FESEM). Electron microscope images of membranes were obtained using a Hitachi S-47OOII field-emission scanning electron microsc0pe. To prepare membrane cross-sections for imaging, samples were cleaved using tweezers, sputter-coated with 5 nm of gold, rinsed with methanol, and dried using nitrogen unless otherwise noted. In the cleaving process, a membrane was held by one pair of tweezers and bent with another pair to break the alumina support. The polymer support ring around the edge of the alumina was then cut away using a pair of scissors. During imaging, a low accelerating voltage (4 kV; beam current, 10 uA) was . . . . . . 30.31 used to minimize the effects of sample chargrng and to provrde better surface detafls. Atomic Force Microscopy (AFM). AF M images were obtained with a Nanoscope IIIa (Digital instruments) instrument using the tapping mode (scan rate = 1 Hz). A cantilever having a nominal spring constant of 20-100 N/m was used along with etched silicon tips. The tips have a nominal radius of curvature of 20-60 nm. Membranes could be directly imaged by AF M without any special preparation. Gas-Transport Measurements. Gas-transport measurements were performed using a permeation cell (MKS Instruments, cell 02910-40) equipped with a pressure relief 34 valve. Permeate flux was measured as a function of inlet pressure (10-40 psig) using a digital soap bubble flowmeter (Fisher scientific, Model 420). The area of the membrane exposed to the gas stream was 2.8 cm2. The perrneabilities of He, N2, O2, CO2, CH4 and SF (2 were determined for each sample. Three different membranes of each type were tested with at least three steady state measurements of flux for each gas at a given pressure. The order in which the permeability of different gases was determined was deliberately varied, and each sample was tested at least twice for the entire set of gases to check the stability of the membrane. The cell was purged several times with the gas of interest at a pressure of 50 psig using the pressure relief valve before performing any measurements with a particular gas. All measurements were obtained after the flux reached a steady-state value to ensure complete purging of the cell. After permeability measurements of membranes coated with PAA layers, samples were removed, analyzed by transmission FTIR, and fluorinated as described in the synthesis section. Permeability of fluorinated PAA layers was then measured. 2.3 Results and Discussion Spectroscopic Characterization of PAA Membranes. Hyperbranched PAA films with up to 6 layers were synthesized on gold-coated porous alumina as shown in Figure 2.2, and the synthetic procedure was monitored by FTIR spectroscopy. In agreement with previous syntheses of hyperbranched PAA on gold-coated silicon wafers,26 external reflection FTIR spectra confirmed the steps of the synthetic process, including formation of a mixed anhydride, attachment of H2NR-PTBA-RNH2 to the surface, and hydrolysis of tert-butyl ester groups. External reflection F TIR spectra of l 35 Ions I coo" a: O C a -E o J m .a < 6PAA 5PAA 3““ A ‘— ._12AA \: \ l I \ 2000 1800 1600 1400 1200 Wavenumber(cm1) Figure 2.3 External reflection F TIR spectra of l, 2, 3, 4, 5, and 6 layers of PAA on porous alumina (0.02 pm pore diameter) that was sputter coated with 63 nm of gold. 36 to 6 layers of PAA grafted onto gold-coated (63 nm) porous alumina indicated that the absorbance due to the C=O band of PAA increases nonlinearly as a function of the number of layers, which suggests hyperbranched growth of these films (Figure 2.3). We estimated the thickness of PAA films on porous alumina by comparing their IR absorbances with absorbances due to films of known ellipsometric thickness on gold wafers. Using this method, the estimated thickness of a 6-layer PAA fihn on porous alumina is about 940 A.32 This compares reasonably well with FESEM images, which show 6-1ayer PAA films to be about 850 A thick. Because alumina substrates are highly porous, direct ellipsometric measurements are not possible on these membranes. We utilized transmission FTIR spectroscopy (Figure 2.4) in an effort to investigate the synthesis of PAA films on porous alumina coated with 5 nm of gold because these surfaces are not sufficiently reflective for external reflection FTIR spectroscopy. Large carbonyl absorbances in the transmission FTIR spectra of PAA- coated supports (absorbance of ~1 .0 at 1,712 cm'1 for 6-layer films) suggest that considerable PAA is adsorbed inside the alumina pores. We confirmed this by control experiments where PAA films were grown with and without MUA monolayers on gold- coated (5 nm) alumina membranes. Absorbance values are comparable for the two cases, which indicates that growth inside the pores is the primary reason for the large IR absorbances. It is not surprising that some PAA would be present in substrate pores because a minimal amount of adsorption will be amplified by hyperbranched growth. During the deposition of PTBA, some physisorption likely occurs. After hydrolysis, interaction of the —COO‘ groups of PAA with alumina may result in even stronger physisorption. The 37 9-05 Amide u / (D O C B 5 COOH U) .0 <1: 3—PAA l l l 1800 1600 1400 Energy (cm'1) Figure 2.4 Transmission FTIR spectra of a 3-layer PAA film before and alter fluorination. The substrate was gold-coated (5 nm) porous alumina (0.02 um pore diameter). 38 high internal surface area of the alumina supports leads to large carbonyl absorbances even when only a small fraction of the pores is filled with PAA. Comparison of IR absorbances of PAA films grown on alumina supports to absorbances of known amounts of PAA physisorbed on alumina supports indicates that there is 1.1 i 0.4 mg of PAA distributed through alumina supports after the synthesis of a 6-layer PAA film. The number of ~0.2 rim-diameter pores in the bulk of porous alumina is 2 x 109/cm2 (determined from SEM images), giving a total internal surface area of 1800 cm2 for these 50 um-thick membranes. If we assume that the PAA is distributed evenly across the internal surface area, there would be a 6 i 2 nm-thick film along the pore walls. This would reduce the inside diameter of pores by < 10%. For 3-layer PAA films, the pore diameter would hardly be affected. FESEM images and gas-permeability studies (vide infra) clearly show that PAA is not filling a substantial fraction of membrane pores. Transmission F TIR spectroscopy confirms modification of 3-layer PAA films with H2NCH2(CF2)6CF3 (Figure 2.4). After derivatization, the acid carbonyl peak (1,712 cm") diminishes and amide I (1,669 cm") and amide 11 (1,537 cm") bands appear. Additionally, a CFx stretching peak appears at 1,254 cm“l . Other CFx stretches cannot be observed because of the large absorbance of alumina below 1,210 cm’l. Although these spectra represent material both within the alumina and on its surface, similar chemistry should occur in both places. Field-Emission Scanning Electron Microscopy (FESEM). Cross-sectional FESEM images of alumina supports before and after PAA growth indicate that PAA films cover substrate pores without filling them. Figure 2.5 (a-f) shows FESEM images of porous alumina before and after coating with 4- and 6-layer PAA films. In the case of 39 6-layer deposition, the PAA is easily observed, and yet, pores are unclogged. Because of their small thickness, 4-1ayer PAA films are not easy to see. The large increase in film thickness on going from four to six layers confirms the non-linear PAA growth shown by FTIR. Prior to film synthesis, we sputter-coated the alumina with either 5 nm (a-c) or 63 nm (d-f) of gold. As seen by comparing images with no PAA films (a, d), the 63 nm- gold layer is clearly distinguishable from the membrane and has a porous structure that does not form an impermeable barrier. All of the images in Figure 2.5 (a-f) were taken on substrates that have a nominal 0.02 pm pore diameter. These substrates have a very thin cake layer on their surface with 0.02 rim-diameter pores, while the bulk of the membrane has 0.2 um-diameter pores. This substrate geometry makes it easier to cover pores without filling them. However, images of substrates with 0.2 um-diameter pores (no cake layer) show that PAA films cover even these large pores (Figure 2.6). A PAA film is clearly visible after deposition of 6 layers on alumina coated with 5 nm of gold. Despite the larger pores, deposition is restricted primarily to the substrate surface. In an effort to ensure that gold coating after cleavage does not affect the electron micrographs, we also imaged a sample covered with 4-layers of PAA (0.02 um-diameter pores coated with 63 nm gold) without coating with gold alter cleavage (Figure 2.7). The image shows the presence of PAA and confirms that the film is, in fact, PAA and not an artifact of sample preparation. FESEM images (cross—sections) of other types of control samples on which we attempted to graft 6-layer PAA and fluorinated 6-layer PAA films on gold-coated (5 nm) alumina without depositing a MUA monolayer do not show significant film growth on the alumina surface. This is consistent with gas permeability 40 data that show minimal inhibition of flux for membranes prepared without first depositing a layer of MUA. 41 llllrlvllllll 4.0kvig&7mm x70.01< SE(U) 1.3/28199 12:17 429nm Figure 2.5 (a). Field-emission scanning electron micrograph (cross-section) of porous alumina (0.02 pm surface pore diameter) before deposition of PAA films. In this image, 5 nm of gold was deposited before PAA growth. 42 PAA Figure 2.5 (b). Field-emission scanning electron micrograph (cross-section) of porous alumina (0.02 pm surface pore diameter) after deposition of 4-layer PAA films. In this image, 5 nm of gold was deposited before PAA growth. After cleaving the membrane, sample was sputter-coated with 5 nm of gold for imaging. 43 PAA :1 “ 1% .1 9 1111111111 42 ‘1 1 mm x70.0k SE[U) 811/99 21 :09 1 I I Figure 2.5 (c). Field-emission scanning electron micrograph (cross—section) of porous alumina (0.02 pm surface pore diameter) after deposition of 6-layer PAA films. In this image, 5 nm of gold was deposited before PAA growth. Alter cleaving the membrane, sample was sputter-coated with 5 nm of gold for imaging. lllllll’llll 4.0kV 12.3mm x70.Dk seru) 5125299 16:21 Figure 2.5 (d). Field-emission scanning electron micrograph (cross-section) of porous alumina (0.02 pm surface pore diameter) before deposition of PAA films. In this image, 63 nm of gold was deposited before PAA growth. 45 PAA Gold 32'. 1. a u ‘2 - ti 3 111111.111! »428nm Figure 2.5 (e). F ield-emission scanning electron micrograph (cross-section) of porous alumina (0.02 pm surface pore diameter) after deposition of 4-layer PAA films. In this image, 63 nm of gold was deposited before PAA growth. After cleaving the membrane, sample was sputter-coated with 5 nm of gold for imaging. 46 Figure 2.5 (f). Field-emission scanning electron micrograph (cross-section) of porous alumina (0.02 pm surface pore diameter) after deposition of 6-layer PAA films. In this image, 63 nm of gold was deposited before PAA growth. After cleaving the membrane, sample was sputter-coated with 5 nm of gold for imaging. 47 Figure 2.6. Field-emission scanning electron micrographs (cross-section) of porous alumina (0.2 pm pore diameter) before and after deposition of 6-PAA layers. 5-nm of gold was deposited before PAA growth, and samples were sputter-coated with 5 nm of gold for imaging alter cleaving. 48 Figure 2.7. Field-emission scanning electron micrograph (cross-section) of a 4- layer PAA film grown on a gold-coated (63 nm) porous alumina membrane (0.02 pm surface pore diameter). The membrane was imaged without coating with gold after cleavage. 49 Atomic Force Microscopy (AF M). AF M images show that 6-layer PAA films cover the porous alumina surface completely. Figure 2.8 contains images of gold-coated porous alumina substrates before and after grafting of 4- and 6-layers of hyperbranched PAA to the surface. With the addition of successive PAA layers, surface features of the porous alumina become less distinct as the film covers the substrate. Prior to PAA coating, images (a and (I) show the pointed surface features of gold-coated alumina substrates. These images are similar to naked porous alumina. Interestingly, the sputtered gold appears to follow the pattern of the underlying substrate even when 63 nm of gold is deposited. This explains why the cross-sectional FESEM images show that thick gold coatings are porous. The pores in the alumina substrates most likely reside between the peaks on the surface. This topology allows hyperbranched PAA to effectively cover pores because the polymer can grow above the pore from all sides. As the images show, covering the underlying alumina with PAA films greatly reduces surface roughness. To quantitate the changes in roughness, we used the instrument software to calculate the average surface roughness. This average roughness, R,, is defined by equation 2.2 where Z, is the deviation between the surface height at any Ra : {7 211211 2.2 point and the average surface height, and N is the number of points within the measured area. The surface of the alumina coated with 5 nm of gold has an average roughness of 40 i 10 run, while the surface of the alumina coated with 63 nm of gold has an average roughness of 19 i 2 nm. In contrast, 6-layer PAA surfaces on 5-nm gold-coated supports 50 have average roughnesses of 7 i 1 nm, and similar films on substrates coated with 63 nm of gold have average roughnesses of 6 i 2 nm. These measurements show that 6-layer PAA films form a smooth surface that covers underlying sharp features. Figure 2.9 shows AF M images of PAA films that were prepared on gold-coated substrates with 0.2 rim-diameter surface pores. Pores in the substrate before coating with PAA (a, c) are an order of magnitude larger than those in Figure 2.9, as would be expected. After deposition of PAA, surface roughness decreased fiom 20 i 2 nm to 9 i 2 nm on the substrate coated with 5 nm of gold and from 18.0 :t 0.4 nm to 8 :1: 1 nm on the surface coated with 63 nm of gold. The figure clearly shows that six layers of PAA completely cover underlying pores. 51 X 0.500 tin/div 2 100.000 nM/diu Figure 2.8 (a). Tapping-mode atomic force micrograph of gold-coated porous alumina before grafting of PAA films. The gold coating on 0.02-um surface pore-diameter-Anopore membrane was 5 nm thick in this image. 52 X 0.500 LIN/dIU 2 100.000 nn/diu Figure 2.8 (b). Tapping-mode atomic force micrograph of gold-coated porous alumina after grafting of 4-layer PAA films. The gold coating on 0.02-um surface pore-diameter-Anopore membrane was 5 nm thick in this image. 53 X 0.500 un/diu 2 100.000 nM/diu Figure 2.8 (c). Tapping-mode atomic force micrograph of gold-coated porous alumina after grafting of 6-layer PAA films. The gold coating on 0.02-um surface pore-diameter-Anopore membrane was 5 nm thick in this image. 54 X 0.500 lJM/diu 2 100.000 hM/dlU Figure 2.8 (d). Tapping-mode atomic force micrograph of gold-coated porous alumina before grafting of PAA films. The gold coating on 0.02-um surface pore-diameter-Anopore membrane was 63 nm thick in this image. 55 X 0.500 un/cliu 2 100.000 nH/diu Figure 2.8 (e). Tapping-mode atomic force micrograph of gold-coated porous alumina after grafting of 4-layer PAA films. The gold coating on 0.02—um surface pore-diameter-Anopore membrane was 63 nm thick in this image. 56 X 0.500 uM/diu 2 100.000 nM/diu Figure 2.8 (f). Tapping-mode atomic force micrograph of gold-coated porous alumina after grafting of 6-layer PAA films. The gold coating on 0.02-um surface pore-diameter-Anopore membrane was 63 nm thick in this image. 57 X 0.500 uM/diu 2 100.000 nM/diu Figure 2.9 (a). Tapping-mode atomic force micrograph of gold-coated alumina before grafting of hyperbranched PAA. The gold coating on 0.2 um pore- diameter-Anopore membrane was 5 nm thick in this image. 58 X 0.500 un/diu 2 100.000 hM/dlU Figure 2.9 (b). Tapping-mode atomic force micrograph of gold-coated alumina after grafting of 6-layers of hyperbranched PAA. The gold coating on 0.2 um pore- diameter-Anopore membrane was 5 nm thick in this image. 59 X 0.500 lm/diu 2 100.122 nn/diu Figure 2.9 (c). Tapping-mode atomic force micrograph of gold-coated alumina before grafting of hyperbranched PAA. The gold coating on 0.2 um pore-diameter- Anopore membrane was 63 nm thick in this image. 60 X 0.500 LIN/div 2 100.000 nn/diu Figure 2.9 (d). Tapping-mode atomic force micrograph of gold-coated alumina after grafting of 6-layers of hyperbranched PAA. The gold coating on 0.2 pm pore-diameter-Anopore membrane was 63 nm thick in this image. 61 Defects in Gold Coatings. Selective transport requires defect-free films because flux through open channels will negate the selectivity of the membrane.20 Unfortunately, during initial permeation studies we noticed that the 63 nm-thick gold coatings had small, but visible holes. Careful inspection under a microscope revealed that there were at least 20 small (1-100 um) holes in the gold layer on a substrate with an area of 3 cm2. We inspected substrates under a microscope immediately after metal deposition, and most of the substrates had holes in the gold layer. Rinsing these samples with methanol and drying them with nitrogen produced additional holes. Optical microscope images (Figure 2.10) often show that the small piece of gold removed from the surface after rinsing sits next to the newly formed hole. Our attempts to improve adhesion of gold by cleaning with hot solvents, dilute ammonium hydroxide, UV/ozone, or Ar plasma did not provide a defect-free metal layer. The use of an adhesion layer (Cr or Ti) to prevent peeling was also unsuccessful. The only method that appeared to give defect-free gold coating was deposition of a much thinner layer of gold (5 nm). We note, however, that it is more difficult to see defects in these thin gold coatings. Selectivities observed in gas-transport measurements (described below) give corroborating evidence that the 5-nm layer was free of defects. Gas-Transport Measurements. In the solution-diffusion model of flux through a membrane, the permeability coefficient of a given analyte is related to its . . . . 5 solubility and diffusion coefficient in the membrane material as shown in Equauon 2.3, P = S x D 2.3 where P is the permeability coefficient, S is the apparent solubility constant for the given gas—membrane pair, and D is the apparent diffusion coefficient. Selectivity is defined as 62 Figure 2.10 Optical micrograph of a defect on a gold-coated (63 nm) porous alumina (0.02 um-diameter surface pores) substrate. 63 the ratio of the permeability coefficients of two different gases. Thus, selectivities between gases can result from differences in either solubility or diffusivity. Using 5 nm gold-coated substrates, we began investigating the gas-permeability of 3-layer hyperbranched PAA films. Figure 2.11 shows how gas flow through a 3-layer PAA film varies with inlet pressure for various gases. Total flux through these PAA membranes depends primarily on the molar mass of the gases; thus, He has the highest flux and SF, the lowest. The selectivity ratios are not very high and fall around or below calculated Knudsen-diffusion values. Knudsen-diffusion of gases results from diffusion through pores with radii that are much smaller than the mean free path of the gases, and flux is inversely proportional to the square root of molar mass.20 These results suggest that 3-layer PAA films are rather porous. We note, however, that the permeability of these membranes is orders of magnitude lower than the permeability of gold-coated porous alumina. When PAA films were modified by fluorination with H2NCH2(CF2)6CF3, selectivities improved significantly, and the relative fluxes did not depend solely on molar mass (Figure 2.12). The enhancement in selectivity occurred because the permeability of some gases (He, O2, and CO2) increased upon fluorination, while other gases were relatively unaffected by derivatization. The selectivity ratios for several of the gas pairs are 3-4 times higher than those obtained with PAA films. Table 2.1 shows selectivity ratios for 3-layer PAA and 3-layer fluorinated PAA membranes, along with calculated values for Knudsen diffusion-based selectivities. The increase in selectivity upon fluorination is not simply due to filling of defects in the PAA films. If that were the case, flux would decrease for all gases and increases in flux for some of the gases upon 64 l I l l 0 He 4 _ A CH4 _ O I N2 [:1 02 o A A co2 : ug 3 I— 0 SF6 . — :1 E. o A 9 co 2 - A r h o g A I _ I Q 0 A L I 151 O 1 - A I 5 O - O A g g Q 0 l l l I 0 10 20 30 4O Inlet Pressure (psig) Figure 2.11 Flow rates of several gases through a three-layer PAA film on gold-coated (5 nm) porous alumina (0.02 pm surface pore diameter) as a function of inlet pressure. The membrane area was 2.8 cm2. Permeability values with standard deviations are listed in the text. 65 I I I I 0 He A CH4 30 " I N2 A " El 02 c A CO2 -- A g 0 SF6 . E 20 - A . " 45‘ s A - a . - E _ O _ 10 A D O D [:1 0 D E] I [1 D I I l I 0 5 H n O Q 0 Q 0 1O 20 30 40 Inlet pressure, psig Figure 2.12 Flow rates of several gases through a fluorinated three-layer PAA film on gold-coated (5 nm ) porous alumina (0.02-um pore diameter) as a function of inlet pressure. The membrane area was 2.8 cm2. Permeability values with standard deviation are listed in the text. 66 fluorination certainly would not occur. Also, the addition of subsequent PAA layers, which should increase coverage, does not improve selectivity. We tested gas permeabilities of 4-, 5-, and 6-layer PAA films. 4- and 5-layer PAA films still showed Knudsen-like selectivities, and 6-layer PAA fihn permeabilities were below the limit of detection of our flow-meter. The fact that additional PAA layers do not improve selectivity suggests that the lack of selectivity in these films is likely a property of hyperbranched PAA and not a manifestation of defects in substrate coverage by these films. Derivatization of PAA films with H2NCH2(CF2)6CF3 probably increases selectivities because of differing solubilities of gases in these films. Formation of a more crystalline membrane may also increase selectivity. The O2/N 2 selectivities we observe with fluorinated PAA agree well with selectivities of other fluorinated membranes. Several studies report higher 02 permeability values and improved O2/N2 selectivities (2 — 6) in a variety of fluorinated polymers such as a fluorine-containing polydimethoxysilane (PDMS)-tn'methylsilylpropyne (PTMSP) block copolymer, polyimides containing trifluoromethyl groups, and a fluorine-containing PDMS-ethyl cellulose graft copolymer.”38 Aoki et al. also showed that adding a small amount (1 wt %) of poly (trifluoromethyl substituted arylacetylene) to PDMS and PTMSP films 3 enhances the oxygen permeability and selectivity of O2/N 2. 9 The increased oxygen permeability of fluorinated membranes is most likely due to the high solubility of oxygen in fluorinated hydrocarbons.39'40 COz/CH4 selectivities of fluorinated PAA are also similar to those in a recent report on gas permeation through the fluorinated polymer poly(2,2-bis(trifluoromethyl)-4,5-difluoro- 1 ,3-dioxole-co-tetrafluoroethylene 67 Table 2.1 Ratios of the fluxes of different gases through 3-layer PAA films and fluorinated 3-layer PAA films.“ H6 CH4 N2 02 C02 SF6 1.3 (2.0) 1.9 (2.6) 2.0 (2.8) 2.0 (3.3) 2.4 (6.1) He — [6.7] [6.3] [2.8] [0.81] [24.4] 1.3 (2.0) 1.4 (1.3) 1.5 (1.4) 1.5 (1.7) 1.8 (3.0) CH4 — [6.7] [0.94] [0.42] [0.12] [3.6] 1.9 (2.6) 1.4 (1.3) 1.1(1.1) 1.1(1.3) 1.3 (2.3) N .— 2 [6.3] [0.94] [0.45] [0.13] [3.9] 2.0 (2.8) 1.5 (1.4) 1.1(1.1) 1.0 (1.2) 1.2 (2.1) O _ 2 [2.8] [0.42] [0.45] [0.29] [8.6] 2.0 (3.3) 1.5 (1.7) 1.1(1.3) 1.0 (1.2) 1.2 (1.8) (:0 — 2 [0.81] [0.12] [0.13] [0.29] [30.0] 2.4 (6.1) 1.8 (3.0) 1.3 (2.3) 1.2 (2.1) 1.2 (1.8) SF _ 6 [24.4] [3.6] [3.9] [8.6] [30.0] a Values for fluorinated films are in brackets. The ratios for Knudsen-diffusion selectivity are given in parentheses. All value represent the flux of the lighter gas divided by the flux of the heavier gas. 68 (TFE/BDD87).4l Of most relevance to this chapter, selectivities are consistent with those of poly(1,1 ’-dihydroperfluorooctyl acrylate), which is very similar in chemical composition to PAA derivatized with H2NCH2(CF2)6CF3.42 Control experiments show that the selectivity shown by fluorinated PAA membranes is due to the film on the surface of the membrane and not to grth inside the pores. After attempted growth of a 3-layer PAA fihn on a control sample (no MUA was deposited prior to PAA grafting) flux was too high to be measured by our flowmeter (>> 50 mL/min). After deposition of a 6-layer PAA films on a control sample, flux was still higher than that for a 3-layer PAA film grown with a MUA monolayer. These 6-layer PAA films did not show selectivity (not even Knudsen selectivity), even after fluorination. As mentioned earlier, FESEM analysis of control samples did not show blocking of the pores or the presence of a film on the surface. This control experiments show that the presence of a monolayer is required for film growth on the surface and effective coverage of pores. Additionally, any growth inside the pores is not affecting permeability results. We calculated permeability coefficients for 3-layer PAA membranes using Equation 1 and assuming a fihn thickness of 20 nm. The film thickness was estimated based on the thickness of 3-layer PAA films on gold-coated silicon wafers and FESEM images of PAA films on gold-coated porous alumina. The permeability coefficients for various gases are He, 2.6 i 0.7 barrer; N2, 1.4 i 0.6 barrer; SFb, 1.1 i 0.5 barrer; 02, 1.3 i 0.6 barrer; CO2, 1.3 i 0.7 barrer; and CH4 2.0 i 0.8 barrer, where 1 barrer is [1010 cm3 (STP) cm/(cm2 s cm(Hg)]. In calculating the permeability coefficients of fluorinated films, we used a thickness of 40 run because film thickness doubles upon fluorination 69 with H2NCH2(CF2)(,CF3.29 After fluorination, the permeability coefficients are higher for CO2 (24 i 3 barter), He (20 i 4 barrer), O2 (7 i 2 barrer), and N2 (3.1i 0.7 barrer), whereas coefficients for CH4 (2.9 i 0.4 barrer) and SF6 (0.8 i 0.1 barrer) are relatively unchanged. The permeability coefficients are 10-50 times higher than those for highly selective, ultrathin, conducting polymer membranes,ll but about 2 orders of magnitude lower than those reported by Merkel et al. for TFE/BDD87.“ The relatively low fluxes of fluorinated PAA may be due to crystallinity resulting from interactions between perfluorooctyl groups. We note that teflon has a low permeability.4l Low fluxes may also result from hydrogen bonding interactions in residual —COOH groups in PAA.” 4" 2.4 Conclusions The graft-on-graft synthesis and derivatization of PAA on gold-coated porous alumina yields selective, ultrathin membranes. Film thickness increases nonlinearly with the number of deposited layers and these films effectively cover porous substrates, presumably because of their hyperbranched structure. FESEM and AF M images clearly show that six-layer PAA fihns cover the surface without filling underlying pores. Gas- permeability measurements indicate that three-layer PAA films show modest Knudsen diffusion-based selectivity, but fluorination of these films provides selective, ultrathin membranes for gas separations. Although the complex synthesis of hyperbranched PAA films will likely prohibit their use in commercial gas separations, these permeability data do show that PAA membranes can be defect-free and easily derivatized. Thus, PAA membranes might prove valuable in small-scale sensing or separations applications. 70 2.5 References (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11) (12) (13) (14) (15) (16) (17) (18) Abelson, P. H. Science 1989, 244, 1421. Henis, J. M. S.; Tripodi, M. K. Science 1983, 220, 11-17. Haggin, J. Chem. Eng. News 1990, 68, 22-26. Spillman, R. W. Chem. Eng. Prog. 1989, 85, 41-62. Freeman, B. D.; Pinnau, I. in Polymer Membranes for Gas and Vapor Separation Freeman, B. D. and Pinnau, 1., Ed.; American Chemical Society: Washington, DC, 1999; pp1-27. Maier, G. Angew. Chem. Int. Ed. 1998, 37, 2960-2974. Freeman, B. D. Macromolecules 1999, 32, 375-380. Robeson, L. M. J. Membr. Sci. 1991, 62, 165-185. Robeson, L. M.; Burgoyne, W. F.; Langsam, M.; Savoca, A. C.; Tien, C. F. Polymer 1994, 35, 4970-4978. Rezac, M. E.; Koros, W. J. J. Appl. Polym. Sci. 1992, 46, 1927-1938. Liu, C.; Martin, C. R. Nature 1991, 352, 50-52. Prasad, R.; Shaner, R. L.; Doshi, K. J. in Polymeric Gas Separation Membranes; Paul, D. R. and Yampol'skii, Y. P., Ed.; CRC: Boca Raton, 1994; pp 513-614. Pinnau, I.; Koros, W. J. Ind. Eng. Chem. Res. 1991, 30, 1837-1840. Gantzel, P. K.; Merten, U. 1&EC Proc. Des. Dev. 1970, 9, 331. Suda, H.; Haraya, K. in Membrane Formation and Modification ; Pinnau, I. and Freeman, B. D., Ed.; American Chemical Society: Washington, DC, 2000; pp 295- 3 l 3. Francis, P. Fabrication and Evaluation of New Ultrathin Reverse Osmosis Membranes; NTIS Report, Pb-l77083, Springfield, VA, 1966. Le Roux, J. D.; Paul, D. R. J. Membr. Sci. 1992, 74, 233-252. Kesting, R. Synthetic Polymeric Membranes: A Structural Perspective; 2nd ed.; John Wiley & Sons, New York, 1985; pp 224-236. 71 (19) (20) (21) (22) (23) (24) (25) (26) (27) (28) (29) (30) (31) (32) (33) (34) Cadotte, J. in Materials Science of Synthetic Membranes ; Lloyd, D. R., Ed.; American Chemical Society: Washington, DC, 1985; pp 273-294. Kesting, R. E.; Fritzsche, A. K. Polymeric Gas Separation Membranes; John Wiley & Sons, New York, 1993; pp 284-318. Conner, M. D.; Janout, V.; Kudelka, I.; Dedek, P.; Zhu, J .; Regen, S. L. Langmuir 1993, 9, 2389-2397. Hendel, R. A.; Nomura, E.; Janout, V.; Regen, S. L. J. Am. Chem. Soc. 1997, 119, 6909-6918. Zhang, L.-H.; Hendel, R. A.; Cozzi, P. G.; Regen, S. L. J. Am. Chem. Soc. 1999, 121,1621-1622. Liu, (3.; Chen, w. J.; Martin, c. R. J. Membr. Sci. 1992,65, 113-128. Shieh, J. J.; Chung, T. 8.; Paul, D. R. Chem. Eng. Sci. 1999, 54, 675-684. Zhou, Y.; Bruening, M. L.; Bergbreiter, D. E.; Crooks, R. M.; Wells, M. J. Am. Chem. Soc. 1996, 118, 3773-3774. Bruening, M. L.; Zhou, Y.; Aguilar, G.; Agee, R.; Bergbreiter, D. E.; Crooks, R. M. Langmuir 1997, 13, 770-778. Odukale, A. A.; MS thesis; Michigan State University: East Lansing, MI, 1999. Zhou, Y.; Bruening, M. L.; Liu, Y.; Crooks, R. M.; Bergbreiter, D. E. Langmuir 1996, 12, 5519-5521. Kim, K. J .; Dickson, M. R.; Chen, V.; Fane, A. G. Micron and Microscopia Acta 1992, 23, 259-271. Coates, V. J. Proc. 40th Ann. Electron Microscopy Soc. Am. 1982, 752-753. A hyperbranched PAA film on a gold-coated wafer with an ellipsometric thickness of 970 A, has a carbonyl absorbance of 0.1 13. Using this data point and the fact that the absorbance due to a PAA membrane on alumina was 0.109, we estimated that the membrane thickness was 940 A. Ellipsometric measurements on gold wafers were performed with a Woollam M-44 rotating analyzer ellipsometer. Substrate parameters (11 and k) were determined prior to film deposition. Aoki, T. Prog. Polym. Sci. 1999, 24, 951-993. Coleman, M. R.; Koros, W. J. J. Membr. Sci. 1990, 50, 285-297. 72 (35) (36) (37) (38) (39) (40) (41) (42) (43) (44) Nagase, Y.; Ochiai, J.; Matsui, K.; Uchikura, M. Polym. Commun. 1988, 29, 10-13. Nagase, Y.; Ochiai, J.; Matsui, K.; Uchikura, M. Polymer 1988, 29, 740-745. Kim, T. H.; Koros, W. J .; Husk, G. R.; O'Brien, K. C. J. Membr. Sci. 1988, 37, 45- 62. Stem, S. A.; Mi, Y.; Yamamoto, H. J. Polym. Sci, Polym. Phys. Ed. 1989, 27, 1887-1909. Aoki, T.; Oikawa, E.; Hayakawa, Y.; Nishida, M. J. Membr. Sci. 1991, 57, 207-216. Inagaki, N.; Kawai, H. J. Polym. Sci, Polym. Chem. Ed. 1986, 24, 3381-3391. Merkel, T. C.; Bondar, V.; Nagai, K.; Freeman, B. D.; Yarnpolskii, Y. P Macromolecules 1999, 32, 8427-8440. Arnold, M. E.; Nagai, K.; Freeman, B. D.; Spontak, R. J .; Betts, D. E.; DeSimone, J. M.; Pinnau, I. Macromolecules 2001, 34, 5611-5619. Bergbreiter, D. E.; Tao, G. J. Polym. Sci. Part A: Polym. Chem. 2000, 38, 3944- 3953. Xiao, K. P.; Harris, J. J.; Park, A.; Martin, C. M.; Pradeep, V.; Bruening, M. L. Langmuir 2001, 17, 8236-8241. 73 Chapter 3 pH-DEPENDEN T GROWTH AND MORPHOLOGY OF MULTILAYER DENDRIMER/POLY(ACRYLIC ACID) FILMS Summary Because hyperbranched poly(amidoamine) dendrimers are well-defined polycations with a high density of surface charge, they provide unique materials for controlled synthesis of films by alternating electrostatic adsorption of polycations and polyanions. Thicknesses of generation 4 (G4) dendrimer/poly(acry1ic acid) (PAA) “bilayers” can be tuned from <10 A to >2,000 A simply by varying deposition pH, while generation 8 (G8) dendrimer/PAA bilayers can be as thick as 4,000 A. The highest thicknesses occur when using deposition pH values at which PAA and dendrimers are only partially charged, and the thickest films show high surface roughnesses, which may account in part for the rapid film growth. Deposition of dendrimer/PAA films on porous alumina supports yields highly gas-penneable membranes, in sharp contrast to several other multilayer polyelectrolyte systems, which show little gas permeability. Dendrimer/PAA films can also be induced to form microporous coatings through pH and salt-induced structural changes, and pore diameters in these structures can be varied from 0.02 11m to 0.4 um. 74 3.1 Introduction Alternating electrostatic adsorption of polycations and polyanions is a convenient, versatile technique for forming ultrathin films."2 Synthesis of these films simply involves altemating immersions of a substrate in solutions containing polycations and polyanions, and possible film constituents include materials as diverse as polymers}4 inorganic nanoparticles,5 and proteins.° Selection of these constituents as well as deposition conditions (i.e. pH,7 polyelectrolyte concentration,3 immersion time,8 and supporting electrolyte concentrationg'”) allows tailoring of fihn properties for possible applications 16,17 in sensing,”’13 separations,”’15 surface protection, controlled release,18 and electronics. 1 9 Because of their unusual hyperbranched structure, dendrimers present unique polycations that can be included in multilayer polyelectrolyte films.”22 Poly(amidoamine) (PAMAM) dendrimers are especially attractive constituents for such films because they have a relatively well-defined shape, a narrow molecular weight distribution, and a high concentration of amine groups at their periphery.”’24 The open 21 ,22,25 interior and dense shell of dendrimers may also impart unique adsorption and 26-28 transport properties to films containing these materials. We should note, however, that the structures of dendrimers depend on pH,29 dendrimer type,30 and whether the - - 25,31 dendrimer 1s bound to a surface. In some cases, backfolding of the dendrimer periphery may occur.30 Several studies demonstrated deposition of dendrimer-containing films in a layer- by-layer fashion.20'22’28’29’32'36 Regen and Watanabe constructed multilayered films using metal-ligand interactions between layers of dendrimers and organometallic complexes of 75 Ptz’L,22 while Liu and co-workers employed covalent bond formation to form films by alternating deposition of dendrimers and a poly(anhydride).28 Examples of alternating electrostatic adsorption with dendrimers includes dendrimer/p0ly(styrenesulfonate) (PSS),35‘3°, dendrimer/ glucose oxidase, dendrimer/p0lyoxometalates, and carboxylate- terminated dendrimer/amino-terminated dendrimer filmszo’z‘ In these cases, film thickness increases with the number of bilayers, but film formation was not studied as a function of deposition conditions. This paper reports the formation of multilayer films by alternating adsorption of protonated PAMAM dendrimers and ionized poly(acrylic acid) (PAA). Because these polyelectrolytes are weak acids and bases, their deposition should be very sensitive to pH.37 Remarkably, the thickness of dendrimer/PAA bilayers in such films can be varied from <10 A to as high as 4000 A by controlling pH and selecting the generation of dendrimer. Rubner and coworkers previously showed that deposition pH affects the thickness of poly(allylamine hydrochloride) (PAH)/PAA films, but the maximum thickness per bilayer in those studies was <150 A.37 The hyperbranched architecture of the dendrimers evidently results in much greater thicknesses. Like PAH/PAA systems, dendrimer/PAA films are capable of forming microporous structures upon exposure to low pH solutions.38 Unlike PAH/PAA, however, dendrimer/PAA films are highly permeable to gases. Experimental Section Materials. G4 amine-terminated PAMAM dendrimer (10 wt % solution in methanol, MW = 14215), poly(acrylic acid) (25 wt % aqueous solution, Mw = 90,000 or 76 powder with Mw = 2000), and 3-mercaptopropionic acid (MPA) were used as received from Aldrich Chemical. G8 amine-terminated PAMAM dendrimer (Mw = 233,383) was obtained from Dendritech (Midland, MI) as a 6.9 wt % aqueous solution. Gold-coated silicon wafers (200 nm of sputtered gold on 20 nm chromium on Si (100)) were used as substrates for ellipsometric measurements and external reflection FTIR spectroscopic studies. AnodiscTM porous alumina membranes with 0.02 [rm-diameter surface pores (Whatman) were used as supports for the preparation of composite membranes. Film preparation. Gold-coated silicon substrates were immersed in piranha solution (3:1 v/v mixture of concentrated H2SO4 and 30 % H202; caution, piranha reacts violently with organic compounds and should not be stored in closed containers) for 1-3 min, rinsed thoroughly with deionized water (Milli-Q, 18 MQcm), and dried with N2. The clean substrates were then immersed in a 1 mM mercaptopropionic acid (MPA) solution in ethanol for 30 min, rinsed with ethanol and deionized water, and dried with N2. This produces a carboxylic acid-containing monolayer that will be negatively charged upon deprotonation. Deposition of one dendrimer layer occurred by immersion of a monolayer-coated substrate in a solution of 10'5 M dendrimer for 10 min, followed by washing with water and drying with N2. The dendrimer-coated substrate was then immersed in 0.02 M PAA (molarity is given with respect to the repeating unit) for 5 min, washed with water, and dried with N2. This procedure was repeated until the desired number of dendrimer/PAA bilayers was deposited. The pH of dendrimer and PAA solutions was adjusted by adding a few drops of dilute HCl or NaOH solutions. 77 Film characterization. Ellipsometry, external reflection F TIR spectroscopy, field-emission scanning electron microscopy (FESEM), atomic force microscopy (AF M), optical microscopy, and gas transport measurements were performed as described in - - . 14.39 prevrous publicatlons. 78 W IAU Mercaptopropionic acid (1) (MPA) qoppppgpp.9920,09 MAPA u C-D carboxylate g} protonated multiply charged PAMAM Dendrimer Figure 3.1 Schematic diagram of the deposition of a dendrimer/PAA bilayer on a Au surface. Repetition of steps 2 and 3 yields multilayer films. 79 3.2 Results and Discussion Film formation and characterization. Figure 1 shows schematically the preparation of multilayer dendrimer/PAA films on gold. External reflection F TIR spectra confirm the formation of these films as shown in Figure 2. The spectra contain strong absorbances at 1550 cm'1 and 1650 cm‘1 that are due to the amide bonds in the dendrimer internal structure, and these bands increase in intensity with the addition of each dendrimer layer. After deposition of PAA layers, an absorbance at 17 30 cm'1 appears due to the carbonyl stretch of the —COOH groups of PAA (spectrum c). Upon subsequent deposition of the dendrimer, the acid carbonyl peak decreases to < 20 % of its initial value (spectrum d), suggesting that most of the —COOH groups are deprotonated during dendrimer adsorption. This is reasonable as the dendrimer was deposited at pH 8. The — C 00' symmetric stretch (1400 cm") appears after deposition of the dendrimer, while the —COO' asymmetric stretch (~1575 cm") is probably masked by the amide II band. Spectra are similar for G8 dendrimer/PAA systems, and these data are consistent with results for PAH/PAA films.7’l7 Effect of deposition pH on the thickness of G4 dendrimer/PAA films. Rubner and coworkers showed that deposition pH has a large effect on the thickness of PAH/PAA films.37 To see if similar effects occur in dendrimer/PAA films, we independently varied the pH of dendrimer and PAA deposition solutions from 2 to 8. Figure 3 shows ellipsometric thicknesses of 2.5-bilayer G4 dendrimer/PAA films as a function of the deposition pH for both dendrimers and PAA. (In this terminology, films with an integer number of bilayers have PAA as the top layer, whereas films with an additional 0.5 bilayer are terminated with dendrimers.) When dendrimers are deposited 80 Absorbance Amide I I 0-05 11 Amide 11 Acid ‘1 c=o A .__..._ C M v. p 1800 1600 1400 1200 Wavenumbers(cm '1) Figure 3.2 Reflection FTIR spectra of a MPA-modified gold substrate coated with (a) 0.5 (b) 1.5 (c) 2 and (d) 2.5 G4 dendrimer/PAA bilayers. The half bilayer results in the dendrimer being the top layer in the film. Deposition pH values were 8 for G4 dendrimers and 4 for PAA 81 Thickness (A) Figure 3.3 Ellipsometric thicknesses of 2.5-bilayer G4 dendrimer/PAA films as a function of the pH used for dendrimer and PAA deposition. Standard deviations of thicknesses measured on three different samples ranged from 3 to 35 % of the measured values. 82 at pH 2, 4, or 6, and PAA is deposited at pH 8, little film growth occurs (thickness <5 A). At these pH values, both dendrimers and PAA are highly charged, and minimal film growth takes place probably because PAA is in a thin, extended conformation parallel to the substrate surface.37 In most other cases, thickness increases when decreasing the deposition pH for PAA or increasing the deposition pH for dendrimers. Higher thicknesses likely result from a lower charge density on both PAA and dendrimers in these solutions and the resultant need for more adsorption to compensate the surface charge.37‘40’4| At very low PAA deposition pH values (pH 2), PAA may desorb from the film,37 and thus some of these coatings are not as thick as films in which PAA was deposited at pH 4. The maximum thickness of 2.5-bilayer G4 dendrimer/PAA films was 1200 A, while reported maximum bilayer thicknesses for PAH/PAA films are < 150 A.37 Thus the architecture of the dendrimer must greatly affect film growth. The trends in Figure 3 also apply as more layers are added to the film. After deposition of 4.5 dendrimer/PAA bilayers, films deposited at a dendrimer pH of 4 and a PAA pH of 6 still have thicknesses less than 10 A. In contrast, Table 1 shows that the thicknesses of G4 dendrimer (pH=8)/PAA (pH=4) fihns grow by 4,300 A on going from a 2.5-bilayer to a 4.5-bilayer film. Table 1 also indicates that film thicknesses initially increase nonlinearly with the number of bilayers, but bilayer thickness becomes relatively constant after deposition of 2.5 dendrimer/PAA bilayers. Such a grth profile is often seen in the build-up of polyelectrolyte multilayers because a consistent surface charge distribution does not fully develop until after adsorption of a few bilayersfl’42 83 Table 3.1 Ellipsometric thicknesses of G4 and G8 PAMAM Dendrimer (pH=8)/PAA (pH=4) fihns on gold as a function of the number of bilayers. G4 Dendrimer G8 Dendrimer iigngnfigifiig/Slifla Thickness/A Thickness/A Thickness/A bilayers) MW of PAA = 90000 Mw of PAA = 90000 MW of PAA= 2000 0.5 9 i 1 14 i 2 10 i- l 1 28i4 36:7 32.11.02 1.5 110:30 310i10 140:10 2 300i30 670i30 3003210 2.5 ll70i90 3990:60 1020i 10 3 2410i 190 4950:80 l930i30 3.5 3060 i 100 7000 i 10 2780 j: 20 4 4540 i 370 8940 i 210 3990 j: 60 4.5 5470 i 600 11390 i 550 Fil’ggjggme “ Half bilayers correspond to dendrimers being the last layer deposited 84 Effect of deposition pH and PAA chain length on the thickness of G8 dendrimer/PAA films. Thicknesses of G8 dendrimer/PAA films provide further evidence that film thickness depends on dendrimer structure. G8 dendrimers have 16 times as many peripheral amine groups as G4 dendrimers and about double the diameter, and this increased size and crowded exterior functionality is clearly reflected in the thicknesses of G8 dendrimer/PAA films (Figure 4). As with G4 PAMAM dendrimers, thickness usually increases with increasing dendrimer deposition pH and decreasing PAA deposition pH. However, 2.5-bilayer G8 dendrimer (pH=8)/PAA (pH=4) films have thicknesses of 4000 A, more than three times that for similar G4 dendrimer/PAA films. As shown in Table l, the thicknesses of G8 dendrimer/PAA fihns grow by as much as 3300 A upon dendrimer deposition. This thickness increase would represent an addition of 30 dendrimer monolayers (assuming a dendrimer diameter of 92 A43). Most likely, the PAA already in the film elongates or rearranges to provide adsorption sites for these dendrimers. Extension of PAA by 3000 A would require at least 1000 monomer units, or a PAA molecular weight of 72000, assuming an extended length of 3 A per monomer. The PAA employed in this study had an average molecular weight of 90000, so such an extension is possible. When we prepared films from PAA with a molecular weight of only 2000, however, we observed much smaller growth as seen in Table l.44 The lower growth would be consistent with the inability of PAA to extend over large distances, but dendrimer deposition still resulted in thickness increases of 700 - 900 A. These thicknesses are ten times longer than the extended length of PAA chains with a molecular weight of 2000. Evidently, the PAA can also redistribute in the films. Previous studies of multilayer polyelectrolyte films such as PAH/PSS and alternating 85 Thickness (A) Figure 3.4 Ellipsometric thicknesses of 2.5-bilayer G8 dendrimer/PAA films as a function of the pH used for dendrimer and PAA deposition. Standard deviations of thicknesses measured on three different samples ranged from 2 to 29 % of the measured values. 86 carboxylate- and amino-terminated dendrimers showed little dependence of thickness on the molecular weights of polycations and polyanions.4'2"45 However in those cases, films were relatively thin and there would have been little need from chains to extend to accommodate incoming polyelectrolytes. Surface Roughness. One might expect that very thick polyelectrolyte films would be quite rough. Figure 5 contains optical microscope images of 2.5-bilayer G8 dendrimer (pH=4)/PAA (pH=4) and 2.5-bilayer G8 dendrimer (pH=8)/PAA (pH=4) films. As the images show, films deposited with the dendrimer solution at pH 4 (thickness=80 A) appear smooth, whereas a dendrimer deposition pH of 8 (thickness=4,000 A) produces a rough surface. Quantitative measurements of surface morphology using AF M also show the roughness of thick films. Differences in the surface morphology can be expressed in terms of a roughness parameter, Rq, which is the standard deviation of the Z values (heights) within a given area. R. : Jan—2.32 Np (1) Values of R, were calculated using equation 1, where Z.- is a specific Z value, Zavg is the average of the Z values within the given area, and Np is the number of points within a given area.46 The surfaces of 2.5-bilayer G8 dendrimer (pH=4)/PAA (pH=4) and 2.5- bilayer G8 dendrimer (pH=8)/PAA (pH=4) films had roughnesses of 2.9 i 0.5 nm and 12.5 i 2.6 nm, respectively for 1 pm x 1 11m spots. This is consistent with the optical microscopy data and suggests that in the case of thicker films, there is some aggregation of dendrimer during film growth. 87 Figure 3.5 Optical microscope images of (a) 2.5-bilayer G8 dendrimer (pH=4)/PAA (pH%) and (b) 2.5-bilayer G8 dendrimer (pH=8)/PAA (pH=4) films. The ellipsometric thicknesses of the films were 80 A and 4000 A, respectively. 88 Microporous film formation. In addition to their pH-dependent growth properties, PAH/PAA fihns are rather unique because they can form microporous structures upon exposure to solutions with different pH values38 or ionic strengths.47 Dendrimer/PAA films are also capable of forming such microporous films. Figures 6(a) and 6(b) show AF M images of 5-bilayer G4 dendrimer (pH=8)/PAA (pH=4) films prepared in the presence of 0.2 M NaCl before and after immersion in a pH 2.5 solution for 1 min and washing with pure water. Exposure to acidic solution results in micropores (0.02 - 0.07 pm in diameter) in the initially homogeneous film. Figure 6(c) shows that larger pores (0.1 - 0.4 um diameter) form when fihns are prepared in the absence of salt. Rubner suggested that these changes in morphology occur because the low pH results in cleavage of interchain ionic bonds through protonation of the carboxylate groups of PAA.”47 This, in turn, allows large scale reorganization and phase separation from the acidic water solvent, yielding micrometer-sized pores. This reorganization appears to be independent of whether the cation is PAH or a dendrimer, but it does depend on the ionic strength used in film deposition. The addition of supporting electrolyte should screen charges on the films and result in a structure with more loops and tails.4 Such a loopy structure could result in the smaller pores seen for films deposited in 0.2 M NaCl and exposed to pH 2.5 solution. 89 10.0 n- Figure 3.6 AFM images of 5-bilayer G4 dendrimer (pH=8)/PAA (pH=4) fihns before (a) and after (b, c) immersion in a pH 2.5 aqueous solution. For images (a) and (b), polyelectrolytes were deposited in the presence of 0.2 M NaCl, and for image (c), no salt was used during film deposition. 90 Permeability of Dendrimer/PAA Layers. Because dendrimers have a relatively open interior, we thought that dendrimer/PAA membranes might be highly permeable to gases. To test the permeability of these films, we deposited 4.5-bilayer G4 dendrimer (pH=8)/PAA (pH=4) films on porous alumina substrates. Cross-sectional field-emission scanning electron microscope (FESEM) images (Figure 7) show that these thick films cover the substrate effectively without filling pores.48 Interestingly, fluxes of all the gases that we tested (He, O2, N2, H2, CO2, and CH4) were larger than the detection limit of our gas flow meter (>030 mL/cm2 sec), even at a gas pressure of 10 psig. This is in stark contrast to 2-bilayer PAH/PAA films (thickness of 100 A), which allow very low fluxes at 10 psig gas inlet pressures (0.003 to 0.04 mL/cm2 sec for He, 02, N2, H2, CO2, and CH4). The dendrimer/PAA fihn is about 35 times thicker than the PAH/PAA and yet it allows fluxes that are several orders of magnitude higher than those through PAH/PAA. We also tested gas permeation through 5, 4, and 3-bilayer PAH/PAA films for this comparison, but fluxes through these films were below the limit of detection of our flow meter (<6 x 10 “5 mL/cm2 sec). At this point, we do not know if the high flux through dendrimer/PAA films is due to an open structure of the dendrimer,” loose packing of the film or the presence of defects. However, FESEM images suggest that the fihns are defect-free. The high flux through dendrimer/PAA films suggests that they may prove useful in providing a continuous substrate on which to deposit highly selective, ultrathin polyelectrolyte gas-separation layers.50 The minimal thickness of the selective layer would allow high flux in addition to high selectivity. 91 jf~ tetfiymm x7008 sew) 6/2319912 17 Figure 3.7. Cross-sectional FESEM images of porous alumina before (a) and alter (b) coating with a 4.5-bilayer G4 dendrimer (pH=8)/PAA (pH=4) film. 92 3.5 Conclusions The thicknesses of multilayer dendrimer/PAA films depend on both dendrimer generation and pH. For a given dendrimer, the thickest films result from deposition of dendrimers at high pH and PAA at pH 4, and films prepared with G8 dendrimers are generally 2 - 4-fold thicker than films made with G4 dendrimers. Because multilayers of dendrimer adsorb to the surface in a single deposition step, PAA likely reorients in the film to accommodate the dendrimers. Lower molecular weight PAA is less capable of accommodating dendrimers and thus yields thinner films. Although dendrimer/PAA films can be very thick, they are also highly permeable to gases, perhaps because of their open structure. Additionally, if desired, microporous structures can be fabricated by brief exposure of films to pH 2.5 aqueous solutions followed by rinsing with water. The high permeability and thicknesses of dendrimer/PAA films show that the structure of these materials is significantly different from that of PAH/PAA. Thus dendrimer-containing films should provide expanded versatility for multilayer polyelectrolyte deposition. 93 3.6 References (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11) (12) (13) (14) (15) (16) (17) (18) Decher, G. Science 1997, 277, 1232-1237. Iler, R. K. J. Colloid Interface Sci. 1966, 21, 569-594. Ferreira, M.; Rubner, M. F. Macromolecules 1995, 28, 7107-7114. Ldsche, M.; Schmitt, J .; Decher, G.; Bouwman, W. G.; Kjaer, K. Macromolecules 1998, 31, 8893-8906. Ostrander, J. W.; Mamedov, A. A.; Kotov, N. A. J. Am. Chem. Soc. 2001, 123, 1 101-1110. Lvov, Y.; Ariga, K.; Ichinose, 1.; Kunitake, T. J. Am. Chem. Soc. 1995, 11 7, 6117-6123. Yoo, D.; Seimei, S.; Shiratori, S. S.; Rubner, M. F. Macromolecules 1998, 31 , 4309-4318. Decher, G.; Hong, J. D.; Schmitt, J. Thin Solid Films 1992, 210, 831-835. Schlenoff, J. B.; Ly, H.; Li, M. J. Am. Chem. Soc. 1998, 120, 7626-7634. Sukhishvili, S. A.; Granick, S. J. Chem. Phys. 1998, 109, 6861-6868. Sukhorukov, G. B.; Schmitt, J .; Decher, G. Ber. Bunsen-Ges. Phys. Chem. 1996, 100, 948-953. Yang, X.; Johnson, 8.; Shi, J.; Holesinger, T.; Swanson, B. Sensors and Actuators, B: Chemical 1997, 45, 87-92. Monterel, M. M.; Sukhorukov, G. B.; Petrov, A. 1.; Shabarchina, I.; Sukhorukov, B. I. Sensors and Actuators B: Chemical 1997, 42, 225-231. Harris, J. J.; Stair, J. L.; Bruening, M. L. Chem. Mater. 2000, 12, 1941-1946. Krasemann, L.; Tieke, B. Langmuir 2000, 16, 287-290. Farhat, T. R.; Schlenoff, J. B. Electrochemical and Solid-State Lett. 2002, 5, B13- B15. Dai, J.; Sullivan, D. M.; Bruening, M. L. Ind. Eng. Chem. Res. 2000, 39, 3528- 3535. Caruso, F.; Yang, W.; Trau, D.; Renneberg, R. Langmuir 2000, 16, 8932-8936. 94 (19) (20) (21) (22) (23) (24) (25) (26) (27) (28) (29) (30) (31) (32) (33) (34) (35) (36) (37) Fou, A. C.; Rubner, M. F. Macromolecules 1995, 28, 7115-7120. Tsukruk, V. V. Adv. Mater. 1998, 10, 253-257. Tsukruk, V. V.; Rinderspacher, F.; Bliznyuk, V. N. Langmuir 1997, 13, 2171- 2176. Watanabe, S.; Regen, S. L. J. Am. Chem. Soc. 1994, 116, 8855-8856. Tomalia, D. A.; Durst, H. D. Top. Curr. Chem, 1993, 165, 193-313. Li, J.; Piehler, L. T.; Qin, D.; Baker, J. R., Jr.; Tomalia, D. A.; Meijer, D. J. Langmuir 2000, 16, 5613-5616. Bliznyuk, V. N.; Rinderspacher, F.; Tsukruk, V. V. Polymer 1998, 39, 5249-5252. Kovvali, A. S.; Sirkar, K. K. Ind. Eng. Chem. Res. 2001, 40, 2502-2511. Dillon, R. E. A.; Shriver, D. F. Chem. Mater. 2001, 13, 1369-1373. Liu, Y.; Bruening, M. L.; Bergbreiter, D. E.; Crooks, R. M. Angew. Chem. Int. Ed. Engl. 1997, 36, 2114-2116. Wang, J.; Chen, J.; Jia, X.; Cao, W.; Li, M. Chem. Commun. 2000, 6, 511-512. Bosman, A. W.; Bruining, M. J .; Kooijman, H.; Spek, A. L.; Janssen, R. A. J .; Meijer, E. W. J. Am. Chem. Soc. 1998, 120, 8547-8548. 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. J. Am. Chem. Soc. 1998, 120, 4492- 4501. Anzai, J. I.; Kobayashi, Y.; Nakamura, N.; Nishimura, M.; Hoshi, T. Langmuir 1999, 15, 221-226. Yoon, H. C.; Kim, H. S. Anal. Chem. 2000, 72, 922-926. Cheng, L.; Cox, J. A. Electrochem. Commun. 2001, 3, 285-289. He, J. A.; Valluzzi, R.; Yang, K.; Dolukhanyan, T.; Sung, C. M.; Kumar, J .; Tripathy, S. K.; Samuelson, L.; Balogh, L.; Tomalia, D. A. Chem. Mater. 1999, 11, 3268-3274. Khopade, A. J .; Caruso, F. Nano Lett. 2002, 2, 415-418. Shiratori, S. S.; Rubner, M. F. Macromolecules 2000, 33, 4213-4219. 95 (38) (39) (40) (41) (42) (43) (44) (45) (46) (47) (48) (49) (50) Mendelsohn, J. D.; Barrett, C. J.; Chan, V. V.; Pal, A. J.; Mayes, A. M.; Rubner, M. F. Langmuir 2000, 16, 5017-5023. Nagale, M.; Kim, B. Y.; Bruening, M. L. J. Am. Chem. Soc. 2000, 122, 11670- 11678. Schlenoff, J. B.; Dubas, S. T. Macromolecules 2001, 34, 592-598. Dubas, S. T.; Schlenoff, J. B. Macromolecules 1999, 32, 8153-8160. Hoogeveen, N. G.; Cohen Stuart, M. A.; Fleer, G. J. Langmuir 1996, 12, 3675- 3682. Tomalia, D. A.; Naylor, A. M.; Goddard III, W. A. Angew. Chem. Int. Ed. Engl. 1990, 29, 138-175. Growth of G4 dendrimer (pH=8)/PAA (pH=4) films using PAA with a molecular weight of 2000 resulted in cloudy films after deposition of only 2 bilayers. Thus we could not determine whether PAA molecular weight affects the thicknesses of these films. Cheung, J. H.; Stockton, W. B.; Rubner, M. F. Macromolecules 1997, 30, 2712- 2716. Singh, S.; Khulbe, K. C.; Matsuura, T.; Rarnamurthy, P. J. Membr. Sci. 1998, 142,111-127. Fery, A.; Scholer, B.; Cassagneau, T.; Caruso, F. Langmuir 2001, 17, 3779-3783. Thicknesses of these films are reasonably consistent with thicknesses of films on gold-coated wafers. Jansen, J. F. G. A.; de Brabander-van den Berg, E. M. M.; Meijer, E. W. Science 1994, 266, 1226-1229. Sullivan, D. M.; Bruening, M. L. Chem. Mater 2002, in press. 96 Chapter 4 THE USE OF DENDRIMER/POLY(ACRYLIC ACID) FILMS AS ADHESIVES Summary Dendrimer-mediated adhesion is a convenient method for bonding substrates because it provides a thin adhesive layer that can be deposited from aqueous solution. This study shows the use of amine-terminated poly(amidoamine) (PAMAM) dendrimers as an adhesive interlayer between poly(acrylic acid) (PAA)-tenninated surfaces. Specifically, gold-coated silicon wafers modified with 5 G4 dendrimer/PAA bilayers on 5 precursor PAH/PAA bilayers are bonded with a G4 dendrimer interlayer. Cross-sectional field-emission scanning electron microscope (FESEM) images show that the dendrimer interlayer bonds the surfaces together without leaving gaps between the substrates. The adhesive forces between the substrates are most likely the strong electrostatic interactions that exist between the exterior amine groups of protonated G4 dendrimers and the —COO' groups of ionized PAA. Tensile tests show that the bonding strength is >1.6 MPa. The thin adhesive layer along with the high bonding strength may allow use of variations of this technique for applications such as wafer bonding to form lab-on-a-chip or micro- electro-mechanical-systems (MEMS). 97 4.1 Introduction Bonding of intrinsically nonadherent materials is generally achieved using a variety of adhesives such as epoxies "2 and cyanoacrylates.3 However, for some application such as wafer bonding to form three-dimensional micromachined devices, deposition of a thick layer of adhesive is not acceptable, and thin adhesion layers may prove useful.4'7 It is, however, difficult to utilize organic monolayers for adhesion due to a lack of contact between opposite monolayers on even slightly rough surfaces.8’9 We surmised that the deposition of slightly thicker PAA/PAMAM dendrimer adhesion layers might be useful for wafer-to-wafer bonding because of the possibility of interrningling of layers and strong electrostatic interactions between protonated dendrimers and ionized PAA. Such a process would require no heating and no organic solvents. Traditionally, wafer-bonding has been achieved in high-temperature processes using direct bonding, field-assisted silicon-to-glass bonding (anodic bonding) or eutectic bonding between silicon wafers via a thin metal interlayerfuo Because of the difficulties associated with these procedures (e. g., the need for strict control over ambient conditions, rigorous chemical cleaning and conditioning of the surfaces, and high temperatures), simple bonding using organic adhesion layers is highly desirable. With respect to traditional wafer-bonding methods, the main advantages of organic adhesives are the relatively low bonding temperatures, the absence of electric voltage or current, and the ability to join practically any kind of wafer materials.5 A few recent studies demonstrated that the multidentate nature of dendrimers can be useful in promoting adhesion. Zamborini and co-workers recently described the use of dendrimer monolayers as adhesion promoters between vapor-deposited gold films and 98 glass or silicon wafers.11 Interaction of the PAMAM dendrimers with both the substrate and the gold fihns should increase with the number of multidentate interactions, and thus the high density of amine groups at the densely packed surface of G8 dendrimers (1,024 primary amine groups per dendrimer) provides strong adhesion. Street et al. also noted that the presence of a self-assembled monolayer of amine-terminated G8 dendrimers on glass results in decreased surface roughness and increased hardness of an evaporated gold film. 12,13 Because dendrimer/silicon and dendrimer/ glass interactions are relatively weak,ll we thought that the use of dendrimer/poly(acrylic acid) (PAA) films might enhance the adhesion properties of dendrimer-containing films and allow wafer-to-wafer bonding. Poly(acrylic) acid adheres well to charged surfaces, and we hoped to use electrostatic interactions (between dendrimers and PAA as well as PAA and the substrate), rather than coordination, to enhance adhesion. In this chapter, we present a preliminary study of adhesion promoted by a protonated G4 dendrimer layer between two surfaces coated with partially ionized PAA-terminated dendrimer/PAA films. Gold-coated silicon wafers are bonded using this dendrimer adhesion interlayer after coating the surface with S-bilayer dendrimer/PAA films using alternating electrostatic adsorption. The ~l .3 um-thick dendrimer/PAA multilayer film and the electrostatic attraction between dendrimers and PAA yields bonding strengths as high as ~1.6 MPa. In fact, epoxied joints between the wafer and the testing apparatus fracture before the adhesion layer does. 4.2 Experimental Materials. Poly(acrylic acid) (MW = 90,000, 25 wt % aqueous solution), poly(allylamine hydrochloride) (Mw=70,000), 3-mercaptopropioinc acid (MPA) and a 10 99 wt % solution of G4 amine-terminated PAMAM dendrimer (Mw = 14,215) in methanol were purchased from Aldrich and were used as received. 18 MQ-cm Milli-Q water was used in solution preparation. Gold-coated silicon wafers (200 nm of sputtered gold on 20 nm chromium on Si (100)) were used as substrates for ellipsometric measurements, external reflection F TIR spectroscopy, and wafer-bonding studies. PAH (0.02 M, pH=4.0) and PAA (0.02 M, pH=4.0) solutions contained 0.5 M NaBr and 0.5 M NaCl, respectively, and solution pH values were adjusted by adding NaOH or HCl. Dendrimer solutions (10'5 M) were adjusted to pH 8.0 and contained no supporting electrolyte. (The molarities of PAA and PAH are given with respect to the repeating unit, while dendrimer molarity is given with respect to polymer molecular weight.) Film Deposition. Gold-coated silicon wafers were initially cleaned in piranha solution (3:1 H2804:H202-caution-strong oxidizer, store waste piranha only in loosely capped bottles) for 1 minute, rinsed with deionized water, and dried with N2. Cleaned gold-coated silicon wafers were first exposed to a solution of 1 mM mercaptopropionic acid (MPA) in ethanol for 30 min, rinsed with ethanol and deionized water, and dried with nitrogen to produce a —COOH-terminated surface. The substrates were then alternately immersed in solutions containing PAH and PAA for 5 min with rinsing (l min with a stream of deionized water) between adsorption of each polyelectrolyte. After formation of S-bilayers of PAH/PAA, the precursor film-coated wafers were immersed in an aqueous 10'5 M G4 dendrimer solution for 10 min, washed with water, and dried with N2. The substrates were then immersed in a PAA solution for 5 min, rinsed, and dried with N2. This procedure was repeated until 5 G4 dendrimer (pH = 8)/PAA (pH=4) bilayers were 100 deposited on the precursor 5-bilayer PAH/PAA film. For the wafer-to-wafer bonding, films were prepared on two gold-coated wafers with areas of 1.6 cm2, and then one drop of 10'5 M G4 dendrimer solution (pH=8, no salt) was placed on one of the wafers, after which the two wafers were simply pressed together by hand for about 20 min. Shorter amount of pressing time may work to make wafer-to-wafer bonding. Characterization. External reflectance FTIR spectra of polyelectrolyte films on gold-coated silicon wafers were obtained using a Nicolet Magna-IR 560 spectrometer. Ellipsometric thickness determinations were made with a rotating analyzer ellipsometer (Model M-44: J. A. Woollam) using WVASE32 software, and thicknesses were determined at three separate areas on each sample. Field-emission scanning electron microscopy (FESEM) (Hitachi S-4700) was used to take cross-sectional images of bonded wafers, and testing of bonding strength was performed using a United SFM-20 testing machine, manufactured by the United Calibration Corporation. A strain-gauge type load cell measures the load applied to the sample, and an extensiometer measures the strain imposed on the sample. Increasing load was applied to the sample at a rate of 158 lbs/min. The bonded wafers were attached to two aluminum posts with epoxy so that a load could be applied to the system. 4.3 Results and Discussion Film formation and characterization. In chapter 3, we showed that the highest thicknesses of dendrimer/PAA films occur when using a deposition pH of 4 for PAA and 8 for dendrimers.l4 These high thicknesses likely result from a lower charge density on both PAA and dendrimers at these pH values and the resultant need for more 101 adsorption to compensate the surface charge.”’” We chose these two pH values for this work so that dendrimer/PAA films would be sufficiently thick for wafer bonding. Figure 4.1 presents schematically multilayer dendrimer/PAA films on two gold-coated silicon wafers prior to bonding. Addition of the dendrimer interlayer presumably zips the films together as shown in Figure 4.1. We used 5-bilayer PAH/PAA films as a precursor because this allows formation of slightly thicker dendrimer/PAA coatings. However, the use of this precursor is probably not necessary. Figure 4.2 shows ellipsometric thicknesses of G4 dendrimer (pH = 8)/PAA (pH = 4) films on a base of 5 PAH (pH = 4)/PAA (pH = 4) bilayers as a function of the number of dendrimer/PAA bilayers. The formation of dendrimer/PAA films on the base layer occurs largely due to the electrostatic attractions between negatively charged PAA and positively charged G4 dendrimers. Ellipsometry demonstrates that thickness increases rapidly and linearly with the addition of each new bilayer after deposition of the base 5- bilayer PAH/PAA film, and suggests that a consistent surface charge distribution has already fully developed after adsorption of one dendrimer/PAA bilayer."”18 The average thickness of dendrimer/PAA bilayers (excluding the first bilayer) was ~ 280 nm. This value is about ~1.3 times that of bilayer thicknesses for similar dendrimer/PAA films deposited directly on gold modified with mercaptopropionic acid. The extra thickness with the base PAH/PAA films may occur because the presence of salt during base layer deposition results in a loopy structure that allows excess polymer to be accommodated over several polymer layers.15 Regardless of the reason, deposition on a base layer easily allows ~l .3 rim-thick films to adsorb onto a gold-coated silicon wafer. 102 5dendrimer/PAA< .1“: ‘M‘Tf' bilayers 5 dendrimer/PAAy bilayers Si Wafer ,. ‘ '+, 11+; base / 5 PAH/PAA bilayers Figure 4.1 Strategy for wafer bonding using dendrimer/PAA films. The inset on the right shows intertwining of dendrimer/PAA bilayers that should occur throughout the film. 103 1400 1200 - A 1000 800 600 Thickness(nm + 400 — 9 O O L = 1 . Base 1 2 3 4 5 Bilayers(dendrimer/PAA) Figure 4.2 Ellipsometric thicknesses of dendrimer/PAA films deposited on a base composed of 5 bilayers of PAH/PAA. Deposition pH values were 8 for G4 dendrimers and 4 for PAA. The PAH/PAA base was deposited at pH 4. 104 FESEM Characterization of Bonded Wafers. After deposition of 5 dendrimer/PAA bilayers on a modified substrate, wafer-to-wafer bonding can be effected by adding a drop of 10'5 M dendrimer solution to one wafer and pressing a second wafer onto this surface using finger pressure for 20 min. Figure 4.3 shows cross-sectional FESEM images of dendrimer/PAA films on gold-coated wafers before and after bonding two wafers together using the insertion of a dendrimer adhesion layer. These images suggest that wafers bond with coatings in intimate contact. Using these FESEM images we estimated film thickness to be ~1.3 um, which is consistent with ellipsometric measurements. In addition to using insertion of a dendrimer interlayer to bond coated wafers, we attempted to directly bond a wafer coated with 5 dendrimer/PAA bilayers to one that was coated with 5.5 dendrimer/PAA bilayers (both dendrimer/PAA films were deposited on a base of 5 PAH/PAA bilayers). We thought that the interaction of a dendrimer- terminated film with a PAA-terminated film might bond the wafers together. We used 1 drop of deionized water to promote interaction between the two films. We found, however, that the two wafers could be separated by hand, suggesting that little interaction between the two films occurred. Figure 4.4 shows a cross-sectional FESEM image of this system before separating the two wafers. This image indicates that the dendrimer- terminated surface and the PAA-terminated surface are not in intimate contact. In fact, the gap between the two wafers coated wafers is ~10 um. Evidently, pressing these films together was not sufficient to initiate full contact between the wafers. Bonding Strength Measurements. The schematic diagram in Figure 4.5 illustrates the configuration used for testing the strength of bonding between wafers. 105 Aluminum rods (2.5 cm diameter, 2 cm height) were attached to each of the bonded wafers with epoxy adhesive, and then force was slowly applied to pull the rods apart. This test yields a value for the tensile strength, 0 ,5, which is expressed by 0 (5 = F max /A, where F max is the maximum tensile load before fracture and A is the bonded area. The dimensions of the bonded area were 1.25 X 1.25 cm. However, when breakage did finally happen, it didn’t occur between the wafers, but rather between the epoxy and the aluminum rod. Thus we could measure only a minimum tensile strength of >1 .6 MPa. However, this value is still very high. The experiment was repeated with at least 2 samples, and they never broke at the junction between the wafers. This strong wafer bond using a dendrimer adhesion layer is likely due to dendrimers that interact electrostatically with both outer PAA layers. Multiple electrostatic interactions should hold the layers together. One question in this bonding is just how much dendrimer was added to the surface, and would this amount be compatible with a fill] intermingling of dendrimers and outer PAA layers. Approximately 0.03 mL of 10’5 M dendrimer solution was applied to the 1.6 cm2 wafer area. Assuming a dendrimer density of 1 g/cm3, this would result in deposition of ~260 A of dendrimer. This is about 20 % of the amount of dendrimer added during deposition of a dendrimer/PAA layer, and thus intermingling of the layers is possible, because a thick interlayer composed solely of dendrimer need not form. 106 Figure 4.3 Cross-sectional field-emission scanning electron microscopy images of (a) 5 dendrimer/PAA bilayers on 5 PAH/PAA bilayers on a gold-coated wafer and (b) two wafers coated with similar films that were pressed together with ~ 0.025 ml of dendrimer solution between them. 107 5.5(dendn'mer/PAA) o? 5(PAH/PAA) 5(dendn’mer/PAA) ofi' '- ’- 5(PAH/PAA) — E 10.0 u m Figure 4.4 Cross-sectional field-emission scanning electron microscopy image of two film-coated wafers that were pressed together with a drop of water between them. The film on the top gold-coated wafer was composed of 5.5 dendrimer/PAA bilayers on 5 PAH/PAA bilayers, while the film on the bottom gold-coated wafer contained 5 dendrimer/PAA bilayers on 5 PAH/PAA bilayers. 108 Applied Force Aluminum Rod / Epoxy Adhesive all” Bonded Wafer Pair { . 7 -. _ .- 3<— Derxlrimer Adhesion Layer .\ / Aluminum Rod Epoxy Adhesive App lied Force Figure 4.5 The tensile strength measurement configuration. A tensile tester is used to pull the bonded wafer apart.10 109 4.4 Conclusions Deposition of G4 dendrimer between dendrimer/PAA films capped with PAA allows preparation of a 3 um-thick adhesive film between wafers. Electrostatic adsorption of protonated dendrimer in the two ionized PAA layers probably results in electrostatic interactions between the films. The adhesion strength of this bonding is higher than that of the epoxy joint used to attach the wafers to the testing apparatus (>1.6 MPa). These results demonstrate the usefulness of dendrimer/PAA interfaces for promoting adhesion. Future work with these systems will include examination of how thick the dendrimer/PAA layers on the two wafers need to allow wafer bonding. Thinner layers will be desirable from many bonding applications. More work also needs to be done to examine the mechanism of bonding in these systems. Some information about this could be gained by changing the functionalities at the dendrimer surface and examining whether adhesion still occurs. Investigation of the use of other polycations to promote adhesion at these interfaces would also be illustrative. Additionally, although there are technical challenges, a spectroscopic investigation of the interface would provide insight into the chemistry that is occurring. 110 4.5 References (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11) (12) (13) (14) (15) (16) (17) (18) Amott, D. R.; Kindermann, M. R. J. Adhesion 1995, 48, 101-119. de Neve, B.; Shanahan, M. E. R. J. Adhesion 1995, 49, 165-176. Okamoto, Y.; Philip, K. J. Adhesion 1993, 40, 81-91. Choi, W. B.; Ju, B. K.; Lee, Y. H.; Haskard, M. R.; Oh, M. H. J. Vac. Sci. T echnol. B 1997, 15, 477-481. Niklaus, B; Anderson, H.; Enoksson, P.; Stemme, G. Sensors and Actuators A 2001, 92, 235-241. Schmidt, M. A. Proceedings of the IEEE. 1998, 86, 1575-1585. Tong, Q. Y.; Cha, G.; Gafiteanu, R.; Gosele, U. J. Microelectromech. Syst. 1994, 3, 29-30. Kim, T.; Chan, K. C.; Crooks, R. M. J. Am. Chem. Soc. 1997, 119, 189-193. Goss, C. A.; Charych, D. H.; Majda, M. Anal. Chem. 1991, 63, 85-83. Tong, Q. Y. Semiconductor Wafer Bonding: Science and Technology; John Wiley & Sons, Inc.: New York, 1998. Baker, L.; Zamborini, F. P.; Sun, L.; Crooks, R. M. Anal. Chem. 1999, 71, 4403-4406. Rar, A.; Zhow, J. N.; Liu, W. J .; Barnard, J. A.; Bennett, A.; Street, C. S. Appl. Surface Sci. 2001, 175, 134-136. Street, C. 8.; Rar, A.; Zhou, J. N.; Liu, W. J.; Barnard, J. A. Chem. Mater. 2001, 13, 3669-3677. Kim, B. Y.; Bruening, M. L. Langmuir 2002, in press. Dubas, S. T.; Schlenoff, J. B. Macromolecules 1999, 32, 8153-8160. Schlenoff, J. B.; Dubas, S. T. Macromolecules 2001, 34, 592-598. Shiratori, S. S.; Rubner, M. F. Macromolecules 1998, 33, 4213-4219. Hoogeveen, N. G.; Cohen Stuart, M. A.; Fleer, G. J. Langmuir 1996, 12, 3675-3682. 111 Chapter 5 CONCLUSIONS AND FUTURE WORK In this dissertation, we have developed of hyperbranched poly(acry1ic acid) (PAA) and layered hyperbranched dendrimer films as membrane skins and adhesives. The work performed in chapter 2 shows that synthesis of hyperbranched PAA films by a grafi-on-grafi strategy yields a selective, ultrathin membrane skin. F ield-emission scanning electron microscope and atomic force microscope images demonstrate that hyperbranched PAA effectively covers porous alumina to form an ultrathin, selective layer. Derivatization of PAA films with H2NCH2(CF2)6CF3 results in increased Oz/Nz and COz/CH4 selectivities for these composite membranes. Upon fluorination, the fluxes of Oz and C02 increase, while the fluxes of other gases such as CH4 and N2 are comparatively unaffected. The selectivity of fluorinated PAA membranes shows that they are free of defects, as even a small percentage of defects would negate selectivity. Because hyperbranched polymer membranes can be derivatized to contain a wide variety of functional groups, they will be attractive for applications where functional, defect-free membranes are required. Although deposition of PAA films on porous alumina substrates yields defect-free materials, the alumina support is, in general, too fragile for practical applications. In the future, it may be possible to construct composite films on polymeric substrates that have increased mechanical stability and durability. Possible supports include porous polysulfone and polycarbonate,"5 but the surface of these materials would first have to be functionalized.°’7 Both Crooks and Bergbreiter showed that hyperbranched films can be formed on functionalized polymeric surfaces.8'9 Should a more continuous surface be 112 needed, films could be deposited on poly[1-trimethylsilyl-1-propyne] (PTMSP) as PTMSP is a glassy, highly permeable support that presents a continuous surface.lO Coverage of underlying pores will not be a problem in this system. However, when working with some polymeric supports, the synthetic procedure will need some modification because the substrates may be sensitive to the organic solvents used in processing. To continue to examine the transport properties of hyperbranched PAA, transport of neutral molecules such as ortho-xylene, meta-xylene and para-xylene could be studied using a non-aqueous solvent. Using an organic solvent such as cyclohexane may allow the xylene isomers to be separated by steric effects, showing that hyperbranched PAA membranes are selective to an organic system. In chapter 3 of this dissertation, the thicknesses and morphology of multilayer dendrimer/PAA films showed the strong effect of deposition pH and dendrimer generation on film formation. The use of a deposition pH of 8 for dendrimers and 4 for PAA resulted in G4 dendrimer/PAA bilayer thickness values as high as 2,000 A, while G8 dendrimer/PAA bilayers were as thick as 4,000 A. At these solution pH values, the relatively low charge densities of both polymers are important factors contributing to the rapid grth of these films. Although both dendrimer/PAA and hyperbranched dendrimer/PAA films contain hyperbranched polymers, the structures of these two types of films are very different. Hyperbranched PAA does not have the controlled architecture that dendrimers do, and thus gas separation membranes prepared from PAA behave similarly to conventional polymers.” In contrast, dendrimer/PAA films are extremely permeable to gases. Thus 113 these films may prove to be very useful supporting layers for even thinner selective layers. For example, Sullivan and Bruening recently demonstrated the formation of ultrathin (35-40 nm) polyimide membranes using alternating polyelectrolyte deposition.12 If such films can be deposited on uniform, dendrimer/PAA films, perhaps defect-free polyimide films as thin as 10 nm could be produced. The wafer bonding studies in chapter 4 suggest that dendrimer mediated adhesion is stronger than that obtained with a commercial epoxy glue. However, much more work needs to be done in this area. First, we would like to limit the number of dendrimer/PAA layers that need to be deposited. This would allow fewer processing steps as well as a thinner adhesive layer. Second, we do not yet know the mechanism of adhesion. Do dendrimers penetrate into each of the PAA surface layers? Further work using XPS results and IR studies is needed in this area. In summary, this dissertation shows the promise of hyperbranched films for both separation and adhesion studies. However, further work is needed to better understand these materials and bring applications to fruition. 114 5.1 References (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11) (12) Hopkins, J .; Badyal, J. P. S. Langmuir 1996, 12, 4205-4210. Schonenberger, C.; van der Zande, B. M.; Fokkink, L. G. J .; Henny, M.; Schmid, C.; Kruger, M.; Bachtold, A.; Huber, R.; Birk, H.; Staufer, U. J. Phys. Chem. B. 1997, 101, 5497-5505. Ramamoorthy, M.; Raju, M. D. Ind. Eng. Chem. Res. 2001, 40, 4815-4820. Reid, B. D.; Ruiz-Trevino, F. A.; Musselman, I. H.; Balkus, K. J ., Jr.; Ferraris, J. P. Chem. Mater. 2001, 13, 2366-2373. Ghosh, K.; Chem, R. T.; Freeman, B. D.; Daly, W. H.; Negulescu, I. I. Macromolecules 1996, 29, 4360-4369. Dauginet, L.; Duwez, A. S.; Legras, R.; Demoustier-Champaign, S. Langmuir 2001, 1 7, 3952-3957. Kim, I. W.; Lee, K. J.; Jho, J. Y.; Park, H. C.; Won, J.; Kang, Y. S.; Guiver, M. D.; Robertson, G. P.; Dai, Y. Macromolecules 2001, 2908-2913. Dermody, D. L.; Peez, R. F.; Bergbreiter, D. E. Langmuir 1999, 15, 885-890. Zhao, M.; Liu, Y.; Crooks, R. M.; Bergbreiter, D. E. J. Am. Chem. Soc. 1999, 121, 923-930. Parthasarathy, A.; Brumlik, c. J .; Martin, C. R.; Collins, G. E. J. Membr. Sci. 1994, 94, 249-254. Toy, L. G.; Freeman, B. D.; Spontak, R. J .; Morisato, A.; Pinnau, I. Macromolecules 1997, 30, 4766-4769. Sullivan, D.; Bruening, M. L. Chem. Mater. 2002, in press. 115 I lllllllll lllllllllllllllllllllll'lll‘lllllllll’l 1293 02328 84