‘1- ..it ‘ a... i . W@; 34...“. . k. . v. .«mfiufix 1 .15: V ‘ 3 £3 , 5. ,. a I!" i :3 .z . . a. .3 A L..\. 1 ‘ r‘ \ :3 . .l. . : .3 K a} h wwwra...a‘h 6,43 a u- . it}: .. .mqufiwi ‘I' . I ,. . L» A urn. Jnflgm.‘ . , .. xuvmmwmrkwwbfi $1.2 . . E. 33%.. a. a uni}. 5.5.3.56” 1. .3: .x .5ch $1.45. 5.. 243.... LIBRARY Michigan State University This is to certify that the dissertation entitled USING NANOPARTICLES TO INHIBIT DEWETI'ING IN THIN POLYMER FILMS FOR POTENTIAL APPLICATIONS IN CHEMICAL SENSORS presented by MELISSA ANN HOLMES has been accepted towards fulfillment of the requirements for the Ph. D. degree in Chemical Engineering and Materials Science W/ W/?/ Major Professor’s’ Signature 6/40? Date MSU is an affinnative—action, equal-opportunity employer o.-a-n-n-o-o-O-.-o-o-u-O-0-I-—p--o-I-t-o---D-0-0-0-l-I-D-I-l-I-l-I-D-u-a-o—n-.---0-l-l-.-l-I-O-I-I-0-0-I-0-0-0-3-I-I-0-0- 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 6/07 p:ICIRC/DateDue.indd-p.1 USING NANOPARTICLES TO INHIBIT DEWE'I'I'ING IN THIN POLYMER FILMS FOR POTENTIAL APPLICATIONS IN CHEMICAL SENSORS By Melissa Ann Holmes A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemical Engineering and Materials Science 2007 ABSTRACT USING NANOPARTICLES TO INHIBIT DEWETTING IN THIN POLYMER FILMS FOR POTENTIAL APPLICATIONS IN CHEMICAL SENSORS By Melissa Ann Holmes Thin polymeric films are used in many new and emerging technologies, including nanolithography resist layers, fuel cells, dielectric coatings and chemical and biological sensors. Successful use of these materials requires that they be continuous, of uniform thickness and remain stable on a variety of organic and inorganic substrates. However, polymer films above their glass transition temperature or plasticized through solvent absorption may dewet from a substrate. The stabilization of these films by addition of various nanoparticles such as fullerenes (060) and polystyrene nanoparticles was investigated in this study. Previous researchers have shown that addition of fullerenes, dendrimers and polystyrene nanoparticles to thin polystyrene films (<50 nm) inhibits dewetting upon high temperature annealing. In this work the dewetting behavior of nanoparticle filled polymer films was investigated upon solvent annealing, similar to that seen in the manufacture and use of chemical sensors. It was found that under certain conditions the thin films were stabilized by the addition of a small amount of nanoparticles upon exposure to a solvent atmosphere just as they are stabilized during thermal annealing. This stabilization is due to the nanoparticles segregating to the substrate surface upon manufacture of the polymer film, forming a jammed state or gel-like network which makes the system stable. In addition, the effect of surface energy on the ability of nanoparticles to inhibit dewetting was investigated. In final studies, nanoparticle-filled polymer films were used in live sensors where the reliability and robustness were studied. This dissertation is dedicated to my family and friends who have supported me over the years. iv ACKNOWLEDGEMENTS I would like to thank my family for all of their love and support. I would like to thank Allan for his love and his patience with me during the stressful times, and for always being willing to let me vent to him about my frustrations. I would like to thank my mom for teaching me to love learning as a young child and for bestowing in me that I could achieve anything I put my mind to. I would like to thank my dad for going above and beyond to make sure I had everything I could every need or want, and for always helping me out whenever I asked. I would like to thank my sister, Michelle, for being such a good friend and roommate to me and helping me survive grad school. I would like to thank Dr. Mackay for his support and guidance over the past 5 years. He has given me a wealth of knowledge of which I am extremely grateful for. Also, I would like to thank Rachel Giunta for being a mentor to me and for helping me survive outside of the walls of academia. Finally, I would like to thank the Mackay Group members for all their help in the lab. I would like to especially thank Tiffany and Dave who have literally been at my sides during the entire journey that was grad school. Thank you to Tiffany for running with me to give me an excuse to get out of the lab. Thank you to Dave for always answering my numerous questions. Table of Contents List of Tables ................................................................................................... viii List of Figures ....... - .................................................................... x Key to Symbols or Abbreviations ................................................................... xv Chapter 1 Introduction ....................................................................................... 1 1.1. Motivation ........................................................................................... 1 1.2. Background ......................................................................................... 3 Chapter 2 Dewetting of Polymer Films under Solvent Annealing .................. 6 2.1. introduction ......................................................................................... 6 2.2. Experimental Method .......................................................................... 9 2. 3. Results and Discussion .................................................................... 11 2.4. Conclusion ........................................................................................ 28 Chapter 3 Dewetting Inhibition of Polymer Films Using Nanoparticles ...... 29 3.1. Introduction ....................................................................................... 29 3.2. Experimental ..................................................................................... 31 3.3. Results and Discussion .................................................................... 35 3.3.1. Polystyrene/Fullerene Systems .................................................... 35 3.3.2. Polystyrene/Polystyrene Nanoparticle Systems ........................... 47 3.4. Conclusion ........................................................................................ 53 Chapter 4 Influence of Substrate Surface Energy on the Dewetting of Polymer/Nanoparticle Films ............................................................................ 54 4.1. Introduction ....................................................................................... 54 4.2. Experimental ..................................................................................... 56 4.3. Results and Discussion .................................................................... 60 4.3.1. Polystyrene/Polystyrene Nanoparticle Systems ........................... 60 4.3.2. Polystyrene/Fullerene Systems .................................................... 68 4.4. Conclusion ........................................................................................ 76 Chapter 5 Application of Technology to Chemical Sensors ......................... 77 5.1. introduction ....................................................................................... 77 5.2. Experimental ..................................................................................... 79 5.3. Results and Discussion .................................................................... 81 5.3.1. Dewetting Inhibition ..................................................................... 81 5.3.2. Sensor Performance .................................................................... 85 5.4. Conclusion ........................................................................................ 89 Chapter 6 Summary and Conclusions ............................................................ 90 vi Appendices ............................................................... 94 Appendix A Principles of Neutron Reflectivity .............................................. 95 Appendix B Neutron Reflectile Data Interpretation .................................. 102 Appendix C Additional Data .................................................................... 105 vii List of Tables Table 2.1. Viscosities based upon dewetting data and rheological characterization for the different molecular weights tested. ................................ 17 Table 2.2. Cross model parameters for the different molecular weight PS samples used. .................................................................................................... 23 Table A.1. Scattering length densities (SLDs) for the materials used in the reflectivity studies. ............................................................................................ 101 Table 8.1. Known parameters used in all reflectivity models. .......................... 104 . Table C.1. Fitted values for the model parameters used in Figure 0.1. ........... 107 Table 0.2. Fitted values for the model parameters used in FigureC.2. ............ 109 Table C.3. Fitted values for the model parameters used for the piranha cleaned sample in Figure 0.3. .......................................................................... 111 Table C.4. Fitted values for the model parameters used for the Sigmacote® treated sample in Figure 0.3 ........................................................ 111 Table C.5. Fitted values for the model parameters used in Figure 0.4. ........... 113 Table 0.6. Fitted values for the model parameters used for the before annealed sample in Figure 0.5 ......................................................................... 115 Table 0.7. Fitted values for the model parameters used for the after annealed sample in Figure 0.5 ......................................................................... 115 Table 0.8. Fitted values for the model parameters used for the normal film before annealing in Figure 3.2. ......................................................................... 116 Table 0.9. Fitted values for the model parameters used for the normal film after annealing in Figure 3.2. ............................................................................ 116 Table 0.10. Fitted values for the model parameters used for the flipped film before annealing in Figure 3.2. .................................................................. 116 Table (3.11. Fitted values for the model parameters used for the film before annealing in Figure 3.4. ......................................................................... 117 Table C.12. Fitted values for the model parameters used for the film after annealing in Figure 3.4. .................................................................................... 117 viii Table C.13. Fitted values for the model parameters used for the film during annealing in Figure 3.5. ......................................................................... 118 Table C.14. Fitted values for the model parameters used for the before annealed sample in Figure 3.7. ........................................................................ 119 Table C.15. Fitted values for the model parameters used for the before annealed sample in Figure 3.7. ........................................................................ 119 Table C.16. Fitted values for the model parameters used for the film before annealing in Figure 4.8. ......................................................................... 120 Table C.17. Fitted values for the model parameters used for the film after annealing in Figure 4.8. .................................................................................... 120 Table C.18. Fitted values for the model parameters used for the flipped film before annealing in Figure 4.9. .................................................................. 121 Table 0.29. Fitted values for the model parameters used for the flipped film after annealing in Figure 4.9. ..................................................................... 121 ix List of Figures Figure 2.1. Optical micrographs showing the progression of dewetting for various molecular weight PS films arranged in columns on silicon substrates under solvent annealing with toluene over time. (a) 5 k0 PS, (b) 19 kD PS, (c) 75 kD PS. Films are all ~36 nm in thickness. From the images one can see the instability of the rims increase with increasing molecular weight. Scale bars are 100 um and the scale is the same for all images. ............................................................................................................... 12 Figure 2.2. Hole growth data for 5 RD, 19 kD and 75 kD PS samples over time. All films are ~36 nm thick. The plot shows that the holes grow at a constant velocity and the rate is dependent on the molecular weight of the film ...................................................................................................................... 13 Figure 2.3. Dynamic frequency sweeps of different molecular weight saturated PS samples with their Cross model fits. ............................................. 16 Figure 2.4. (a) Hole growth data for 5 k0 PS samples of various thicknesses over time. Film thicknesses are given in the legend. (b) Plot of the dewetting rates versus film thickness for the 5 kD PS films .......................... 20 Figure 2.5. (a) Hole growth data for 19 kD PS samples of various thicknesses over time. Film thicknesses are given in the legend. (b) Plot of the dewetting rates versus film thickness for the 19 kD PS films ........................ 21 Figure 2.6. (a) Hole growth data for 75 kD PS samples of various thicknesses over time. Film thicknesses are give in the legend. (b) Plot of the dewetting rates versus film thickness for the 75 kD PS films ........................ 22 Figure 2.7. Optical micrographs of the evolution of a dewetting hole for a 75 kD PS film. The fingering instabilities of the rim are apparent as the hole grows over time from 5 minutes (left) to 10 minutes (middle) to 20 minutes (right). Scale bars are 50 um. ............................................................... 26 Figure 3.1. a) Pure PS film after annealing with a saturated toluene atmosphere for 1 hour. b) PS with 3% C60 by mass after annealing with a saturated toluene atmosphere for 3 hours. Films are ~30 nm thick. ................... 35 Figure 3.2. a) Neutron reflectivity data and models for polystyrene films with 3 wt% C60. Normal films refer to films that were spin-coated directly onto silicon substrates while the flipped film was spin-coated onto mica and then flipped onto a silicon substrate. b) Scattering length density profiles for the models used to fit the data. Reflectivity data were shifted to allow for easier viewing. ..................................................................................... 37 Figure 3.3. Optical micrographs of spin-cast films (~30 nm) of PS with various concentrations of 060 after exposure to a saturated toluene atmosphere. a) Pure PS after 1 hour. b) PS + 0.1 wt% Cso after 5 hours. c) PS + 3 wt% Ceo after 3 hours. d) PS + 10 wt% 060 after 1 hour. TEM micrographs of e) PS film with 3 wt% Cso and f) PS film with 10 wt% Ceo (inset shows diffraction pattern for a cluster of fullerenes). ................................ 40 Figure 3.4. a) Neutron reflectivity data and models and b) scattering length density profiles for a polystyrene film with 7 wt% C50. The film was measured before (a) and after (0) annealing under a saturated toluene atmosphere. The film was ~18 nm thick. After annealed reflectivity data were shifted up a decade to allow for easier viewing. ........................................ 43 Figure 3.5. a) Neutron reflectivity data and model and b) scatting length density profile for the same film as Figure 3.4 during exposure to a saturated toluene atmosphere. The horizontal lines correspond to the SLD of pure d-Toluene and PS. ................................................................................. 44 Figure 3.6. PS + 10 wt% 25 kD PSNP film after annealing with a saturated toluene atmosphere for 5 hours. Inset shows film before annealing. Film is ~30 nm thick. ...................................................................................................... 47 Figure 3.7. a) Neutron reflectivity data and models for two polystyrene films with 10 wt% 25 kD PSNP. One film is immediately after spin-coating (o) and the other has been exposed to a saturated toluene atmosphere for 5 hours (0). b) Scattering length density profile of the models used to fit the data. The SLD for pure dPS and PSNP are shown on the graph. For the after annealing sample, it is clearly seen from the SLD profile that the nanoparticles migrate to the substrate, as shown. ............................................. 48 Figure 3.8. Optical micrographs of PS films with varying concentrations of 25 kD PSNP annealed under a saturated toluene atmosphere for 1 hour. a) Pure PS film. b) PS + 5% 25 kD PSNP film. 0) PS + 15% 25 kD PSNP film. d) PS + 50% 25 kD PSNP film. All films are ~30 nm thick. ........................ 50 Figure 4.1. Optical micrographs of PS + PSNP films (~30 nm thick) on piranha cleaned wafers annealed for 5 hours under a saturated toluene atmosphere. a) PS + 17% 60 kD PSNP (0:1.2). b) PS + 23% 140 kD PSNP (0:12). C) A schematic of how the monolayer of PSNP stabilizes the film against the adverse van der Waals forces. The arrows represent the extent of the van der Waals force interaction. .............................................. 62 Figure 4.2. Optical micrographs of a pure PS film (~30 nm thick) on a silanized water after annealing under a saturated toluene atmosphere for 15 minutes. Inset shows film prior to annealing. ................................................. 63 xi Figure 4.3. Optical micrographs of PS + 25 kD PSNP films on a silanized wafer after annealing under a saturated toluene atmosphere. a) PS 4» 5% 25 kD PSNP after annealing for 15 minutes (0:05). b) PS + 10% 25 kD PSNP after annealing for 15 minutes (0:1.0). 0) PS + 15% 25 kD PSNP after annealing for 5 hours (9:1.4). d) PS + 26% 25 kD PSNP after annealing for 5 hours (0:25). All films are ~30 nm thick .................................... 64 Figure 4.4. Optical micrographs of PS + 60 kD PSNP films on silanized wafers annealed under a saturated toluene atmosphere. a) PS + 12% 60 kD PSNP after annealing 30 minutes (0:09). b) PS + 20% 140 kD PSNP after annealing 1 day (0:1.4). Films are ~30 nm thick ........................................ 65 Figure 4.5. Optical micrographs of PS + 140 kD PSNP films on silanized wafers annealed under a saturated toluene atmosphere. a) PS + 16% 60 kD PSNP after annealing 30 minutes (0:09). b) PS + 23% 140 kD PSNP after annealing 20 hours (0:12). Films are ~30 nm thick. ................................. 66 Figure 4.6. Optical micrographs of PS + 3 wt% C60 films (~30 nm thick) spin-coated onto mica and floated onto a) a piranha cleaned wafer and annealed for 5 hours and b) a silanized wafer and annealed for 15 minutes. Annealing was performed by exposing the films to a saturated toluene atmosphere. Insets show films prior to annealing. ................................. 68 Figure 4.7. Optical micrographs of PS + 3 wt% C60 films (~30 nm thick) spin-coated onto mica and flipped onto a) a piranha cleaned wafer and annealed for 1 hour and b) a silanized wafer and annealed for 15 minutes. Annealing was performed by exposing the films to a saturated toluene atmosphere. Insets show films prior to annealing ............................................... 70 Figure 4.8. a) Neutron reflectivity data and models for a PS + 60 kD PSNP + 3 wt% Cso flipped film before and after annealing under at saturated toluene atmosphere for 2.5 hours. b) The scattering length density profiles for the models used in a). The film was formulated to ensure that the PSNP are at sufficient concentration to form a monolayer. ....... 72 Figure 4.9. a) Neutron reflectivity data and models for a PS + 140 kD PSNP + 3 wt% C60 film before and after annealing under a saturated toluene atmosphere for 5 hours. b) The scattering length density profiles for the models used in a). The film was formulated to ensure that the PSNP are at sufficient concentration to form a monolayer. ................................ 73 Figure 5.1. Optical micrographs of spin-cast (a) pure PECH and (b) PECH + 1 wt% Ceo films after exposure to a saturated methanol atmosphere for 20 hours. Insets show films prior to annealing ................................................... 81 xii Figure 5.2. Optical micrographs of spray-coated films after exposure to a saturated methanol atmosphere. (a) Pure PECH after 42 hours and (b) PECH + 10 wt% C50 after 47 hours of exposure. Insets show films prior to annealing. ........................................................................................................... 83 Figure 5.3. Optical micrographs of spray-coated (a) pure DKAP and (b) DKAP + 10 wt% C50 films after exposure to a saturated methanol atmosphere for 17 hours. Insets show films prior to annealing ........................... 84 Figure 5.4. Optical micrograph of a live SAW sensor. The left side (Line 1) is coated with PECH + 10 wt% Cso. The right side (Line 2) is coated with pure PECH. ................................................................................................. 85 Figure 5.5. Response curves for a live SAW sensor. (a) Response to challenges for line 2 coated with pure PECH (A) immediately after spray- coating and (B) after sitting for 1 day. (b) Response to challenges for Line 1 coated with PECH with 10 wt% 060 (C) immediately after spray-coating, (D) after sitting for 3 days and (E) after spraying more solution 4 days after the initial deposition. ........................................................................................... 87 Figure A.1. Schematic of the POSY2 reflectometer at Argonne National Laboratory. ......................................................................................................... 98 Figure A.2. Schematic of SPEAR at Los Alamos National Laboratory. ............. 98 Figure A.3. Schematic of the reflection and refraction of a beam of neutrons through a thin film (1) on a substrate (2) in air (0) ................................ 99 Figure 3.1. Illustration of a reflectivity model for a film having m layers, each with a distinct thickness (d) and neutron refractive index (n). .................. 102 Figure 8.2. Image of the reflectivity panel window in Motofit. .......................... 104 Figure 6.1. a) Neutron reflectivity data and model for a polystyrene film on a piranha cleaned silicon wafer. b) Scattering length density profile of the model used to fit the data. ................................................................................ 106 Figure C.2. a) Neutron reflectivity data and model for a deuterated polystyrene film on a piranha cleaned silicon wafer. b) Scattering length density profile of the model used to fit the data. ............................................... 108 Figure C.3. a) Neutron reflectivity data and models for deuterated polystyrene film. One was spin-coated onto a piranha cleaned silicon water while the other was floated onto a Sigmacote® treated wafer. b) Scattering length density profiles of the models used to fit the data. ................ 110 xiii Figure C.4. a) Neutron reflectivity data and models for polystyrene with 3 wt% 060 films on piranha cleaned silicon wafers. One was tested without annealing and the other was first annealed with a saturated toluene vapor. b) Scattering length density profiles of the models used to fit the data ............. 112 Figure C.5. a) Neutron reflectivity data and models for polystyrene with 3 wt% C60 films on piranha cleaned silicon wafers. One was tested without annealing and the other was first annealed with a saturated toluene vapor. b) Scattering length density profiles of the models used to fit the data ............. 114 xiv ACS i’sr. Key to Symbols or Abbreviations Nanoparticle radius Degrees of freedom gained by releasing a monomer unit from substrate constraints Hamaker constant for a 3 layer assembly American Chemical Society Coherent scattering length Background added to the reflectivity fit Concentration Parameter used to determine the goodness of a fit Fullerenes Film thickness Film thickness Diffusivity Characteristic diameter Diameter of a growing hole A poly(dimethylsiloxane) derivative Dimethyl methyl phosphonate Deuterated polystyrene Monomeric interaction energy Volume fraction of i Surface tension Liquid surface tension Solid surface energy Interfacial energy between a solid and liquid Shear rate Film thickness Characteristic film thickness defined as the absolute value of the spreading parameter divided by the plateau modulus Initial film thickness Packing density Viscosity Complex viscosity Terminal viscosity Specific viscosity Intrinsic viscosity Hydrofluoric acid Dimensionless constant Huggins constant Boltzmann constant Constant equal to 3.2x10'8 ° '3 Kilodalton Coefficient dependent on n, kand h,- XV I], j+1 rpm .02 0'0 SAW SLD Neutron wavelength Modulated differential scanning calorimetry Methanol Methyl salicylate Weight average molecular weight Power law index Atomic number density Neutron refractive index of medium j Polydispersity index (weight average molecular weight/number average molecular weight) Polyepichlorohydrin Polystyrene Polystyrene nanoparticles Polytetrafluoroethylene (commonly referred to as Teflon“) Surface area coverage Angle of incidence Critical angle of incidence below which total external reflection occurs Equilibrium contact angle Angle of incidence Neutron momentum transfer Critical value of qz below which total external reflection occurs Neutron momentum transfer perpendicular to the surface Reflectivity Fresnel reflection coefficient Revolutions per minute Scattering length density at depth 2 Kuhn or monomer length Roughness Plateau modulus Surface acoustic wave Scattering length density Time Relaxation time Crossover time, below which Dd scales linearly with t Equilibration time Temperature Glass transition temperature Polymer glass transition temperature Solvent glass transition temperature Transmission electron microscopy Rate of dewetting Weight fraction Weight percent Coefficient dependent on n ranging from 1 to 00 Depth in the direction perpendicular to the surface xvi Chapter 1 Introduction 1.1. Motivation In the aftermath of the attacks on September 11, 2001, the level of concern and awareness of a terrorist attack has significantly risen. Specifically, concerns over the contamination of the nation’s water supply as well as harmful agents being released into the atmosphere have been discussed. These types of chemical and biological warfare have the possibility to kill millions of people and thus, the need for a way to detect and protect against such possible attacks has surfaced. Numerous types of chemical and biological sensors are currently being developed. Many of these sensors employ the use of thin polymer films to aid in the detection of harmful pathogens. At the forefront of this effort is Sandia National Laboratories, which has developed a chemical sensor it calls the pChemLabTM. This sensor is a portable chemical analysis system consisting of three different stages. The first stage of the device collects and concentrates the vapor sample. Next, a micro-gas chromatograph separates the sample components based on size. Finally, the sample is fed to the sensing stage. The sensor used isva surface acoustic wave (SAW) device which uses a thin polymer film deposited on the quartz as the detector. The polymer film is designed to selectively absorb a specific analyte and is deposited onto the substrate along the wave’s propagation path. When the analyte is present, it is absorbed by the polymer film causing a minute increase in the film’s mass that can then be measured by a shift in the SAW operating frequency relative to an uncoated reference device. 1.2. Background Thin polymeric films (<100 nm) are used in many of today’s new and emerging technologies. These include nanolithography resist layers, fuel cells, dielectric coatings and chemical and biological sensors to name a few." 2 For the majority of the applications, successful use of these materials requires that they be continuous, of uniform thickness and remain stable on a variety of organic and inorganic substrates. However, for films of this thickness, the film may be unstable or metastable, leading to the dewetting (or beading up) of the polymer on the substrate under the appropriate conditions. Dewetting can occur either as a spinodal process3 or by nucleation and hole growth.4 In the first, thermal fluctuations at the polymer-air interface increase in amplitude until they reach the substrate and create a hole. In the latter, a nucleation site is formed due to either defects or impurities in the film. In both cases, once the holes are formed, they continue to grow in size until they coalesce to form a honeycomb of polymer fibrils, eventually decaying into spherical droplets through Rayleigh instabilities.5 In order for a polymer film to dewet, it must be in a liquid state, and thus above its glass transition temperature, or T9. One method to induce dewetting is to thermally anneal the polymer film well above its T9, thereby allowing it to flow. In this case, long-range van der Waals forces lead to the polymer film instability and its eventual breakup.6' 7 Another method to induce dewetting is to lower the effective T9 of the polymer film by vapor annealing. In this case, solvent vapor is introduced into the system effectively plasticizing the film, leading to a depression in the Tg as well as decreasing the viscosity.8 Vapor annealing differs from thermal annealing by the addition of a third component to the system, increasing the possible surface and bulk interactions, and to date has been given very little aflenfion. In the present work, we look at the issues associated with polymer film dewetting that occur upon vapor annealing, similar to the situation for sensor applications. The polymer films are exposed to a saturated solvent atmosphere which plasticizes the polymer and reduces its T9, allowing dewetting to occur. This type of dewetting is seen in the use of chemical sensors,9 such as surface acoustic wave (SAW) sensors, which are devices with an input and output transducer patterned on a piezoelectric substrate. When the input transducer is excited, it generates a surface acoustic wave that propagates across the substrate creating an output signal at the receiving transducer. To make the device a sensor, a polymer coating designed to selectively absorb a specific analyte is deposited onto the substrate along the propagation path. When the analyte is present, it is absorbed by the polymer coating causing a minute increase in the mass of the film that can then be measured by a shift in the SAW operating frequency relative to an uncoated reference device. When dewetting occurs in a sensor, the output frequency can become erratic or stop altogether due to the reduced analyte detection area, leading to a faulty sensor.9 The typical approaches for stabilizing thin polymer films often involve modification of the polymer, the substrate or the interaction between the two. These may include crosslinking of the polymer,10 sulfonation and metal 1 surface treatment of the substrate to alter its complexation of the polymer,1 surface energy and thus promote wetting12 or grafting of the polymer chains directly to the substrate.13 However, for some applications it is not feasible to utilize these methods as they may impact the desired properties of the film or hinder the ability for rapid manufacture of devices. Consequently, an alternate method for eliminating dewetting is desired. Recently it has been shown that addition of fullerenes,14 dendrimers15 and other nanoparticles‘a'18 to polymer films inhibits dewetting during thermal annealing. These studies suggest that the nanosized fillers segregate to the substrate surface, which leads to stabilization of the film. To support this, recent simulations have shown that the distribution and mobility of the nanoparticles as well as the polymer-filler attraction are important factors in the inhibition of dewetting.19 This stabilization effect is hypothesized to be the result of contact line pinning by fractal-like structures formed at the substrate as well as due to modification of the polymer-substrate interaction as a result of an increased surface (nano)roughness and the concomitant surface energy modification due to a fractal-like polymer molecule conformation.“ In the work presented here, this novel strategy of using nanoparticles to inhibit dewetting is adopted for use in vapor annealed polymer films such as those in chemical sensors. This technique will be used with a standard, well characterized polymer (polystyrene) as well as polymers associated with sensor applications to determine if the above hypothesis is valid. In addition, sensor performance will be tested with and without nanoparticles to determine their influence. Chapter 2 Dewetting of Polymer Films under Solvent Annealing 2.1 . Introduction The stability of thin polymer films (<100 nm) has been studied at length due to their numerous applications in many new and emerging technologies such as dielectric coatings, nanolithography resist layers and chemical and biological sensors, to name a few." 2 The applications of these films require that they be continuous, of uniform thickness and remain stable on a variety of substrates. However, depending on the intermolecular forces present, as well as the wettability of the liquid on the substrate, the film may dewet. Currently, there are two separate mechanisms which are believed to initiate the onset of dewetting. One mechanism involves the influence of heterogeneous defects or impurities in the film, such as dust particles." 2° These defects create a dry patch, or nucleation site, where dewetting can begin. The other mechanism is a spontaneous rupturing of the film, or spinodal dewetting,3' 21'2“ due to the fact that it is comparable to spinodal decomposition seen in fluid mixtures. Here, modulations at the free surface of the film induce a pressure gradient resulting from a competition between the Laplace pressure that will level the film, and the disjoining pressure which thickens and roughens the film. Depending on the system, these modulations may be amplified exponentially until they reach the substrate, spontaneously rupturing the film. In either case, heterogeneous or homogeneous dewetting, once the holes are formed they continue to grow in size. Much work has been done investigating the dynamics of growing holes by considering the balance between the capillary driving forces and the viscous dissipation forces. For a sample where the viscous dissipation is dominant, one finds that the growth rate remains constant with time (0.25 On the other hand, if there is very little viscous dissipation and instead the energy is dissipated by friction, then the polymer has been seen to slip on the substrate.26' 27 In this case, the diameter of the hole grows proportional to 12’3 rather than linearly. Some studies have shown that there is a combination of these two cases, where slip and viscous dissipation are present in the hole gromrth.28' 29 Still, other work, involving the growth of holes in free-standing films, found an exponential growth of the hole size.3°' 3‘ Work has also been performed with regard to the rim stability of the growing holes. In most instances, as the holes grow, the removed polymer accumulates at the edges into rims. Eventually the holes will coalesce which results in the formation of a honeycomb of polymer ribbons. These ribbons then decay into spherical droplets through Rayleigh instabilities.5 Other studies have shown that the rims undergo fingering instabilities as they grow, such as those seen in Hele-Shaw flow patterns?" 32' 33 Although there has been much research in the area of polymer films dewetting from a silicon substrate, much of it has been done using high- temperature annealingf” 4' 20' 2" 33'” Little work has been done in the area of solvent annealed films. This is a current area of interest due to the fact that polymer films used in sensor applications are exposed to solvents and must remain stable to ensure satisfactory performance.38 In one previous study, the instability of the film was attributed to the short-range polar interactions as opposed to the long-range van der Waals forces.39 Also, they suggest there are two types of rim instabilities in solvent dewetting depending on the polarity of the solvent used. Dewetting of block copolymer films induced by solvent vapor has also been considered as well as the resulting morphologies.” 4‘ However, none of these studies have specifically focused on the dynamics of the dewetting holes over time. In the current work, we investigate the dewetting of thin polystyrene films on silicon substrates under a saturated toluene atmosphere. Solvent annealed - dewetting has been seen in the manufacture of sensors, and therefore the aim of this work is to better understand the stability and dynamics of polymer films exposed to solvent vapor. The dewetting dynamics of films are investigated for various molecular weights of polystyrene as well as different initial film thicknesses. In addition, a look at the instability of the moving rim is studied. This work differs from many of the other dewetting dynamic studies due to the fact that a third component is introduced, the solvent, which changes the interfacial energies of the system as well as the flow properties. 2.2. Experimental Method Polystyrene (PS) standards were purchased from Scientific Polymer Products, Inc. with weight average molecular weights (MW) of 5.1, 19.3 and 75.7 kD with respective polydispersity indexes (PDI, weight to number average mass ratio) of 1.07, 1.07 and 1.17. These molecular weights were chosen to allow for samples which are unentangled (5.1 kD and 19.3 kD) and entangled (75.7 kD) since the critical molecular weight for entanglement coupling is 35 kD. The polymers were used as received. Solutions were made using ACS reagent grade toluene. Stock solutions were first made and then diluted to the desired concentration and were filtered using a 0.2 pm PTFE filter before use. The substrates used for the experiments were silicon wafers with native oxide on their surface and were purchased from Wafer World Inc. The wafers were cleaved to the desired size (~1 cm x 1 cm) and then cleaned for 1 hour with an acid wash consisting of a 70:30 by volume mixture of 96% H2SO4 and 30% H202, respectively, to remove any residual organic contaminants. After the acid wash, the wafers were rinsed with copious amounts of Millipore water and dried under a nitrogen stream. Wafers were used immediately after cleaning. Thin films were made by spin-coating the polymer solutions onto the substrates at a speed of 5000 rpm for 40 seconds. Film thickness was controlled by the total mass concentration of the solution and ranged from 20-50 nm as determined using ellipsometry (JA Woollam M-44 ellipsometer). Films were found to be uniform and continuous upon inspection with optical microscopy and were annealed by exposing them to a saturated toluene environment in a closed vessel at room temperature to observe the dewetting process. Reflective mode optical images (bright field) were obtained in situ using an Edmund Industrial Optics optical microscope with a JAI camera attachment. Rheology experiments were carried out using a Rheometrics ARES rheometer using 8 mm parallel plates set at a gap of approximately 0.3 mm. Dynamic frequency sweeps in the range of 1-100 radians/second were carried out at room temperature (25°C 11°) under saturated vapor conditions. Bulk P8 was placed on the rheometer plates and a glass chamber filled with toluene vapor was placed around the sample and allowed to sit for several days to insure equilibrium had been achieved. Several measurements were taken and then averaged. 10 2.3. Results and Discussion In the experiments presented here, annealing of the polymer films was done by introducing solvent vaporinto the system, effectively plasticizing the film leading to a depression in the Tg as well as decreasing the viscosity.8 The evolution of holes for different molecular weight polystyrene films of the same thickness are shown in Figure 2.1 as they dewet from silicon substrates. After a certain time, holes spontaneously form and, as can be seen from the images for all the molecular weights tested, the holes have rims as soon as they are visible with the optical microscope. Also, one can see that the time for holes to form, as well as grow, varies significantly with molecular weight. The growth of hole diameters for the three molecular weight PS films studied are shown in Figure 2.2 where all films in this case are 36 :1: 1 nm thick. To insure an accurate value for the diameters of the holes, five measurements were taken at different orientations around the hole, neglecting any fingering that occurred, and then averaged to give the reported value. For all molecular weights a linear growth rate is observed, meaning that viscous forces dominate and there is no slip. The rate of dewetting is also shown to be dependent on the polymer molecular weight in agreement with previous studies which suggest that the rate of dewetting is inversely proportional to the viscosity.42 We of course find that the higher molecular weight sample, which is expected to have a higher viscosity, as discussed below, has a smaller dewetting rate. 11 Figure 2.1. Optical micrographs showing the progression of dewetting for various molecular weight PS films arranged in columns on silicon substrates under solvent annealing with toluene over time. (a) 5 k0 PS, (b) 19 kD PS, (0) 75 kD PS. Films are all ~36 nm in thickness. From the images one can see the instability of the rims increase with increasing molecular weight. Scale bars are 100 pm and the scale is the same for all images. 12 Hole Diameter (pm) 16030 Time (seconds) Figure 2.2. Hole growth data for 5 k0, 19 kD and 75 kD PS samples over time. All films are ~36 nm thick. The plot shows that the holes grow at a constant velocity and the rate is dependent on the molecular weight of the film. To more fully understand the flow mechanism of the dewetting films, experiments were performed using bulk samples of polystyrene powders of all three molecular weights to determine the amount of toluene absorbed by the polymers. Polystyrene was used in powder form as received, placed in a pan and weighed both before and after exposure to a saturated toluene atmosphere. Upon reaching equilibrium, these experiments show that the approximate equilibrium polymer concentration is 0.73 :I: 0.01 g/ml for all molecular weights after one week, and is assumed to be equal to the concentration of the PS/toluene films used for the dewetting experiments. We expect equilibrium to be achieved in a much shorter time in the thin film and the equilibration time can be estimated by 43 Dr A25 >1 . (2.1) where D is the toluene diffusivity in PS (~10'6 cm2/s) 44, A, the film thickness (~30 nm) and 15, the equilibration time. Inserting these values, a time of ~10 us is found and thus, solvent exposure creates an essentially instantaneously equilibrated film. The dewetting time scale is on the order of minutes and so it is expected that dewetting occurs under equilibrium concentration conditions. Knowing the concentrations of our films, the solution glass transition temperature can be estimated using the Fox equation“ _1_ _ w 1— w T -—P 3 T8 T8 T8 (2.2) Here, w is the weight fraction of the polymer in solution (~0.70), and T9,P and T5,S are the polymer (379 K for 75 kD PS) and solvent (117 K) ‘5 glass transition temperatures. At this concentration, the T9 is found to be -46°C. The fact that the T9 of the solution is well below room temperature, it is expected that the viscosity will be low and rationalizes why rapid dewetting occurs when the polystyrene film is exposed to a saturated toluene atmosphere even at room temperature. Further, the solution surface energy is expected to be ~31 mJ/m2 and similar to that of pure PS (~28.5 mJ/mz) under thermal annealing and so dewetting is expected.46 Thus, the flow behavior may be different during vapor exposure compared to thermal annealing since different final patterns are 14 produced, yet, dewetting is expected under surface energy arguments and the ultimate result is the same with the appearance of gross dewetting. To help quantify the dewetting time we first estimate the viscosity of the polymer/solvent mixture using the Martin equation47 for the 75 kD molecular weight sample :73 = ek'InIc 477] (2.3, where 175,, is the specific viscosity {solution viscosity/solvent viscosity - 1, where the solvent viscosity is taken as 0.56 GP for toluene at 25°C“}, 0, the polymer concentration (0.73 g/ml), [77], the polymer intrinsic viscosity (calculated to be ~38 ng using the Mark-Houwink-Sakurada equation with the appropriate constants for the 75 kD molecular weight polystyrene) and k', Huggins constant (~0.33).49 The resulting solution viscosity using this method is 150 Pa-s which is equivalent to the viscosity of pure 75 kD polystyrene at a temperature of ~210°C.5°' 5‘ For the two lower molecular weight samples the results obtained using the Martin equation were much lower with values of 0.2 Pa-s for the 19 kD PS and 7x10'3 Pa-s for the 5 k0 PS. Viscosity measurements were carried out under saturated vapor conditions to compare to those calculated using the Martin equation and are shown in Figure 2.3. We can see that the viscosity of the saturated films is dependent upon the molecular weight as expected, and in agreement with our results for the dewetting rates given in Figure 2.2 where higher molecular weights dewet slower. We were unable to reach the terminal region for the highest 15 ,I l 1 lllllll 1 111 ' T ' 17 I f I U ' U W V V : i o 75kDPS] " :, O 19kDPS 2 e ' u - A 5kDPS u _ , 4 u ._.., 10 .. n .. - n ’6; 4?" b n "‘ " ll (6 I " u Q. U " 9; 2 .'l‘.. Q; " H II a: 3 . “ " ._‘~\‘\“~.. " II II c 10 ' \\-[ ._. “ u ”- - “ \x u j. 5 ‘I “\ n 1 4 . “\..\“fi' 2 2' 102 . ' é éis'éié" 2'» :3 15655” 1 10 100 Frequency (rad/sec) Figure 2.3. Dynamic frequency sweeps of different molecular weight saturated PS samples with their Cross model fits. molecular weight polymer so we can only speculate as to its zero-shear viscosity by fitting with the Cross model discussed below (see Table 2.1), however, we estimated the viscosity of the films to be 2800, 6700 and 67 000 Pa-s, in order of increasing molecular weight. These are much higher than those calculated from predictions using the Martin equation, which may be explained by the fact that we have fairly concentrated solutions (~0.73 g/ml) while the Martin equation is valid for dilute or semi-dilute solutions. For no-slip conditions, as our experiments show, the rate of dewetting (V) is expected to be inversely proportional to the viscosity through the following relation:42 16 _ 3 V —k —7 9. where r; is the viscosity, y is the surface tension, 8., is the equilibrium contact . angle and k9 is a constant equal to 3.2x10‘8 degrees'a.52 Using the dewetting rates calculated from the data in Figure 2.2, a surface tension of 31 mN/m53 and an equilibrium contact angle of 15° (which was found for a sample of saturated PS on a silicon water), we can calculate the Viscosities for the three different molecular weight samples which are given in Table 2.1. From the table we can see that the Viscosities are less than those found in the rheological characterization experiments perhaps indicating that the viscosity of thin films may differ from the bulk“ as has been seen for various other bulk properties (i.e. glass transition temperature55'57 ) or that equation 2.4 is not applicable to our system. In Chapter 4 we will show through neutron reflectivity experiments that the concentration of the polymer in thin films upon absorption of toluene is much less than that calculated from the bulk experiments, which would lead to a lower Table 2.1. Viscosities based upon dewetting data and rheological characterization for the different molecular weights tested. Molecular Dewetting Dewetting Chiraetgtggzcaetiilon Weight Rate Data Viscosity Viscosity 75.7 0.015 220 67 000 19.3 0.13 25 6700 5.1 1 .43 2.3 2800 l7 viscosity compared to that measured by rheological characterization. This would explain why we would predict a much lower viscosity based on the dewetting rates. Consider now the influence of film thickness on the rate of dewetting for the polymer films. The dewetting hole growth data for 5 kD PS films of various initial film thicknesses is shown in Figure 2.4a. Holes rupture only a few minutes after being exposed to the saturated toluene atmosphere and over time grow in size. At shorter times, an interesting observation is made, the rate of dewetting appears to follow a nonlinear trend before becoming linear. This may be clue to the fact that the dewetting occurs so rapidly that it has not become an equilibrium process at this point, yet, ultimately the rate of hole growth becomes linear. To characterize the rate of dewetting, the data were fitted to a linear model at longer times neglecting the initial hole diameters which do not follow this trend. From this analysis the influence of the initial film thickness on the observed rate of dewetting is found to be inversely proportional to the initial film thickness under solvent annealing at odds with the thickness independent prediction of equation 2.4 (see Figure 2.4b). By comparison, the hole growth data for the higher molecular weight PS films, 19 kD and 75 kD, shown in Figures 2.5a and 2.6a, respectively, always show a linear trend. It takes much longer in these cases for the holes to rupture and once the holes are formed and visible, rims are also immediately observable. For all film thicknesses we again see a linear growth rate, yet for higher molecular weight films, we do not observe the brief non-linear dewetting at 18 shorter times as was seen in the 5 k0 PS films. The influence of initial film thickness on the rate of dewetting is shown in Figures 2.5b and 2.6b where again a linear trend is seen, similar to that in Figure 2.4b. Previous studies have considered the influence of film thickness on the rate of dewetting. In one study liquid-liquid dewetting was performed where the liquid substrate thickness is shown to significantly influence the rate of dewetting.58 The other study showed polystyrene films on silicon substrates of varying initial thicknesses had different rates of dewetting upon thermal annealing59 which was attributed to a difference in the viscosity of the films based on the film thickness. A purely Newtonian fluid (equation 2.4) would not give a film thickness effect, however, our results in Figure 2.3 suggest that the fluids exhibit a shear- thinning effect, which has been previously considered by Saulnier et al.60 Shear- thinning was investigated due to the fact that the extremely thin films experience high shear rates at even moderate dewetting rates. We can determine an apparent shear rate by dividing the rate of dewetting by the film thickness, which yields a shear rate on the order of 100 5". According to Saulnier et al., dewetting holes in thin films which experience shear-thinning behavior should grow linearly with time, and in addition, the rate of dewetting should be faster for thinner films, both of which are seen in our data. For shear-thinning, the viscosity of the film can be approximated using the Cross model: . n 2.5 1+4”) ( ) 19 A m v 500 cli- . -. ..... ,- ‘- I . 2 ; f i 2 I ..t I l I I . Hole Diameter (um) l b 0 CI ' 91 ~ ‘4’: >->->« ' o b 200 300 400 500 600 ( ) Time (seconds) 1.8 l .' . -........~_ . .. .......q_ I x l 2 I E C a . i 1 5 db"........-......-..v.......‘,..............-...........-....;.-.... ............-.....,-...i........-......_.....,..,_.... - ‘ : : . . . . . . . . . . . . x z z 1 4 .--- o.-.:-..... ...- ... ..¢ . - O . . l I I Rate of Dewetting (urn/s) dp.-.-........ ........... 1.2 100 200 300 400 500 Film Thickness (A) Figure 2.4. (a) Hole growth data for 5 k0 PS samples of various thicknesses over time. Film thicknesses are given in the legend. (b) Plot of the dewetting rates versus film thickness for the 5 RD PS films. 20 A m v 500 300 200 Hole Diameter (pm) 100 '- (b) 0 1000 2000 3000 4000 Time (seconds) 028 0.26 024 022 020 0.18 0.16 Rate of Dewetting (pm/s) 0.14 0.12 150 200 250 300 350 400 Film Thickness (A) Figure 2.5. (a) Hole growth data for 19 kD PS samples of various thicknesses over time. Film thicknesses are given in the legend. (b) Plot of the dewetting rates versus film thickness for the 19 kD PS films. 21 A N V E 300~~- .6. g a 200 ._ .. ._- Q a 3 A 135 100- c. 261A 0 329A . . :1 356A 0 I I . . b 0 5CDO 10000 15000 20000 25000 ( ) Tlrne (seconds) 020 I i : I 4. a g . 0.10 : i .. 3 ; 8 \ *6 i \ 0 : K g 0.05 ; . \ . ,,.,,..._L 0.00 r \ I l 100 150 200 250 300 350 400 Film Thickness (A) Figure 2.6. (a) Hole growth data for 75 kD PS samples of various thicknesses over time. Film thicknesses are give in the legend. (b) Plot of the dewetting rates versus film thickness for the 75 kD PS films. 22 Table 2.2. Cross model parameters for the different molecular weight PS samples used. Parameter 5 kD PS 19 kD PS 75 kD PS 0'0 5200 Pa 11 300 Pa 40 000 Pa 7 0.54 s 0.62 s 1.7 s 0.25 0.70 0.49 n 0.86 0.80 0.99 where 00 is taken as the plateau modulus, ris the relaxation time taken as 170/00 (where no is the terminal viscosity that was found from the data, see Table 2.1 ), n is a power law index found at higher shear rates (7}), which is between 0 and 1 for shear-thinning fluids and k is a dimensionless constant. Applying equation 2.5 to the rheological data (see Figure 2.3) we can determine the parameters of the Cross model which are given in Table 2.2. For non-Newtonian thin films, the diameter of the growing hole (0..) should follow the equation: Ddz001+{(1—-5’l’] (r! -1)Z (2'6) where Do is a characteristic diameter, n is the power-law index from the Cross model, i]! is a coefficient dependent on n and to is a characteristic time, called the crossover time, in which dewetting follows the linear trend and is proportional to the initial film thickness and is given by the equation: (27) 23 with 1 A = k W2 __1 fl: K-n wl'" +1 h. (2-8) I where h* is a characteristic length, defined as the absolute value of the spreading parameter divided by the plateau modulus, and h,- is the initial thickness of the film. The crossover times for the systems in order of increasing molecular weight are on the order of 1 second, 10 seconds, and nearing infinity. We would expect the crossover times to be larger than the values calculated for the two lower molecular weight systems since we observe a linear dewetting rate at times above these values. Applying equation 2.6 to the dewetting data gives reasonably good fits and realistic values of Do for the 5 RD and 19 kD PS films (~3um). For the 75 kD PS films the characteristic diameter that must be used is unrealistically high for the length scales seen in these experiments. The reason for this is due to the fact that n is very close to unity for the high molecular weight system and l// diverges to infinity at n=1. However, overall the shear-thinning model for the dewetting of polymer films is able to closely follow the experimental results much better than that of a Newtonian fluid since a film thickness effect is predicted and seen. Yet, agreement between the model and our results is not quantitative. Reasons for discrepancies, particularly for the high molecular weight system, may be due to the fact that under solvent annealing, the conditions of the films during dewetting are not precisely known. The exact film thickness during annealing is not known due to the fact that the polymer swells when the toluene 24 is absorbed into it.61 Further one may expect the composition of the film to vary with film thickness especially near the solid substrate which would allow for a viscosity gradient. The viscosity gradient would be dependent upon the thickness of the film, and therefore may explain the results in Figures 2.4-2.6 where the different dewetting rates are seen for different film thicknesses since the rate of dewetting is dependent upon the viscosity of the film. Also, it is possible that there may be a condensed layer of pure toluene on top of the film and/or a layer of toluene at the substrate which forms during the annealing process.62 Another interesting observation is that the 75 kD PS film, which would be an entangled film in pure form, does not exhibit slip on the substrate. de Gennes predicted that entangled polymers slip on smooth surfaces, such as silicon substrates,63 as has been seen by us“ 65 and others.66 Slip has been apparent in numerous thermally annealed film dewetting studies”: 27' 67 and if the polymer film were to slip on the substrate, the growth rate should follow a 12’3 power law. However, this is not seen in our solvent annealed samples, suggesting that surface frictional forces are not dominant in our system, but instead the energy is dissipated through viscous effects perhaps by an extremely thin toluene layer near the silicon wafer surface.61 However, we believe the shear-thinning is important due to the magnitude of the shear rate and the fact that the model of Saulnier et al. gives a reasonable representation of the data. Rim instability is apparent in all three molecular weight films, however it is most apparent in the 75 kD sample. The fingering instability of a rim is shown for 25 w._ _ v 'w i. J” I. . " a- C . . /'--'.. . . -fh.‘ . fl . o -" _' -.‘* If . \. . . 'I" o. - _ ‘ ~ \ n ‘ II E ' o —ij_‘4 q l... Oh Figure 2.7. Optical micrographs of the evolution of a dewetting hole for a 75 kD PS film. The fingering instabilities of the rim are apparent as the hole grows over time from 5 minutes (left) to 10 minutes (middle) to 20 minutes (right). Scale bars are 50 pm. a dewetting hole for this system in Figure 2.7. As can be seen, the fingers grow from the unstable rim and are left behind. They eventually break from the rim and decay into spherical droplets through Rayleigh instabilities, just as is seen in the final dewetting stage for all polymer films. The fingering instabilities that are seen mimic those patterns seen in Hele- Shaw flow that develop when a low-viscosity fluid is injected into a high-viscosity medium in a confined space.32 Our system is analogous to this where the polymer film is confined to the substrate and a lower viscosity fluid, in our case air saturated with toluene, is displacing a higher viscosity fluid, a polystyrene/toluene solution. The interface between the two is unstable and the saturated air intrudes into the polymer solution in the form of fingers.68 The observation that the fingers form more readily in the higher molecular weight sample can be attributed to the higher elasticity. Numerous studies have investigated viscous fingering in shear-thinning fluids in Hele-Shaw cells,6971 however, there has been much debate over the exact effect that an increased elasticity has on fingering. Our observation is consistent with Reiter21 however, who suggested that higher molecular weight samples are more susceptible to 26 fingering, due to the fact that the higher viscosity and entanglements of the molecules make the fingers more stable. 27 2.4. Conclusion An experimental observation into the dynamics and instabilities of polymer films by solvent annealing has been performed. Solvent annealing adds a new dimension to the dynamics of polymer film dewetting by the introduction of a third component which is apparent in both the rate of hole growth, as well as the rim instabilities. It has been found that the dewetting dynamics of solvent annealed polystyrene films follows those of non-slipping films for all molecular weights tested. Also, both the molecular weight and initial film thickness greatly influence the rate of dewetting where the smaller viscosity of the lower molecular weight films allows for faster dewetting. Also, the dewetting rates of the saturated films are inversely proportional to the initial film thickness. In addition, we have found that solvent annealing of polymer films increases the instability of the growing rims, and that for higher molecular weight polystyrene, these instabilities lead for finger formation, such as that seen in Hele-Shaw flows. This is attributed to the difference in the Viscosities of the liquid and vapor components in our system. 28 Chapter 3 Dewetting Inhibition of Polymer Fllms Using Nanoparticles 3.1 . Introduction Numerous technological applications, such as fuel cells, dielectric coatings and nanolithography resist layers, require the use of thin polymer films on a variety of organic and inorganic substrates." 2 As the film is made thinner and thinner, the ability to manufacture stable, uniform films lessens. Due to capillary forces, the polymer films may rupture and dewet from the substrate under certain 27' 33 which is detrimental to the effectiveness of the polymer film and conditions may render devices useless.9' 38 Several methods have been employed to stabilize thin polymer films on substrates. Some of these methods involve chemically modifying the polymer or the substrate to enhance their interaction with each other.“"13 In recent years, a novel approach to inhibit dewetting through the addition of nanoparticles has been studied. Nanoparticles such as fullerenes,“ dendrimers,72 polystyrene nanoparticles18 and other nano-sized fillers16'17'73‘75 have been shown to inhibit or eliminate dewetting in thin polymer films under high temperature annealing. Barnes 91 al.14 first showed that the addition of fullerene nanoparticles to polymer films eliminates dewetting due to a pinning of the contact line at the polymer- substrate interface. In addition, simulations have been performed which have supported the idea of nanoparticles inhibiting the dewetting of thin polymer films through not only a pinning effect, but also due to local viscosity changes in the film by the addition of the nanofillers.19 It is suggested that the size, mobility and 29 the interaction of the nanoparticle with the polymer all influence the ability to inhibit dewetting of polymer films. We have previously reported that the addition of fullerenes to a variety of polymer films can also inhibit dewetting under solvent annealing?8 Solvent annealing differs from high-temperature annealing due to the introduction of a third component that plasticizes the polymer film rather than heating above the polymer’s glass transition temperature.76 In the present study, we will extend this work as well as compare the ability of polystyrene nanoparticles to inhibit dewetting of thin polystyrene films annealed under a saturated toluene atmosphere. These nanoparticles differ from fullerenes both in size and chemical composition. The fullerenes are less than 1 nm in diameter and chemically dissimilar to linear polystyrene used to form the film, while the polystyrene nanoparticles are larger in diameter (4-5 nm) and chemically similar to the bulk film. Since the only difference between linear and nanoparticle polystyrene is their molecular morphology, this allows for minimal enthalpic interactions between the polymer and nanoparticle on the monomer-monomer length scale. Specifically, fullerenes and polystyrene nanoparticles will be added to polystyrene films at various concentrations to determine their effectiveness for dewetting inhibition. The films will be exposed to a saturated toluene atmosphere, plasticizing the film to achieve potential dewetting conditions. Neutron reflectivity experiments will also be performed to determine the location of the nanoparticles in the film, to gain insight into the cause for dewetting elimination. 30 3.2. Experimental The polystyrene (PS) standard obtained from Scientific Polymer Products, Inc. (weight average molecular weight (M) = 75.7 kD, polydispersity index (PDI) = 1.17) was used as received. Fullerenes (060 or buckyballs) were purchased from Term-USA and were used as received. Polystyrene nanoparticles (PSNP) were synthesized from a linear precursor (M. = 25 kD) and characterized as described elsewhere.” 78 The procedure to make the nanoparticle is to drip a solution of the linear precursor, which contains 20 mol% crosslinking groups, into hot solvent which activates the crosslinking process to produce a nanoparticle from a single molecule. The PSNP were rid of any silicon containing impurities before use. This was done by first digesting any silicon compounds by stirring the nanoparticles in a hydrofluoric acid (HF) bath for several hours. Toluene was then added to the bath to dissolve the nanoparticles. The mixture was then allowed to settle and separate, then the HF layer was decanted off, leaving only the organic solution. The solution was then dripped into methanol to precipitate the silicon-free nanoparticles which were dried under vacuum at 40°C until all the solvent had been removed. This procedure was performed to ensure that silicon contaminants did not alter the surface energy of the PSNP which may have led to false wettability of the nanoparticle filled films. A previous study in our lab demonstrated some presence of silicon containing compounds in PSNP when x- ray photon spectroscopy (XPS) was performed necessitating the above procedure to ensure accurate results. 31 Polystyrene/nanoparticle solutions were made using ACS reagent grade toluene. Stock solutions of both polystyrene and the nanoparticles were first made, and then mixed at the correct amounts to get the desired overall mass concentration as well as the desired mass concentration of nanoparticles relative to polystyrene (0.01 - 20%). The resulting solutions were filtered using a 0.2 pm PTFE filter before use. In addition, all solutions were sonicated immediately before use to ensure the nanoparticles were dispersed. Solutions were spin-coated onto silicon wafers to produce thin films. The silicon wafers (Wafer World Inc.) were first cleaned with an acid solution of 70:30 by volume 96% H2804 and 30% H202 for 1 hour to remove any organic contaminants. The wafers were then rinsed with Millipore water and dried under nitrogen. Solutions were spin-coated at a speed of 5000 rpm for 40 seconds. Film thickness was controlled by the overall mass concentration of the solution and ranged from 20-50 nm as determined using ellipsometry (JA Woollam M-44 ellipsometer). Films were determined to be uniform and continuous upon inspection with optical microscopy. Annealing of the films was performed by exposing them to a saturated toluene environment to observe the dewetting process. Reflective mode optical images (bright and dark field) were obtained in situ using an Edmund Industrial Optics optical microscope with a JAI camera attachment. Samples for neutron reflectivity measurements were prepared on 2 inch diameter silicon wafers that were 1 mm thick to ensure the substrate would not warp and give erroneous results. Samples were prepared in the same way as 32 described above. Deuterated polystyrene (dPS) standards were purchased from Scientific Polymer Products, Inc. and Polymer Source with weight average molecular weights of 63 kD (PDI = 1.10) and 82 kD (PDI = 1.15), respectively. A silanizing agent Sigmacote® (Sigma-Aldrich), which is a clear, colorless solution of a chlorinated organopolysiloxane ((SiCl2C4H9)20, 2.5%) in heptane (97.5%) was employed to prepare the silanzed surface for the flipped film reflectivity experiments. The silicon wafer was flooded with the solution for 1 minute and then the excess was rinsed off and the wafer was dried under nitrogen. The surface energy of silicon wafer treated with Sigmacote® was found to be 28 :I: 1 mN/m based on contact angle measurements with polar (water) and non-polar (methylene iodide) liquids. Neutron reflectivity experiments were carried out at two different facilities. The first was the POSY2 reflectometer at the Intense Pulsed Neutron Source at Argonne National Laboratory. The intensity of the instrument is 100 neutrons per pulse with wavelengths (i...) in the range of 2.5 - 16 A. The second was the SPEAR reflectometer at the Los Alamos Neutron Scattering Center at Los Alamos National Laboratory. Reflectance (R) measurements were taken at different angles of incidence (0° < e < 3°) and then spliced together by fitting the data in the partially overlapping wave vector (q) ranges. The overall q range that could be obtained from the wavelengths and angles possible was 0 — 0.25 A“, where q = 41tSiI'l9/An. The reduced reflectivity t.79 To do this, models were data was analyzed using the program Motofi approximated which consisted of a series of layers of constant concentrations with discrete parameters. In addition, layer roughness was included in the 33 modeling using values on the order of 1-10 A. The reflectance of the model was then calculated and then the model adjusted until the best fit was determined by minimizing 76" using the Levenberg-Marquardt algorithm. 34 3.3. Results and Discussion 3.3.1. Polystyrene/Fullerene Systems Dewetting experiments were performed using polystyrene films since it is a well characterized and readily available polymer. The polymer films were exposed to a saturated toluene atmosphere to induce dewetting with results for a pure PS film and a PS film with C60 added shown in Figure 3.1. The pure PS film has fully dewetted after only one hour of exposure to the toluene vapor while the film which has only a small amount, 3 wt%, Cso added shows no signs of dewetting after 3 hours of exposure. Previous studies have suggested that the location of the nanoparticles in the film may be the reason that dewetting is inhibited.” 18' ‘9' 38' 8° In order to confirm this, neutron reflectivity measurements were performed and the results 200 um Figure 3.1. a) Pure PS film after annealing with a saturated toluene atmosphere for 1 hour. b) PS with 3% 060 by mass after annealing with a saturated toluene atmosphere for 3 hours. Films are ~30 nm thick. 35 are presented in Figure 3.2. Three samples were considered; one where the 3 wt% Cso solution was spin-coated onto a wafer (normal film before annealing), the same composition film after solvent annealing (normal film after annealing) and a flipped film which will be described below. To allow for easier viewing, the data for the normal annealed film have been shifted up one decade and the data for the flipped film has been shifted up two decades for the plot of Rq4 versus q in Figure 3.2a. The reason for using this plot is that if sharp interfaces are present then the curves are approximately horizontal. Using this plot and due to the large difference in the neutron scattering length densities (SLD) of PS (1.41x10'6 AZ) and Ceo (5.73x10'6 A'Z) we are able to determine the location of the nanoparticles in the film despite the low concentration. The first sample was a film which had been spin-coated directly onto a silicon substrate and had not been exposed to toluene vapor. The data were fitted using a 2 layer model consisting of a region having high SLD near the substrate and low elsewhere as demonstrated by the solid lines in the figure. From the SLD profile in Figure 3.2b, for the normal unannealed film, one can see that the nanoparticles are located at the substrate in a ~2 nm C60 enriched layer (~ 29 vol% 060). This can be explained by the fact that the fullerenes are much less soluble in toluene than polystyrene is. The maximum solubility of fullerenes in toluene has been reported in the range of 2.90 mg/ml,81 while polystyrene at the molecular weight used in the current work is more than two orders of magnitude more soluble. Therefore, the fullerenes are thought to phase separate to the substrate during the spin-coating process forming the enriched layer. The 36 -e ' A Flipped Film Before Annealing 10 ' " 0 Normal Film After Annealing - ; Cl Normal Film Before Annealing Normal Film Before Annealing Normal Film After - - Film Before 0 50 100 150 200 250 300 <--— air h(A) substrate --—> Figure 3.2. a) Neutron reflectivity data and models for polystyrene films with 3 wt% Ceo. Normal films refer to films that were spin- coated directly onto silicon substrates while the flipped film was spin-coated onto mica and then flipped onto a silicon substrate. b) Scattering length density profiles for the models used to fit the data. Reflectivity data were shifted to allow for easier viewing. 37 second sample tested was exposed to toluene vapor for 3 hours (normal film after annealing). Again, the data were fitted using a 2 layer model and it is seen from the SLD profile in Figure 3.2b that the nanoparticles are in a C60 enriched layer at the substrate and remain there after annealing with little change. The fullerenes’ location in the film and the fact that they do not move throughout the film when exposed to the saturated toluene atmosphere can be explained by the effective interfacial potential of the system. Confirming that the Hamaker constant of the 3 layer assembly is negative will ensure that the system is stable and will not reorder to achieve a lower energy state.80 One can determine the sign of the Hamaker constant (A132) for a 3 layer assembly where component 3 is between component 1 and component 2 using the equation: A132 "' (”12 "”3 anz —n32) (3.1) For our system, n1 (air) is 1.0, ng (PS+Cso) is 1.77, and n3 (PS) is 1.59. The value for the PS+C60 layer was determined through an average developed from the reflectivity modeling. Using these values the Hamaker constant is calculated to be less than zero. Due to the negative value, this arrangement of layers will remain stable. If we were to change the order and have the nanoparticles move to the air interface upon annealing, the Hamaker constant would become positive, which would lead to a negative interfacial potential and result in an unstable film that dewets.82' ‘33 To confirm these results and to make sure that the nanoparticles were at the substrate immediately after spin-coating, a third sample was characterized. This sample was spin-coated onto freshly cleaved mica and then “flipped” onto a 38 silanized silicon wafer by placing the mica with the film facing down onto a silicon wafer and using water to separate the film from the mica thereby transferring it onto the silicon wafer. By flipping the film, one would expect that the nanoparticles would now be located at the polymer-air interface, which is what modeling of the data suggests. Modeling performed on this sample confirms that there is a 060 enriched layer at the air interface as seen in Figure 3.2b (flipped film before annealing), as expected. Therefore, the nanoparticles assemble at the polymer-substrate interface during the spin-coating process and subsequently remain there through annealing. We were unable to perform neutron reflectivity on an annealed flipped film to determine if the nanoparticles would separate to the substrate, since the film was not stable and dewetted, thus strengthening the argument that the location of the nanoparticles at the polymer-substrate interface is the cause for dewetting inhibition.” ‘8' ‘9' 38 The concentration of Cso used is important to induce wetting of PS films. Previously we found that a gel-like layer rich in fullerenes forms near the substrate shielding adverse van der Waals forces which we to produced a stable trilayer above.38 Here we discover there is a optimum concentration of fullerenes to achieve elimination of dewetting for a given system, as can be seen in Figure 3.3. In Figure 3.3a, a pure PS film that has fully dewetted is again shown. Adding a very small concentration of nanoparticles, such as 0.1 wt% shown in Figure 3.3b, retards the dewetting process. In this case we are below the optimum concentration to eliminate dewetting, which allows for holes to form 39 Figure 3.3. Optical micrographs of spin-cast films (~30 nm) of P3 with various concentrations of 060 after exposure to a saturated toluene atmosphere. a) Pure PS after 1 hour. b) PS + 0.1 wt% Cso after 5 hours. 0) PS + 3 wt% 060 after 3 hours. (I) PS + 10 wt% 060 after 1 hour. TEM micrographs of e) PS film with 3 wt% 060 and f) PS film with 10 wt% Ceo (inset shows diffraction pattern for a cluster of fullerenes). and grow, but they eventually stop growing even upon further annealing. In Figure 3.3c we are at the optimum concentration of 3 wt% for this system and therefore no dewetting is seen upon toluene vapor exposure. We estimate the optimum concentration is ~3 wt% through extensive experimentation at this film thickness where below this concentration partial dewetting occurs. Above the optimum concentration, dewetting can be seen again which is apparent in Figure 3.3d, where a PS film with 10 wt% 060 has fully dewetted after annealing. It is believed that for smaller concentrations of fullerenes, there are large gaps present between the nanoparticles at the substrate so there are too few nanoparticles to form the gel-like layer, and thUs holes are able to form in these spots. Yet they will only grow until they reach an area with a high enough concentration to form the gel-like layer as postulated by Barnes et al. for thermal annealing.“ For large fullerene concentration it is hypothesized that the fullerenes aggregate84 due to their high concentration in the solution and limited solubility81 forming larger agglomerates that can act as nucleating sites leading to dewetting. Note we sonicated all solutions just prior to spin-coating to ensure maximum dispersion. Microscopy was performed to test this hypothesis as shown in Figures 3.3e and 3.3f where TEM micrographs of PS films with 3 wt% and 10 wt% C60 are presented. There are very few agglomerates present in the PS film with 3 wt% film (fullerene concentration in solution of 0.21 mg/ml) as shown in Figure 3.3e. In addition, the agglomerates present are much less than 100 nm in size. This is not the case for the PS film with 10 wt% C60 (fullerene concentration in solution of 0.70 mg/ml) shown in Figure 3.3f. Here, the 41 agglomerates are in the micron size range, and clustered together into fractal-like networks several micrometers in size. The electron diffraction pattern shown in the inset demonstrates the crystalline order within an agglomerate. Since dewetting occurs while the polymer is plasticized with toluene vapor we measured the reflectivity while exposed to the vapor to determine the solvent concentration and if it was homogeneous. We know the film will become saturated when the polymer and solvent are eventually at thermodynamic equilibrium,""5 yet, since the film is so thin equilibrium is expected to rapidly occur. Unfortunately, because of this, it is not possible to determine the mass uptake over time in order to determine a diffusion coefficient as performed in many other swelling studies of polymers.”87 So, here, we present only equilibrium measurements. The films are expected to undergo a large degree of swelling to negatively affect the reflectivity modeling since many fringes are produced which are close together in q-space, so, a much thinner film, initially 18 nm thick, was prepared. To do this, a lower total (PS + C60) solution concentration was used during the spin-coating operation. The sample was characterized before, during and after exposure to a saturated deuterated toluene atmosphere, and the results are presented in Figures 3.4 and 3.5. The reflectivity data for the film before and after annealing are very similar (Figure 3.4a) and the SLD profiles shown in Figure 3.4b confirm that the C60 are located in a thin layer at the substrate just as seen in Figure 3.2. Note, the modeling was performed such that there is a constant 42 El BeforeAnnealing 0 After ------ Before Annealing — After 0 50 100 150 <--- air h(A) substrate ---> Figure 3.4. a) Neutron reflectivity data and models and b) scattering length density profiles for a polystyrene film with 7 wt% 060. The film was measured before (a) and after (0) annealing under a saturated toluene atmosphere. The film was ~18 nm thick. After annealed reflectivity data were shifted up a decade to allow for easier viewing. 43 (a) lnhornogeneous Distribution Distribution Rq‘ ill") 0.01 0.02 0.03 0.04 0.05 0.06 (b) - d-Toluene SLD p /10'°(A'2) Inhomogeneous PS SLD Distribution 0 200 400 600 8m <--- air h(A) substrate --—> Figure 3.5. a) Neutron reflectivity data and model and b) scatting length density profile for the same film as Figure 3.4 during exposure to a saturated toluene atmosphere. The horizontal lines correspond to the SLD of pure d-Toluene and PS. overall concentration of 060 in the film before and after annealing. Deuterated toluene (SLD = 5.67x106 A'z) was used to allow for contrast with the PS during annealing. Due to the fact that the SLDs of d-toluene and 060 are so close, and due to the large swelling of the polymer film, it is not possible to locate the nanoparticles during annealing for this system. However, since there is such a low concentration of C60 in the film, and there is a high concentration of d-toluene present, the results for the swollen film are not significantly influenced by the presence of the 060. During exposure to the d-toluene vapor, the film swells significantly, going from an initial film thickness of 18 nm to a thickness of 87 nm, as shown in Figures 3.5a and 3.5b. The reflectivity data were fitted using a model with eight layers each with a thickness of 10 nm and varying concentration, while the ninth layers (i.e. next to the air interface) thickness and concentration were allowed to vary. The SLDs for pure PS and d-toluene are shown in Figure 3.5b with an interesting observation that there is a decrease in the d-toluene concentration closer to the substrate (curve labeled inhomogeneous distribution). Previous studies that have considered the swelling of films upon exposure to solvents have also reported this result.“““ 88 The lower solvent concentration at the substrate is a result of the fact that the polymer chains are more confined at the substrate and so they are not able to swell as much as those chains at the air interface. Also, there appears to be a ~5 nm thick layer of pure d-toluene that has condensed on top of the PS film, which is possible due to the fact that the 45 surface tension of d-toluene is slightly lower than that of polystyrene, allowing for the solvent to wet the film. From the SLD data, the equilibrium concentration of the swollen film can be approximated as 80% by mass of solvent which agrees well with the actual degree of swelling (79% = [87 nm — 18 nm] / 87 nm). To confirm that there is an inhomogeneous distribution of toluene, a reflectivity model assuming a homogeneous distribution of d-toluene of 80% by mass throughout the film (SLD = 4.82x10'6 AZ) is also shown in Figures 3.5a and 3.5b, which does not represent the data. Note the concentration of d-toluene determined by reflectivity is much higher than the equilibrium mass uptake of solvent found in bulk measurements (~27 wt%), indicating that the absorption of solvent in thin films is different from that in bulk perhaps due to the increase in the mobility of the chains at the air surface. In addition, the high concentration of solvent in the film will lead to a large decrease in the viscosity of the film, which explains the ability of the pure PS film to dewet rapidly when annealed in a saturated toluene atmosphere. A final note is that merely putting a layer of fullerenes on the substrate then placing a layer of polymer on t0p does not result in a stable film. Dewetting occurs and so it appears that the delicate gel-like layer of fullerenes and polymer formed during spin-coating is necessary to inhibit dewetting. 46 3.3.2. Polystyrene/Polystyrene Nanoparticle Systems Previous work has shown that nanoparticles made of collapsed and cross- linked polystyrene chains have inhibited dewetting of thin polystyrene films under high temperature annealing.18 These nanoparticles were used in the current research to determine their ability to inhibit dewetting under solvent annealing. An optical micrograph of a PS film containing 10 wt% 25 kD PSNP which has been exposed to a saturated toluene atmosphere for 5 hours is shown in Figure 3.6. The image can be used to show that the addition of the nanoparticles inhibits dewetting under solvent annealing just as in thermal annealing. Neutron reflectivity experiments were also performed to determine the location of the nanoparticles before and after annealing, and the results are presented in Figure 3.7. In this case the nanoparticle and the polymer are chemically similar and each is fairly soluble in the solvent, which was not the case for the polystyrene/fullerene system. Therefore, we would not anticipate the Figure 3.6. PS + 10 wt% 25 kD PSNP film after annealing with a saturated toluene atmosphere for 5 hours. Inset shows film before annealing. Film is ~30 nm thick. 47 El BeforeAnnealing 0 After 0.02 0.04 OIB 0.08 0.10 (b) q A“) Before Annealing dPS SLD “9’ p /10“’(A'2) PSNP SLD PSNP 0 50 100 150 200 250 300 <--- air h (A) substrate --—> Figure 3.7. a) Neutron reflectivity data and models for two polystyrene films with 10 wt% 25 kD PSNP. One film is immediately after spin-coating (u) and the other has been exposed to a saturated toluene atmosphere for 5 hours (0). b) Scattering length density profile of the models used to fit the data. The SLD for pure dPS and PSNP are shown on the graph. For the after annealing sample, it is clearly seen from the SLD profile that the nanoparticles migrate to the substrate, as shown. 48 nanoparticle preferentially segregating to the solid substrate interface during the spin-coating process according to our hypothesis presented above. This is what we found after fitting the data shown in Figure 3.7a, for a sample before annealing, which indicates that the nanoparticles are initially uniformly dispersed throughout the film as can be seen in the scattering length density profile shown in Figure 3.7b. However, after a sample has been exposed to a saturated toluene atmosphere we see a change in the reflectivity profile of the film. The reflectivity curve after exposure to toluene vapor was fitted as a 2 layer film, with the layer near the substrate being the same thickness as the diameter of the nanoparticle, 4 nm. The scattering length density profile for the model used to fit the data indicates that the nanoparticles have migrated to a layer near the substrate, leaving primarily dPS in the top layer. The volume fraction of nanoparticles in the layer near the substrate can be calculated from the SLD for that layer and is found to be ~75. As expected, this large volume fraction of nanoparticles at the substrate results in a densely packed layer. Once again, the concentration of nanoparticles in the polymer film affects their ability to inhibit dewetting since the pure polymer film would rapidly dewet (Figure 3.1). To understand this phenomenon in more detail several films containing varying concentrations of 25 kD PSNP have been solvent annealed and optical micrographs are presented in Figure 3.8. It can be seen that smaller concentrations do not result in dewetting inhibition. However, at larger concentrations, dewetting is always inhibited, unlike the previous case with the 49 Figure 3.8. Optical micrographs of PS films with varying concentrations of 25 kD PSNP annealed under a saturated toluene atmosphere for 1 hour. a) Pure PS film. b) PS + 5% 25 kD PSNP film. c) PS + 15% 25 kD PSNP film. d) PS + 50% 25 kD PSNP film. All films are ~30 nm thick. fullerenes where dewetting occurred again once the optimum concentration was exceeded. Knowing that the nanoparticles migrate to the substrate upon annealing it is possible to determine the substrate surface coverage. The surface area coverage (6) of the nanoparticles can be calculated from the equation: 3 h =- _ ¢PSNPX77 2 2“ (3.2) 50 where h is the film thickness, a is the nanoparticle radius, ¢ps~p is the total volume fraction of nanoparticles in the film and his the packing density of spheres in two-dimensions, assumed to be 0.9 for maximum 2-dimensional packing. From Figure 3.8, we find that dewetting is inhibited once the nanoparticle concentration in the film reaches ~10%. With a volume fraction of 0.1, we find that the surface coverage of the nanoparticles is ~1, indicating there is a complete monolayer of nanoparticles at the substrate. Therefore, at concentrations large enough so that at least a monolayer of nanoparticles can form at the substrate, dewetting will be inhibited. A previous study has explained the migration of nanoparticles to the substrate in a polymer film through a gain in the system’s total entropy.80 For segregation to occur, the following relation should be met: 3 2 akBTl:£:l > kBT + 61:1] 0 o where or represents the degrees of freedom gained by releasing a polymer (3.3) molecule’s monomer unit from substrate constraints (i.e. being next to the hard substrate rather than the softer nanoparticle layer), which is of order 1, k3 is Boltzmann’s constant, T is the temperature, a is the components’ monomeric interaction energy,89 and a is a Kuhn or monomer length (~0.1-1 nm). This relation is the result of a balance between conformational entropy gained by the polymer chain moving away from the wall and sitting on the rougher nanoparticle layer (Mafia/013) with the loss of translational entropy available to the nanoparticle (~kBT) and the mixing enthalpy loss between the nanoparticle and 51 polymer molecule (£[a/oj2).89 For the system studied, the interaction energy term for the PSNP in PS is of order 0.1-1 kBT for dispersion forces. Therefore, segregation will occur for a nanoparticle radius of order 1-10 rim and dewetting can be inhibited assuming the concentration of nanoparticles is high enough for a monolayer to form. Therefore, polystyrene nanoparticles can be added to inhibit dewetting in solvent annealed PS films. The nanoparticles are entropically driven to the substrate during the annealing process to form a jammed, closely packed monolayer, just as seen in thermally annealed films.18 This layer shields the adverse van der Waals forces which lead to dewetting, thereby stabilizing the film by acting as an effective solid coating to promote wetting. 52 3.4. Conclusion We have shown that addition of nanoparticles to thin polymer films inhibits and even eliminates dewetting upon exposure to a saturated solvent atmosphere at room temperature, as has been seen before in high temperature annealing. Exposure of a pure polymer film shows solvent is rapidly absorbed lowering the viscosity allowing for dewetting under equilibrium solvent concentration conditions. However, the presence of nanoparticles in polymer films severely retards dewetting and appears to be a general phenomenon that is not limited to the type of nanoparticle used, as we have shown through the use of two different nanoparticles, one that is chemically dissimilar to the polymer matrix (fullerenes) and one that is the chemically the same (PSNP). We hypothesize that the fullerenes form a ~2 nm thick gel-like layer next to the solid substrate that shields the unfavorable substrate-polymer interactions thereby eliminating dewetting. The gel-like layer apparently forms during the spin-coating process due to phase separation of the fullerenes to the substrate. The concentration of fullerenes used is important to inhibit dewetting, as there is a optimum concentration where both above and below dewetting will occur. Additionally, we have shown that using PSNPs in PS films can also inhibit dewetting upon solvent annealing, just as has been seen upon high-temperature annealing. In this case, the nanoparticles are initially homogeneously distributed in the film than entropically driven to the substrate during the annealing process to form a jammed monolayer which shields the‘adverse van der Waals forces which lead to dewetting. 53 Chapter 4 Influence of Substrate Surface Energy on the Dewetting of Polymer/Nanoparticle Films 4.1 . Introduction Manufacturing uniform, thin polymer films which remain stable on a variety of solid substrates has become of technological importance. Numerous applications employ these films, including dielectric coatings, nanolithography resist layers and chemical and biological sensors." 2 Difficulties arise, however, due to the fact that in many cases molecularly thin films can easily become unstable, resulting in dewetting of the film. Whether a liquid will wet and remain stable on a solid substrate is dependent upon the spreading coefficient which is a thermodynamic parameter defined as:90 5:73—7L-75L (4.1) where ys and n are the solid and liquid surface free energies, respectively, and 731. is the interfacial free energy between the two. If the spreading coefficient is positive for a given system, the liquid will spread over the substrate, and under equilibrium conditions, remain stable. A negative value for the spreading coefficient results in a nonwetting film. From this, one can see that a substrate with a lower surface energy is more likely to have a negative spreading coefficient, and thus more likely for a film to dewet from it. 54 A number of recent studies have shown that the addition of nanoparticles to polymer films can retard and even eliminate dewetting. Nanoparticles such as fullerenes,14 dendrimers,72 polystyrene nanoparticles18 and even quantum dots"0 have been shown to inhibit dewetting upon thermal annealing. Fullerenes and polystyrene nanoparticles have also been shown to inhibit dewetting of polymer films upon solvent annealing.” 9‘ The ability of the nanoparticles to inhibit dewetting is dependent upon their location in the film as well as their interaction with the polymer, which has been demonstrated through molecular dynamics simulations.19 In the current work, we investigate the ability of nanoparticles to inhibit dewetting of polystyrene films on hydrophilic and hydrophobic substrates. In addition, the films were annealed under a saturated solvent atmosphere. Two different nanoparticles were used, with the first being polystyrene nanoparticles, so there were minimum enthalpic interactions between the polymer and the nanoparticles at the monomer level, leading to enhanced entropic interactions. Also, by altering the molecular weight of the nanoparticle, it was possible to alter its size, giving nanoparticles which range in size from ~5-8 nm. The second type of nanoparticle was fullerenes. These nanoparticles are chemically dissimilar to the polystyrene film, allowing for possible phase separation, as well as being much smaller in size (~ 0.7 nm) than the polystyrene nanoparticles. 55 4.2. Experimental The polystyrene standard obtained from Scientific Polymer Products, Inc (weight average molecular weight (M) = 75.7 kD, polydispersity index (PDI) = 1.17) was used as received. Fullerenes (060 or buckyballs) were purchased from Term-USA and were used as received. Polystyrene nanoparticles (PSNP) were synthesized from linear precursors (M... = 25, 60 and 140 kD).77 The procedure to make the nanoparticle is to drip a solution of the linear precursor, which contains 20 mol% crosslinking groups, into hot solvent to activate the crosslinking process and produce a nanoparticle from a single molecule. By varying the molecular weight, and thus the size of the chain of the linear precursor, we are able to obtain different size nanoparticles. The nanoparticles were rid of any silicon containing impurities using a hydrofluoric acid wash. This was done by first digesting any silicon compound by stirring the nanoparticles in a hydrofluoric acid (HF) bath for several hours. Toluene was then added to the bath to dissolve the nanoparticles. The mixture was then allowed to settle and separate, then the HF layer was decanted off, leaving only the organic solution. The solution was then dripped into methanol to precipitate the silicon-free nanoparticles which were dried under vacuum at 40°C until all the solvent had been removed. As discussed above the procedure was done to ensure that silicon contaminants did not alter the surface energy of the PSNP which may have led to false wettability of the nanoparticle filled films. A previous study in our lab demonstrated some presence of silicon containing compounds in PSNP samples when x-ray photon 56 spectroscopy was performed necessitating the above procedure to ensure accurate results. Polystyrene/nanoparticle solutions were made using ACS reagent grade toluene. Stock solutions of both polystyrene and the nanoparticles were first made, and then mixed at the correct amount to get the desired overall mass concentration as well as the desired mass concentration of nanoparticles to polystyrene (0.01-20%). The resulting solutions were filtered using a 0.2 pm PTFE filter before use. In addition, all solutions were sonicated immediately before use to ensure the nanoparticles were dispersed. Hydrophilic and hydrophobic silicon wafers were used to support the polymer films. The hydrophilic silicon substrates were cleaned with an acid solution of 70:30 by volume 96% H280. and 30% H202 for 1 hour to remove any organic contaminants. The wafers were then rinsed with copious amounts of Millipore water and dried under nitrogen. For these substrates, the solution was spin-coated directly onto the wafer to produce the polymer film for subsequent study. A silanizing agent Sigmacote® (Sigma-Aldrich), which is a clear, colorless solution of a chlorinated organopolysiloxane ((SiCI2C4H9)20, 2.5%) in heptane (97.5%) was employed to prepare the hydrophobic surfaces. The silicon wafer was flooded with the solution for 1 minute and then the excess was rinsed off and the wafer was dried under nitrogen. The surface energy of silicon wafers treated with Sigmacote® was found to be 28 s: 1 mN/m based on contact angle measurements with polar (water) and non-polar (methylene iodide) liquids while 57 the surface energy of a bare wafer was 71 :l: 1 mN/m. All wafers were used within one day of treatment. Spin-coating directly onto the hydrophobic substrates was not possible due to their poor wettability. Instead, solutions were spin-coated onto freshly cleaved mica sheets (Ted Pella, Inc.). The films were then transferred to the hydrophobic substrates by one of two methods. The first was by floating the film onto the surface of clean, deionized water and picking the film up with the silanized wafer. The second method involved “flipping” the film onto the silanized wafer. The mica sheet was first placed with the film side down onto the top of the silanized wafer. Water was then dispensed into the gap, displacing the film and causing it to fall to the hydrophobic substrate. All films were dried under nitrogen to remove any water left behind by the transfer process. Film thicknesses were determined to be 20-50 nm thick by ellipsometry. Films were determined to be uniform and continuous upon inspection with optical microscopy. Annealing of the films was performed by exposing them to a saturated toluene environment to observe the dewetting process. Reflective mode optical images (bright and dark field) were obtained in situ using an Edmund Industrial Optics optical microscope with a JAI camera attachment. Samples for neutron reflectivity measurements were prepared on 2 inch diameter silicon wafers that were 1 mm thick to ensure the substrate would not warp and give erroneous results. Samples were prepared in the same way as described above for samples used in the dewetting experiments. Neutron reflectivity experiments were carried out at two different facilities. The first was 58 the POSY2 reflectometer at the Intense Pulsed Neutron Source at Argonne National Laboratory. The intensity of the instrument is 100 neutrons per pulse with wavelengths (An) in the range of 2.5 - 16 A. The second was the SPEAR reflectometer at the Los Alamos Neutron Scattering Center at Los Alamos National Laboratory. Measurements of the reflectance (R) were taken at different angles of incidence (0° < 0 < 3°) and then spliced together by fitting the data in the partially overlapping wave vector (q) ranges. The overall q range that could be obtained from the wavelengths and angles possible was 0 - 0.25 A'1 where q = 41tsin0/itn. The reduced reflectivity data was analyzed using the program Motofit.79 To do this, models were approximated which consisted of a series of layers of constant concentration with discrete parameters. In addition, layer roughness was included in the modeling using values on the order of 1-10 A. The reflectivity of the model was then calculated and then the model adjusted until the best fit was determined by minimizing 78 using the Levenberg-Marquardt algorithm. 59 4.3. Results and Discussion 4.3.1 . Polystyrene/Polystyrene Nanoparticle Systems We have reported in Chapter 3 that 25 kD PSNP (radius (a) = 2.1 nm) can inhibit dewetting of thin PS films on a high energy substrate (piranha cleaned silicon wafer) under solvent annealing.91 This inhibition occurs due to the fact that the nanoparticles, which are initially uniformly dispersed throughout the film, migrate to the polymer-substrate interface creating a jammed layer.18 If the concentration of the nanoparticles is large enough, to ensure that at least a tightly packed monolayer will form at the substrate, dewetting is inhibited. The nanoparticle migration is caused by an overall system entropy gain.80 A polymer chain next to the substrate loses many conformations and so suffers a large entropy loss. By moving a nanoparticle to the substrate a translational entropy loss of only ~kBT occurs, where k3 is Boltzmann’s constant and T, temperature, which is much less than the entropy loss endured by each monomer unit in the polymer chain near the hard substrate. This is actually quite a strong driving force that scales with the nanoparticle size which is tempered a little by the loss of mixing enthalpy between the polymer and nanoparticle at the monomer level.89 Regardless we have found that polystyrene nanoparticles strongly segregate to ‘8' 3° which we will also demonstrate the solid substrate during thermal annealing occUrs during solvent annealing below. The surface area coverage (0 of the segregated nanoparticles can be calculated through a simple mass balance from the equation: 60 3 h 6 = E£ZJ¢PSNP X 77 (4.2) where h is the film thickness, ¢ps~p is the total volume fraction of nanoparticles in the film and 77 is the packing denSity of spheres in two-dimensions, assumed to be 0.9 for maximum 2-dimensional packing. Using equation 4.2, we can calculate that the volume fractions needed to form a monolayer for larger nanoparticles as discussed below. Optical micrographs of PS + PSNP films on silicon wafers, which have been annealed under a saturated toluene atmosphere, are shown in Figure 4.1. The concentrations used in these cases are 17% for the 60 kD PSNP (a = 2.8 nm, 8: 1.2) and 23% for the 140 kD PSNP (a = 3.8 nm, 8: 1.2). It is clear from the figure that the larger nanoparticles inhibit dewetting on a high energy substrate just as the 25 kD PSNP has been shown to do in our previous work.91 The nanoparticles migrate to the substrate upon annealing, where they form a jammed layer which shields the adverse van der Waals forces (Figure 4.1c).‘8' 3° For a low energy substrate, the adverse van der Waals forces emanating from the substrate are much stronger, and therefore dewetting is anticipated to occur much faster than with an unstabilized film. In other words, the spreading coefficient is highly negative (see equation 4.1) to result in a much faster dewetting rate.27 An image of a pure PS film which was spin-coated onto freshly cleaved mica and then flipped onto a wafer treated with Sigmacote® is presented in Figure 4.2. Complete dewetting of PS on a high energy, silicon wafer substrate occurs on the order of an hour of exposure to a saturated toluene atmosphere. 61 Figure 4.1. Optical micrographs of PS + PSNP films (~30 nm thick) on piranha cleaned wafers annealed for 5 hours under a saturated toluene atmosphere. a) P3 + 17% 60 kD PSNP (0:1 .2). b) PS + 23% 140 kD PSNP (9:1 .2). c) A schematic of how the monolayer of PSNP stabilizes the film against the adverse van der Waals forces. The arrows represent the extent of the van der Waals force interaction. 62 Figure 4.2. Optical micrographs of a pure PS film (~30 nm thick) on a silanized wafer after annealing under a saturated toluene atmosphere for 15 minutes. Inset shows film prior to annealing. From Figure 4.2, one can see that complete dewetting is achieved much more rapidly for the lower energy substrate. After only 15 minutes, the once uniform film has dewetted into an array of polymer droplets. The ability of the PSNP to create a stable film on the lower energy substrate was investigated. Images of annealed PS + 25 kD PSNP films are shown in Figure 4.3 where we find that the concentration of nanoparticles is crucial to the ability of the films to remain stable. From Figure 4.3, one can see that although only a monolayer of nanoparticles is needed to inhibit dewetting on the high energy substrates, more than a monolayer is needed to inhibit dewetting for the low energy substrate. In Figure 4.3b, the concentration of nanoparticles would result in a monolayer of nanoparticles at the substrate once they migrate. 63 Figure 4.3. Optical micrographs of PS + 25 kD PSNP films on a silanized wafer after annealing under a saturated toluene atmosphere. a) PS + 5% 25 kD PSNP after annealing for 15 minutes (0:05). b) PS + 10% 25 kD PSNP after annealing for 15 minutes (6:1 .0.) 0) PS + 15% 25 kD PSNP after annealing for 5 hours (6:1 4.) d) PS + 26% 25 kD PSNP after annealing for 5 hours (0:2. 5). All films are ~30 nm thick. However, dewetting occurs on the same time frame as that of pure PS. Increasing the concentration slightly retards dewetting, however holes still form, as seen in Figure 4.3c. If the concentration is raised high enough so that 2 layers are present at the substrate, as in Figure 4.3d, inhibition of dewetting is achieved. The stronger forces present from the lower energy substrate require a larger concentration and hence a thicker layer of nanoparticles to inhibit dewetting. Larger nanoparticles were then investigated to determine if they would inhibit dewetting on low energy substrates. Images of PS films with larger PSNP are presented in Figures 4.4 and 4.5. In Figure 4.4, one can see that for the 60 kD PSNP filled film, below a monolayer concentration (8: 0.9) dewetting is seen (Figure 4.4a), as expected. At a concentration slightly above that needed to form a monolayer (8: 1.4), dewetting is inhibited as shown in Figure 4.4b. The same is true for the largest nanoparticles used, 140 kD. Below a monolayer concentration (8 = 0.9) dewetting is observed (Figure 4.5a). Once the concentration reaches that needed to form a monolayer (8 = 1.2), dewetting is inhibited on the silanized wafer, as seen in Figure 4.5b. Figure 4.4. Optical micrographs of PS + 60 kD PSNP films on silanized wafers annealed under a saturated toluene atmosphere. a) PS + 12% 60 kD PSNP after annealing 30 minutes (0:0.9). 0) PS + 20% 140 kD PSNP after annealing 1 day (0:1 .4). Films are ~30 nm thick. 65 Figure 4.5. Optical micrographs of PS + 140 kD PSNP films on silanized wafers annealed under a saturated toluene atmosphere. a) PS + 16% 60 kD PSNP after annealing 30 minutes (0:09). b) PS + 23% 140 kD PSNP after annealing 20 hours (0:1 .2). Films are ~30 nm thick. The stability of polymer films on silicon substrates is dependent upon the interplay of short- and long- range van der Waals forces acting across the film. Seemann, et al. first investigated these interactions through the effective interfacial potential of the system.” 83 They found that increasing the size of the oxide layer of the silicon wafer can alter the stability of the film. Krishnan, et al. furthered this work by including terms to incorporate an alkane chain brush as well as a nanoparticle layer in the film, leading to a five layer van der Waals theory for the interfacial potential within the non-retarded approximation.92 They found that increasing the size of the alkane chain at the substrate enhances dewetting, which is similar to what we observe through the addition of a silanized layer (Sigmacote®) on our silicon wafers. In addition, they found that increasing the size of the nanoparticle layer leads to stabilization of the film, consistent with our results. We have found that the layer thickness needed to shield the van der Waals forces and stabilize the film against dewetting on a Sigmacote® treated 66 wafer is on the order of 5 nm, the size of the 60 kD PSNP or a double layer of 25 kD PSNP. Therefore, PSNPs are able to inhibit dewetting on both low and high energy substrates under solvent annealing. This inhibition is dependent upon creating a layer thick enough to shield the adverse van der Waals forces when the nanoparticles migrate to the polymer-substrate interface. In the case of the high energy substrate, these forces are not as strong, so even the smallest PSNP can inhibit dewetting assuming a full monolayer is present to create a jammed state. This jammed layer acts as a buffer at the substrate, shielding the adverse forces and immobilizing the film, eliminating dewetting from occurring. However, for the lower surface energy, the forces are much stronger, therefore, either a second layer of nanoparticles is needed (25 kD PSNP) or a single layer of larger nanoparticles (60 or 140 kD PSNP) are required to shield or retard the forces. 67 4.3.2. Polystyrene/Fullerene Systems We have previously reported that the addition of 060 nanoparticles to thin polystyrene films can inhibit dewetting upon solvent annealing.” 9‘ The ability of the fullerenes to inhibit dewetting is dependent upon the formation of a fullerene- enriched layer at the substrate which shields the adverse surface forces emanating from the substrate, as well as increasing the local viscosity prohibiting the film from moving. Simply putting a layer of fullerenes on the substrate and then placing a pure PS film on top does not inhibit dewetting, therefore this is not merely a surface energy effect. Fullerenes were added to PS films in order to determine their ability to promote wetting on high energy versus low energy substrates. Optical micrographs of PS with 3 wt% 030 films are presented in Figure 4.6. It was not possible to produce a uniform film on the low energy substrate by directly spin- coating the polymer solution. Therefore, these films were spin-coated onto Figure 4.6. Optical micrographs of PS + 3 wt% 060 films (~30 nm thick) spin-coated onto mica and floated onto a) a piranha cleaned wafer and annealed for 5 hours and b) a silanized wafer and annealed for 15 minutes. Annealing was performed by exposing the films to a saturated toluene atmosphere. Insets show films prior to annealing. 68 freshly cleaved mica, floated onto filtered water and then picked up with the substrate. This was done for both the low and high energy substrates to ensure that there would be no difference in the films. We have previously shown through neutron reflectivity experiments that the fullerenes separate to the polymer- substrate interface upon spin-coating. Therefore, there is a fullerene-enriched layer ~ 2 nm thick at the substrate containing ~30 vol% 060 fullerenes.38 For the high energy substrate (piranha cleaned silicon wafer) shown in Figure 4.6a, it is clear that dewetting is inhibited as seen before for films spin-coated directly onto a wafer. Therefore, the manufacturing process of spin-coating onto mica and transferring the film onto a wafer does not alter the ability of the nanoparticles to inhibit dewetting. However, a different result is seen in Figure 4.6b. Here, the fullerenes do not inhibit dewetting on the low energy (Sigmacote®) treated silicon wafer. Due to the fact that the surface energy is much lower in this case, the adverse forces initiating dewetting are much stronger, as discussed earlier and quantified by Krishnan, et al.92 Because these forces are much stronger in the case of the Sigmacote® substrate, the 2 nm fullerene enriched layer does not shield them promoting dewetting. This is expected due to the fact that the 25 kD PSNP, which are larger than 2 nm, were not able to shield these forces at a monolayer coverage. An alternative method to manufacture the films was tried to see if dewetting could be eliminated. Spin-coating of the polymer solution was done on mica and then the film was transferred to the wafer by flipping it over. Our 69 Figure 4.7. Optical micrographs of PS + 3 wt% 060 films (~30 nm thick) spin-coated onto mica and flipped onto a) a piranha cleaned wafer and annealed for 1 hour and b) a silanized wafer and annealed for 15 minutes. Annealing was performed by exposing the films to a saturated toluene atmosphere. Insets show films prior to annealing. previous work has shown that by flipping the film onto the silanized wafer, the fullerene-enriched layer is located at the air interface prior to annealing.91 Optical micrographs of PS with 3 wt% 060 films flipped onto the substrate are presented in Figure 4.7. In this case, the films on both the high energy and low energy substrates dewet upon exposure to a saturated toluene atmosphere. It was expected that the nanoparticles would migrate to the substrate to inhibit dewetting as has been seen with the polystyrene nanoparticles,91 however dewetting still occurred, even in the case of the high energy substrate. In order to determine the reasoning behind dewetting for the nanoparticle polymer films where the fullerenes were initially located at the air interface, neutron reflectivity experiments were performed on a variety of substrates. Films composed of linear polystyrene, 060 and PSNP were spin-coated onto mica and 70 flipped onto silanized wafers. The low energy substrates were used to produce a smoother film to decrease error and noise in the reflectivity data. A low energy substrate results in a smoother film due to the fact that when the film is transferred to the substrate from mica with water, the water will be preferentially attracted to the high energy or hydrophilic surface, mica, displacing the film evenly. If a high energy substrate had been used, the water would not be preferentially attracted to one surface, resulting in an uneven transfer of the film. The PSNP were added to ensure that the film would remain stable on the silanized water as discussed above. Two different PSNP were used to eliminate dewetting, 60 kD and 140 kD, chosen because we have previously shown that they both are large enough to shield the adverse van der Waals forces and eliminate dewetting on the low energy substrate for a monolayer coverage, the condition we use here. Due to the fact that the background film is composed of protonated PS, it is not possible to distinguish between the bulk film and the PSNP due to the fact that they have the same scattering length density (SLD) of 1.41x106 A'2 to the neutrons. However, it is possible to determine the location of the fullerenes in the film because there is sufficient contrast in the SLD of 060 (5.73x10'6 A'2) compared to protonated PS. In both cases, similar results were found and are presented in Figures 4.8 and 4.9, with the data and model fits in Figures 4.8a and 4.9a and the scattering length density profiles in Figures 4.8b and 4.90. Before annealing, the fullerenes are located in an enriched layer at the air interface, as seen before, since the film was flipped.91 This can be seen by 71 (a) 10'8 1:1 PS+60 kD PSNP+3% 060 before annealing o PS+60 kD PSNP+3% C60 after annealing Rq4 (if) orb 002 004 _1 0 0.08 0.10 (b) q (A I 3.0 : NA 1 ' 60 1 ii , L—r so 1.5 _-_., 1" t . E 1’ i 1.0 -- -- ~ ~- 0-5 " — PS+60 kD PSNP+3% 06,, before annealing ‘ ---- PS+60 kD PSNP+3% C60 after annealing 0.0 ' " : . . . . 0 50 1CD 150 200 250 <--- air h (A) substrate -—-> Figure 4.8. a) Neutron reflectivity data and models for a PS + 60 kD PSNP + 3 wt% Ceo flipped film before and after annealing under at saturated toluene atmosphere for 2.5 hours. b) The scattering length density profiles for the models used in a). The film was formulated to ensure that the PSNP are at sufficient concentration to form a monolayer. U PS+14OK PSNP+3% 060 before annealing o PS+140K PSNP+3% C60 after annealing 0.02 0.04 A4 0.06 0.08 0.10 (b) GI I . 3.0 _ NA I 60] . =5, I , «92 1.5 ‘, . E I 1.0 0-5 ""' — PS+140K PSNP+3% 06,, before annealing ' ---- PS+140K PSNP+3% 060 before annealing 0'0 _ .. . . . 0 100 2(1) 300 <--- air h(A) substrate -—-> Figure 4.9. a) Neutron reflectivity data and models for a P8 + 140 kD PSNP + 3 wt% 060 film before and after annealing under a saturated toluene atmosphere for 5 hours; b) The scattering length density profiles for the models used in a). The film was formulated to ensure that the PSNP are at sufficient concentration to form a monolayer. 73 the large SLD at the air interface followed by a homogeneous SLD profile within the film. However, after annealing an interesting observation is made. There is a layer of pure polystyrene located at the substrate interface for both cases, with a thickness approximately the same as the diameter of the PSNP used (5.5 nm for 60 kD and 7.5 nm for 140 kD). This is suggested from the SLD profile by the dip in the SLD next to the substrate to a value equal to that for pure protonated polystyrene. Using other SLD profiles did not represent the data as well which we confirmed through extensive modeling. It is assumed that these layers are the polystyrene nanoparticles which have migrated to the substrate to inhibit dewetting and have excluded the Can fullerenes. The fullerenes, on the other hand, move away from the air interface, but they do not diffuse down to the substrate during the annealing time as anticipated. Instead, a uniform dispersion of the nanoparticles in the film is seen located above the PSNP layers. Thus, the fullerenes are soluble in the polystyrene matrix at this concentration and do not preferentially migrate to the substrate most likely by a combination of enthalpic and entropic forces. Entropically, there is not much gain for the polymer to push the fullerenes to the substrate since they are so small and almost equivalent to the monomer size. Enthalpically, there may be enhanced mixing energy between the polystyrene monomer units with the fullerene surface since we have found it is possible to disperse ~2 vol% fullerenes (~3.5 wt%) in polystyrene.89 The surface energy of a pure fullerene layer is quite high and so during the PSNP stabilized annealing period they are driven from the air interface and 74 mix with the PS presumably by the arguments given above. We find it interesting though that the spin-coated layer of fullerenes ~2 nm thick on the silicon wafer interface remains stable during annealing and in fact eliminates dewetting. The spin-coating process must create a rather delicate layer containing ~30 vol% fullerenes38 that is stable, or metastable and trapped in this configuration, to produce a macroscopically stable system. 75 4.4. Conclusion In this work we have shown that dewetting of thin polymer films upon solvent annealing can be inhibited using nanoparticles. PSNP can inhibit dewetting in PS films on both hydrophilic and hydrophobic substrates. In the case of hydrophilic substrates, a monolayer of nanoparticles must be present at the substrate to form a jammed state, shielding the adverse van der Waals forces. On a lower energy substrate, these forces are stronger. Therefore, either a higher concentration of smaller nanoparticles are required (25 kD PSNP) or a monolayer of larger nanoparticles are needed (60 or 140 kD PSNP). Fullerenes, which have been shown to inhibit dewetting on a hydrophilic substrate in a previous study,91 are unable to inhibit dewetting on the lower energy substrates. The ~2 nm fullerene enriched layer that forms at the substrate upon spin-coating is not able to retard the stronger van der Waals forces present from the hydrophobic substrate. Additionally, if the film is initially flipped so that the fullerene layer is at the air interface, dewetting occurs for both the hydrophilic and hydrophobic substrates under a saturated toluene atmosphere. This is due to the fact that the fullerenes do not segregate to the polymer-substrate interface upon annealing. A unique system containing fullerenes, polystyrene nanoparticles and linear polymer was used to argue this where we found the fullerenes will homogeneously distribute through the film during annealing. Thus, during spin- coating a delicate but robust thin layer of fullerenes forms which can stabilize a polymer film under not too adverse conditions. 76 Chapter 5 Application of Technology to Chemical Sensors 5.1. Introduction In the present work, we look at the issues associated with polymer film dewetting in sensor applications. Upon exposure to saturated solvent atmospheres, polymer films are plasticized, reducing their T9 and allowing for dewetting to occur. This type of dewetting has been seen in the development of chemical sensors,9 since the sensors are tested, or “challenged” with streams of saturated solvents to test their response before use. One common type of sensor used is a surface acoustic wave (SAW) device. These devices have input and output transducers patterned on a piezoelectric substrate. When the input transducer is excited, it generates a surface acoustic wave that propagates across the substrate creating an output signal at the receiving transducer. To make the device a sensor, a polymer coating designed to selectively absorb a specific analyte is deposited onto the substrate along the propagation path. When the analyte is present, it is absorbed by the polymer coating causing a minute increase in the mass of the film that can then be measured by a shift in the SAW operating frequency relative to an uncoated reference device. When dewetting occurs in a sensor, the output frequency can become erratic or stop altogether due to the reduced analyte detection area, leading to a faulty sensor.9 The. usual approaches for stabilizing thin polymer films often involve modification of the polymer, the substrate or the interaction between the two, 77 such as crosslinking,1o sulfonation and metal complexation,11 surface treatments,12 or grafting."3 These methods, however, may impact the absorption sensitivity of the film or may not be appropriate for rapid sensor manufacture and thus cannot be used in sensor applications. Consequently, an alternate method for eliminating dewetting is needed. In this work it has been shown that the addition of nanoparticles to polymer films can inhibit dewetting under solvent annealing. The location of the nanoparticles at the polymer-substrate interface upon annealing leads to stabilization of the film. in the work presented here, this novel strategy of using nanoparticles to inhibit dewetting is adopted for use in vapor annealed polymer films used in chemical sensors. Fullerenes (060) will be added to various polymers used in sensor applications to determine their ability to eliminate dewetting when the polymer films are exposed to various solvent vapors. In addition, sensor performance will be tested with and without nanoparticles to determine their influence. 78 5.2. Experimental Polyepichlorohydrin (PECH) was purchased from Scientific Polymer Products, Inc. with an approximate value of weight average molecular weight (M) = 700 kD. Modulated differential scanning calorimetry (MDSC) was used to determine the glass transition temperature (T g) of PECH which was approximately -22.5QC (TA Instruments Q1000). DKAP polymer, a poly(dimethylsiloxane) derivative utilized in sensor applications was synthesized at and provided by Sandia National Laboratories with a reported T9 of ~ 20°C. BSP3, another polymer with a T9 below room temperature which is used in sensing applications, was also provided by Sandia National Laboratories. Fullerenes (Cso or buckyballs) were purchased from Aldrich Chemical Company and were used as received. Polymer/nanoparticle solutions were made using ACS reagent grade chloroform or toluene. Stock solutions of both polymers and the nanoparticles were first made, and then mixed at the correct amounts to get the desired overall mass concentration as well as the desired mass concentration of nanoparticles to polymer (0.01-20%). The resulting solutions were filtered using a 0.2 pm PTFE filter before use. In addition, all solutions were sonicated immediately before use to ensure the nanoparticles were well dispersed. Initial dewetting samples were manufactured by spin-coating the solutions onto silicon wafers cleaned with an acid solution of 70:30 volume ratio of 96% H2804 and 30% H202 for one hour, rinsed with Millipore water and dried with nitrogen. Spin-coating was done using a speed of 5000 rpm for 40 seconds, 79 producing films 20-50 nm thick per ellipsometry. In addition, spray-coating was utilized to prepare samples to mimic the actual manufacture of chemical sensors. This was performed using a syringe pump which fed polymer solution to a nebulizer (Burgener Research Inc.) at a flow rate of 5 thr along with a regulated stream of nitrogen at 50 psi. The resulting mist was sprayed through a 300 by 800 um shadow mask for various times to produce differing film thicknesses. In addition to coating acid cleaned silicon wafers, the polymer solutions were also sprayed onto as-received functional sensors supplied by Sandia National Laboratories to determine the influence of the coating procedure on their performance. 80 5.3. Results and Discussion 5.3.1. Dewetting Inhibition Polymers commonly used in sensor applications were studied to determine if the dewetting inhibition effect demonstrated by fullerenes upon solvent exposure was a phenomenon specific to polystyrene, or a general phenomenon independent of the polymer system. Dark field optical micrographs of the dewetting process for one of the polymers of interest in sensor applications, polyepichlorohydrin (PECH), are shown in Figure 5.1. The films were spin-coated onto silicon wafers and then exposed to a saturated methanol atmosphere which is a solvent for PECH. As seen with the pure polystyrene, the pure PECH film dewets upon exposure to the solvent atmosphere as shown in Figure 5.1a. After 45 minutes, hole formation is observed and after 20 hours of exposure to methanol vapor a distinct breakup of the film is seen. Absorption experiments using PECH and methanol were performed as described previously in Section 2.3. It is found that the amount of methanol absorbed by the PECH is - . _ .‘_-, ._ . ' L 351 om [,- . y . ‘ . Figure 5.1. Optical micrographs of spin-cast (a) pure PECH and (b) PECH + 1 wt% 060 films after exposure to a saturated methanol atmosphere for 20 hours. Insets show films prior to annealing. 81 much less than that seen in the case of toluene and PS, leading to a much higher polymer concentration (~1035 mg/ml). Because of this, the solution To is calculated to be ~ -65°C (assuming a similar T,I for methanol to that for toluene), and the viscosity is several orders of magnitude larger, explaining why the hole growth is much slower for this system.“ A PECH film with 1% by mass addition of fullerenes after exposure to methanol for 17 hours is shown in Figure 5.1b. Again, addition of the fullerenes inhibits, and at this time scale, eliminates dewetting of the polymer film. The white dots represent dust particles deposited during the annealing process exemplifying the robustness of this technology since particles frequently act as nucleation sites for dewetting." 3‘ Evaluating the influence of fullerene nanoparticles on sensor performance requires spray-coating of the polymer solutions onto the SAW substrate. Unlike spin-coating, spray-coating does not result in a uniform, smooth film, at least for the conditions used in the current procedure. In addition, we were not able to observe the coating as it is being deposited; therefore it is unclear whether a poor coating is inherent to the spraying process or results from rapid dewetting once the polymer is on the substrate. Despite this film non-uniformity the above spray-coating procedure was followed to mimic an actual procedure used to manufacture SAW sensors. Bright field optical micrographs of spray-coated PECH onto piranha-cleaned silicon wafers are shown in Figure 5.2. As demonstrated in the insets, the coatings are inhomogeneous and of varying 82 Figure 5.2. Optical micrographs of spray-coated films after exposure to a saturated methanol atmosphere. (a) Pure PECH after 42 hours and (b) PECH + 10 wt% 060 after 47 hours of exposure. Insets show films prior to annealing. thicknesses with gaps between deposited polymer on the silicon wafer. Once the pure PECH coating is exposed to a saturated methanol vapor the same dewetting effect is present with the polymer dewetting from the surface forming droplets, just as seen polystyrene films as shown in the previous chapters. In the case of PECH with 10% fullerenes by mass (Figure 5.2b), the dewetting is inhibited completely. Here, it is possible to use 10 wt% fullerenes due to the fact that the initial fullerene concentration in the solution was only 0.05 mg/ml so aggregates do not form. It is clear that gross dewetting, seen in Figure 5.2a, does not occur when nanoparticles are present. In fact, in the presence of fullerenes, it appears that exposure to the saturated solvent atmosphere has increased the overall wetted area of the polymer on the substrate. To further generalize this phenomenon another system was studied with a bright field optical micrograph of the wetting/dewetting process for the polymer DKAP shown in Figure 5.3. A solution of pure DKAP in chloroform was spray- coated onto a silicon wafer and exposed to methanol vapor for 17 hours (Figure 83 ated (a) pure DKAP and (b) DKAP + 10 wt% C50 films after exposure to a saturated methanol atmosphere for 17 hours. Insets show films prior to annealing. 5.3a), with the inset showing the coating immediately after spray-coating. It is clear that the coating has dewetted and the polymer is agglomerated for less surface coverage. However, as shown in Figure 5.3b, the coating consisting of DKAP with 10% fullerenes by mass demonstrates a different behavior. Again, this coating was spray-coated onto a silicon wafer prior to exposure to a saturated methanol atmosphere for 17 hours. No appreciable signs of dewetting are seen and the coating surface coverage appears to have remained constant, thus the presence of the nanoparticles has again inhibited dewetting. So, dewetting is prevented regardless of whether the coatings are initially smooth and uniform (spin-coating) or non-uniform (spray-coating) and it is suspected that the solid-like layer also forms during the spray-coating procedure. 84 5.3.2. Sensor Performance Live sensors that exhibit dewetting under similar conditions to those above were tested to discern the influence of nanoparticles on wetting behavior as well as sensor performance. SAW sensors used in Sandia National Laboratories’ uChemLabTM chemical sensor were used for the tests. A bright field optical micrograph of a live SAW sensor with polymer films produced by spray coating is shown in Figure 5.4. The left side of the sensor has a coating of PECH with 10% fullerenes by mass while the right side has been spray-coated with pure PECH. It is clear through observation of the figure that the coating of polymer with fullerenes has wetted the substrate to a greater degree than the coating consisting of pure polymer, which produces a very poor coating. Once again, it is unknown whether the poor film results during the coating process or from dewetting aftenNards. Nonetheless, it is clear that the presence of the nanoparticles results in a superior film. A goal of this work was to determine if the addition of nanoparticles to Figure 5.4. Optical micrograph of a live SAW sensor. The left side (Line 1) is coated with PECH + 10 wt% 060. The right side (Line 2) is coated with pure PECH. 85 polymer coatings can result in more stable, reproducible sensors. To this end, the response of live sensors having coatings with and without the addition of fullerenes was observed. The sensors were exposed to various solvents to test their response, including: dimethyl methyl phosphonate (DMMP), methyl salicylate (MS), water and methanol (MeOH), in that order. The compounds DMMP and MS are surrogates for aggressive biochemical agents. Methanol vapor represents a chemical that produces gross dewetting in most sensors and water was tested due to its ubiquitous presence. A saturated vapor stream of each solvent was sprayed across the live sensor for a period of five seconds, three times each. The response curve for a live SAW sensor is shown in Figure 5.5 with one line containing pure PECH and another line with PECH containing 10% fullerenes by mass (Figure 5.4). The curves have been shifted to allow for easier viewing with the peaks in the response curve corresponding to each challenge with solvent vapor. The response curves for the pure PECH film are shown in Figure 5.5a with curve A showing the response immediately after film deposition. The response is very erratic with an unstable baseline, leading to the belief that the polymer is dewetting from the sensor surface during the exposure. After one day, the same sensor was challenged, and as can be seen in curve B, the sensor shows no response at all and the sensor is effectively inoperable. Comparing this result to the optical micrographs shown in Figure 5.2 as well as Figure 5.4, it is clear that a non-uniform film that has dewetted from the substrate is not desirable for satisfactory sensor performance. 86 A ID v d N l l ”WE“ us EH,o EMeOHg ..L '0 .° 00 .0 ts Reference Voltage (VDC) 0 O) O N I Q I l I I l I . I - I I I I I I I I I I O 0 z ’ i . a B --_ _ 2 i I - I I I id . — : I 1 i i i i 1 00 200 300 400 Reference Time (seconds) .o o o A 0' v .a. N i i i fr —L 'o l l I I I t I . . . . . I I I D I I I I l I .-. .. .. . .. .. .. .,. .. .-. .. .. .-. .. .. ...-,. .-. ...- --. .. .. . .. .4.-.. .. .-. .. .. .-. .. "3.. .. .- I I I I l I I I I I I I . . I I . . l I I I I I I I I I I I l l O 9 II. I ' n I‘ I. DD. '0. a... II. .0 .- I P a: I I E a a a Reference Voltage (VDC) 9 o N O) I o 'o L. o 1 do 260 360 400 Reference Time (seconds) Figure 5.5. Response curves for a live SAW sensor. (a) Response to challenges for line 2 coated with pure PECH (A) immediately after spray-coating and (B) after sitting for 1 day. (b) Response to challenges for Line 1 coated with PECH with 10 wt% 060 (C) immediately after spray-coating, (D) after sitting for 3 days and (E) after spraying more solution 4 days after ' the initial deposition. 87 The line with the fullerene/PECH film was challenged and the response curves are shown in Figure 5.5b. Curve C was the response immediately after the film was sprayed. Three days later the same sensor was challenged again and the response (curve D) demonstrates that the sensitivity to DMMP has actually gotten stronger while the others have decreased in magnitude. In curve E, more fullerene/PECH solution was sprayed onto the line four days after the initial deposition, to determine if a stronger sensor response occurred, and the device was challenged again clearly revealing a stable response with increased sensitivity to DMMP. 88 5.4. Conclusion The ability to inhibit dewetting of polymer films with the addition of nanoparticles has been found to be possible with the use of a variety of different polymers. Therefore, this appears to be a general phenomenon which would allow for the possibility of use on many types of sensor polymers. Thus, nanoparticles, in particular fullerenes, can greatly increase the robustness of SAW and other sensors that require thin polymer films due to their greater stability. By using this technique, it is possible to manufacture more robust, reliable sensors. Note that the films are extremely thin, requiring only a minute amount of nanoparticles, and so the overall cost of the particles is not prohibitive to their utilization. 89 Chapter 6 Summary and Conclusions In this work we have shown that when a polymer is exposed to a saturated solvent atmosphere, the film is plasticized, lowering both its viscosity and glass transition temperature, potentially leading to dewetting. When dewetting occurs, holes first rupture and then grow in size until they coalesce, eventually forming a pattern of droplets through Rayleigh instabilities. The time scale of dewetting for polystyrene films annealed under a saturated toluene atmosphere has been shown to be dependent upon both the molecular weight of the films and the initial film thickness, different from what has been seen in thermal annealing. Here, the films exhibit shear-thinning behavior due to the large strain rate they experience as they move across the substrate. Dewetting of the polymer films can be very detrimental due to the fact that many applications require the films to remain stable over time. It has been shown in this work that the addition of nanoparticles to polymer films can eliminate dewetting. Fullerene nanoparticles eliminate dewetting of polystyrene films on silicon wafers upon exposure to a saturated toluene atmosphere due to presence of a fullerene enriched layer approximately 2 nm in thickness at the polymer-substrate interface, as determined by neutron reflectivity. The enriched layer forms during the spin-coating process due to phase separation of the fullerenes because of their poorer solubility. The fullerenes remain in this layer even after the film is plasticized by the solvent atmosphere due to the favorable interfacial interactions between the fullerenes 9O and the substrate. It is speculated that the polymer chains are not wholly confined in this layer, as its thickness is less than the radius of gyration of a polystyrene chain (~7 nm). Therefore, the fullerenes are dispersed in segments of polymer chains that also extend into the pure polystyrene layer above. There is an optimum concentration of fullerenes which must be used in order to eliminate dewetting. Below this concentration, holes will develop in the film and grow, although complete dewetting may not occur. Above the optimal concentration, the fullerenes will aggregate forming micron sized particles which can act as nucleating sites and induce dewetting. It has also been shown in the current work that polystyrene nanoparticles can also inhibit dewetting of thin polystyrene films. In this case, there are minimum monomeric enthalpic interactions between the nanoparticles and the polymer to represent a more ideal system. In addition these nanoparticles are more soluble in the solvent than fullerenes, therefore phase separation does not occur during the spin-coating process. Reflectivity data has verified that the polystyrene nanoparticles are initially homogeneously dispersed throughout the film. Upon annealing, however, the reflectivity data shows that the nanoparticles migrate to the polymer-substrate interface and inhibit dewetting of the film provided there is a large enough concentration to form a jammed layer at the substrate. The segregation of nanoparticles can be attributed to a gain in entropy due to the nanoparticles located at the substrate as opposed to the linear polymer chain. Also, through the calculation of the Hamaker constant of the film, 91 it is verified that with the nanoparticles at the substrate, the film will remain stable, as opposed to being unstable if they were to move to the air interface. The fact that the fullerenes and polystyrene nanoparticles inhibit dewetting of polystyrene films through different mechanisms can be explained due to their chemical and physical differences. Fullerenes are chemically dissimilar to polystyrene and are less soluble in toluene. Therefore, the fact that we see a phase separated layer upon spin-coating is understandable. In addition, they are found to aggregate, which at high enough concentrations can result in the formation of clusters of aggregates that are micrometers in size. Polystyrene nanoparticles, however, have essentially the same chemical makeup as the polymer chains in which they are dispersed and are more soluble in solution. Because of this, phase separation does not occur during spin-coating and the polystyrene nanoparticles do not aggregate as the fullerenes do to form colloidal scale particles, which can initiate dewetting. Therefore, there is no maximum concentration of nanoparticles that can be used to eliminate dewetting in the films. The surface energy of a substrate is important to the wetting behavior induced by the nanoparticles. Due to the much lower surface energy of a silanized water, there are stronger adverse van der Waals forces to initiate dewetting. Unfortunately, it was found in this work that the fullerene nanoparticles were not able to eliminate dewetting on the lower energy substrates. The 2 nm layer that is formed during spin-coating is not thick enough to shield the polymer film from the substrate, and so it is able to rupture and dewet. Polystyrene 92 nanoparticles are much larger however, and therefore they have been shown to inhibit dewetting on a silanized wafer. Again, the mechanism for dewetting inhibition is the nanoparticles segregating to the substrate forming a jammed layer. Finally, we have been able to apply the result found herein to the application of chemical sensors. Fullerenes were shown to be able to inhibit the dewetting of several specific polymers used for detection of harmful agents in sensor applications. When the polymers are exposed to a variety of saturated vapors, they dewet in a manner similar to that of polystyrene. However, the addition of the fullerenes to these polymers eliminated dewetting regardless of whether the films were initially smooth and uniform (spin-coated) or not (spray- coated). Additionally, tests on a live SAW sensor showed that the film with fullerenes gave a more stable response and proved more reliable than the film without nanoparticles. 93 Appendices 94 Appendix A Principles of Neutron Reflectivity With technology moving to the nano-scale, many different techniques are required to characterize materials with sub-nanometer accuracy. One such technique which is very popular is that of neutron reflectivity. Neutron reflectivity is a technique which uses neutrons to determine the depth-composition profile for thin planar samples, giving detailed information about the structure and interface of a sample.93 It is similar to the techniques of x-ray reflectivity and light ellipsometry, however it can be more useful in the characterization of polymers. The reason for this is due to the significant difference in which a neutron interacts with different isotopes of elements. The neutron analog to the difference in electron density when working with x-rays and refractive index with light is the coherent neutron scattering length (be). This value is significantly different for a deuterium atom versus a hydrogen atom, allowing for the possibility to label specific materials and produce a large contrast compared to the protonated counterpart. In addition, neutrons are able to penetrate through many engineering materials without being absorbed, which is a problem in the other techniques. Finally, neutron reflectivity is a non-destructive technique that does not require special operating conditions, making it applicable to many experiments. Previously it was stated that neutron reflectivity operates similar to reflectivity using x-rays and light. Light reflects and refracts depending upon the refractive index of the material. This is the same case for neutron reflectivity, 95 where here the neutrons reflect and refract depending upon the neutron refractive index (n,) of the material as given by the equation: 1_ lip. 2n "-= J (A-1) where xln is the neutron wavelength and ,02 is a quantity known as the scattering length density (SLD). The SLD of a material is given by the sum of the products of the atomic number density (N) and the coherent scattering length (be) for each element. When a beam of neutrons at an incident angle of 6, hits an interface between two media with neutron refractive indices n,- and n,-+1, it will either reflect at an angle equal to 6), or will refract with an angle 3+1 according to Snell’s law: n J. cos 6?]. = nj+1 cos 6’].+1 (A2) At the air interface of the sample (no=1) this equation becomes: cos «9, = r21"1 cos 60 (A.3) There is a critical angle (60) below which the angle of refraction is zero, thus cosflc=n1. The position of this critical angle provides information about the average SLD of the sample. Above this angle, reflection and refraction of the incident beam occurs and the reflectivity measured contains information about the change in SLD over the depth of the film. One can use the Fresnel reflection coefficient given by: nj Sln 6’]. - n].+1 sm l9].+1 r. = n+1 _ . - (A.4) nj sm (9]. + nj+1 sm (9}.+1 to determine the reflectivity, R, given by: 96 R = r 13141 (AS) (the actual Fresnel reflectivity is given by the product of the Fresnel reflection coefficient and its complex conjugate, but because the materials used in this study do not have complex components, it is simply the reflection coefficient squared). By imposing the boundary condition that the wave vector transfer in the direction perpendicular to the surface (qz) must be equal on both sides of the interface, we can write equation A5 in terms of Q2 in the form: - ~2 % R(q) = )6 (A5) Where qc is the critical value below which point total reflection occurs and q2 is given by: 47: . 61z = 73111 5’ (A.7) n where 0 is the angle of incidence. An interesting note is that above qc the reflectivity drops off proportional to qz" at large values of qz for sharp interfaces. Therefore, reflectivity is usually plotted as qu‘ versus qz as the data asymptotically reach a limiting value. Two different instruments were used to collect neutron reflectivity data in the work presented here. The first was the POSY2 reflectometer94 (see Figure A.1) of the Intense Pulsed Neutron Source at Argonne National Laboratory and 97 Shutters Collimator . Goniometer 3He Linear PS Figure A.1. Schematic of the POSY2 reflectometer at Argonne National Laboratory. Target Moderator Slits Detector Incoming protons f / Beam Sample Changer/ Stop Goniometer l Collimator Choppers Mirror Figure A.2. Schematic of SPEAR at Los Alamos National Laboratory. 98 the other was the Surface Profile Analysis Reflectometer (SPEAR, see Figure A2) of the Lujan Center at Los Alamos National Laboratory. Both instruments are pulsed source time-of-flight reflectometers. The neutrons are produced by the method of spallation, where protons hit a metal target, releasing neutrons from the target’s nucleus. The neutrons then pass through moderators, choppers and collimators which select only neutrons with the appropriate energy and wavelength as well as focus the beam to the desired size. These instruments use a neutron beam which has a range of wavelengths of roughly 2-14 A. To obtain a greater q range, measurements are also taken at various angles of incidence (0.4-2.5°) and the data are combined. These values allow for a possible q range of about 0.007-0.25 A“. After the neutrons hit the sample, the sum of the reflected waves (see Figure A3) is measured using a 3He linear position sensitive detector. Based upon the time to reach the detector and the position, the intensity of the sum of the reflected neutrons can be determined. Samples for neutron reflectivity were prepared on 2 inch diameter silicon wafers. The samples were either spin-coated directly onto a clean silicon wafer, or spin-coated onto a freshly cleaved mica surface and transferred to a hydrophobic silicon wafer through the use of water. The wafers used for Figure A.3. Schematic of the reflection and refraction of a beam of neutrons through a thin film (1) on a substrate (2) in air (0). 99 experiments were at least 1 mm in thickness to avoid the potential for the wafer to warp, thus ensuring that only one reflection would be seen. The films used ranged in thickness from about 15-40 nm. The POSY2 instrument required the samples to be positioned vertically, while the samples were placed horizontally in the SPEAR instrument. Due to the low flux available to the POSY2 instrument, measurement times ranged from 24-40 hours depending on the sample composition. For SPEAR, measurement times were much shorter at 2-3 hours per sample. In addition, the SPEAR instrument allowed for the possibility to measure samples in situ through the use of an aluminum chamber. The sample was placed on a platform while a reservoir below the sample was filled with solvent. The chamber was then closed allowing it to become saturated with solvent vapor. The chamber was invisible to the incident neutrons and therefore only the sample reflectivity was measured. Experiments were designed so that there is sufficient contrast between the different materials used in order to “see” nanostructures using neutrons. By using materials with different SLDs, and thus different neutron refractive indices, it is possible to determine the composition throughout the film. The different materials used in this work and their scattering length densities are listed in Table A.1. From the table one can see that there is a large difference between the PS and dPS SLDs. By simple isotopic substitution, it is possible to create a material that is chemically very similar, but can still be differentiated through the use of neutrons. Another observation is that there is also a significant difference in the SLD of 060 compared to PS. Therefore, the samples used for the experiments 100 Table A.1. Scattering length densities (SLDs) for the materials used in the reflectivity studies. Material SLD (x106) A-z PS X 1.41 dPS 6.42 060 5.73 d-Toluene 5.67 presented in this work consisted of PS with 060 nanoparticles and dPS with PS nanoparticles. The films which used a background of PS required longer measurement times due to the fact that the hydrogen atoms created a large amount of incoherent scattering. Once data are collected by the detector, they must be properly reduced. First, the reflected beam intensity is normalized to the intensity of the incident beam. This is done by measuring the direct beam in what is called a transmission run, where no sample is present and the beam hits the same detector. Transmission runs are taken every time the beam is started up or any time the detector is changed to ensure accurate results. Next, the data sets for the same sample which used different angles of incidence are combined. The data sets are merged together by scaling the reflectivities of the higher angle runs so that they match those of the lower angle runs in the overlapping regions. Only the lowest angle run is not shifted because this is where the total reflection region, Fl=1, exists. Once the data are properly reduced it can be analyzed. 101 Appendix B Neutron Reflectivity Data Interpretation Due to the highly complex math involved for even a single interface, neutron reflectivity data are interpreted through the use of models and fitting routines. The models may consist of various layers, each having its own thickness (d) and SLD (see Figure 8.1). It is also possible to give each layer a specific roughness (a) to allow for a diffuse interface. There are several programs available to the interpretation of reflectivity data, and the one used in the present work is called Motofit.79' 95 Motofit is a macro specifically designed to work within the analysis package IGOR Pro.96 It works by having the user input the known parameters in the system as well as estimated values for the unknown parameters. The known parameters are held constant while the unknowns are unconstrained. Due to the fact that there can be several solutions for the data depending on the number of parameters left unconstrained, it is assumed that one knows as much as possible 0 air no 1 I d1 n1 2 I 02 n2 ”14 t dm-t nm-1 -; \u \ \ . . .t- i 1. s -' '.\ .- ‘. ~, w \ ‘13 “(3‘3 _ 1 .' V --~—'."_ 9* ';_ ‘. a”: ‘5‘ kfibcrafi‘érst‘trm931's::l‘.'\;::t1 3:.:v'.\'«3“.\9~:122:1:-‘—::..~.;_;‘ :‘,~:.“;3,;::_:.: ‘ , : ,3 2: Qty. .41" .-lr‘.;.. (5.. r'.'..\. g ‘u 7 q. '_.' ‘_ -.;. i ,[K .--‘ ~ . ...f. -‘.‘\“,\ ‘41:... .; .\. .. .' .. 1:1: 1. 3 .3. 4 , . .ag..~.-; . .- .. . x)... ~ ' x ;.;,.:.. 5. . . . 1.. t..: 4. ,1 .7 "1+ V ‘ ‘ ' ‘ -‘ 1' ‘ ‘ , ‘ V ‘ "' ~ ‘ 'v ‘ '- ‘ ‘ r _. .\'.....‘-.,. . ... ‘.... >-.I‘,““<.-:,171IL~-‘1\u|":-’v.h :.....: “‘1 ... .1‘ .. ... ......_.. C 1 '11‘3 _. .‘ .‘_<;.. _ .‘H’.’ \‘ .,'_ . . . . 1,... .. '. ’ . ' "x.'. , ‘ . . .‘ ‘ , ' - x . . ' :.."\ 4»- ~. I. <<<<< .... ._...,.... «a ,"— .134. ._._g ..A, Q : a:.‘~.\‘}:~:-.lx~,:~:~“\.-';~:\:$._ ::.-.\2; :. v\~.::.—‘;::--- ‘ 1 —.--- .- -. . ~ 3.. Figure 8.1. Illustration of a reflectivity model for a film having m layers, each with a distinct thickness (d) and neutron refractive index (n). 102 about the sample through the use of other techniques to compliment the reflectivity data. Using a least squares method (Levenberg-Marquardt), the program is used to determine the best fit to the data. Once the data have been appropriately modeled, the SLD ‘ profile (p(z)) can be deduced and then be converted into a volume fraction profile (¢(z)) through the use of the equations: pz-p ¢1 (z) = _§_)__z ,0] "pz ¢1(2)+¢2 (z)=1 (8.2) where the subscripts 1 and 2 denote the two different materials that the film is (3.1) composed of. This enables one to know the composition of the film throughout the entire depth. The reduced data obtained from the experiments were loaded into Motofit in the form of 3 columns, q, F? and error in Fl. Once the data were loaded it was automatically displayed in a plot. The reflectivity fit panel (see Figure 8.2) was then used to enter in the model parameters desired. All data were fitted against Fl vs. q and weighted with the error in R. In this case there is some bias towards the low q region so it is important to make sure that the critical edge is fit well or it could lead to a large increase in the chi squared value. For the model, there are several constant parameters which the values are already known (see Table 8.1). These values are input into the model and locked. Next, initial values for the other parameters are entered into the model. Some of these values, such as the SLD for pure materials, are known and can therefore be locked as well. The unknown values are left unconstrained and an 103 Figure 3.2. Image of the reflectivity panel window in Motofit. initial fit to the data is made. At first, the roughness of each layer is locked at 4 A while the other unknown parameters are fit. This is due to the fact that the data isn’t as sensitive to the roughness, especially when the data does not extend to large q values. Until a good initial fit to the data has been obtained, the roughness can oscillate considerably to give an unrealistic resurt. Once there is a good initial fit to the data, the roughness of the layers can then be fit to yield a realistic value. Table 8.1. Known parameters used in all reflectivity models. Parameter Value Scale 1 SLD top 0 SLD base 2.07 Sigma_base 4 Solvent 0 penetration 104 Appendix C Additional Data The following figures and tables are examples of reflectivity data collected and the models fitted using the Motofit program. The data was collected to supplement the data presented in Chapters 3 and 4 of this work. Also included, are the parameters used to fit the reflectivity data presented in Chapters 3 and 4. 105 (a) 10'8 El 75 kD PS on piranha wafer (b) 0.02 0.04 1) 0.06 0.08 0.10 p /10'6(A'2) I— 75 kD PS on piranha wafer 50 100 150 200 250 300 <—— air 2 (A) substrate -—-> Figure 0.1. a) Neutron reflectivity data and model for a polystyrene film on a piranha cleaned silicon wafer. b) Scattering length density profile of the model used to fit the data. 106 Table C.1. Fitted values for the model parameters used in Figure C.1. Layer “$23988 p (21:6 (X) bkg x2 0 - 0 - 1 297 1.41 5.9 2.09E-06 57.15 base - 2.07 4 107 El 63 kD dPS on piranha wafer 0.02 0.0 0.08 0.10 (b) 4 _1 0.06 NA ) p/1of< '2) — 63 kD dPS on 0 50 100 150 200 <—— air 2 (A) substrate ——-> Figure C.2. a) Neutron reflectivity data and model for a deuterated polystyrene film on a piranha cleaned silicon wafer. b) Scattering length density profile of the model used to fit the data. 108 Table C.2. Fitted values for the model parameters used in FigureC.2. Layer Thltz’lzqess pa 1:6 (X) bkg x2 0 - 0 - 1 215 6.41 8. 2 2.07E-06 48.00 base - 2.07 4 109 El 82 kD dPS on piranha wafer o 82 kD dPS on wafer ‘\ 6 .- 7‘ _ I I 5 ._ 1 . .. 1 A I' N '.< 4-- . I. .. V 1° '. o \ E 3 .-. ,.. ‘. - Q I I 2 “ — 82 kD dPS on piranha wafer ---- 82 kD dPS on Sigmacote® wafer 50 100 150 200 250 <—— air 2 (A) substrate —--> Figure C.3. a) Neutron reflectivity data and models for deuterated polystyrene film. One was spin-coated onto a piranha cleaned silicon wafer while the other was floated onto a Sigmacote® treated wafer. b) Scattering length density profiles of the models used to fit the data. 110 Table C.3. Fitted vaers for the model parameters used for the piranha cleaned sample in Figure C.3. - 6 Layer “$231988 pay; (2) bkg x2 0 - 0 - 1 257 6.41 5.7 1.70E-05 22.92 base - 2.07 4 Table C.4. Fitted values for the model parameters used for the Sigmacote® treated sample in Figure C.3. Thickness p x106 0 2 La er . bk y (A) (A2) (A) 9 x 0 - 0 - 1 224 6.41 8.7 1.40E-05 1 10. 7 base 2.07 4 111 El 75 kD PS + 3% 060 after annealing Hq‘ (K4) _L l 0.02 o. 0.08 0.10 (b) 04 _1 0.06 (NA ) p /10"’(A‘2) — 75 kD PS + 3% 060 after annealing 0 100 200 300 400 <—- air 2 (A) substrate -—-> Figure C.4. a) Neutron reflectivity data and models for polystyrene with 3 wt% C60 films on piranha cleaned silicon wafers. One was tested without annealing and the other was first annealed with a saturated toluene vapor. b) Scattering length density profiles of the models used to fit the data. 112 Table C.5. Fitted values for the model parameters used in Figure C.4. Thickness p x106 0 2 La er _ bk y (A) (A2) (A) 9 x o - of - 1 387 1.41 3 4.62E-08 188.9 2 27 2.5814 3.6 base - 2.07 4 113 (a) 10'8 El 75 kD PS + 3% C60 before annealing o 75 kD PS + 3% 060 after annealing Rq“ (Aj) orb 0.02 0 04 0 0.08 0.10 -1 (b) q (A ) 3.0 g f f I | 5 \1_ A 2.0 __...,...._ ' 1 N I ; S i I 9 ‘90 1.5 ,. : 1 ...,....-.. a f ' 1.0 E 0'5 — 75 kD PS + 3% C60 before annealing ' ---- 75 kD PS + 3% C60 after annealing 0.0 , - - - 0 50 100 150 200 250 300 <—— air z(A) substrate -—-> Figure C.5. a) Neutron reflectivity data and models for polystyrene with 3 wt% C60 films on piranha cleaned silicon wafers. One was tested without annealing and the other was first annealed with a saturated toluene vapor. b) Scattering length density profiles of the models used to fit the data. 114 Table C.6. Fitted values for the model parameters used for the before annealed sample in Figure C.5. Thickness p x106 0 2 La er . bk y (A) (A 2) (A) 9 X 0 - 0 - 1 293 1.41 7.4 8.33E-O7 53.14 2 20 2.3 3.3 base - 2.07 4 Table 0.7. Fitted values for the model parameters used for the after annealed sample in Figure 0.5. Thickness p x106 6 2 La er - bk y (A) (A2) (A) 9 x O - 0 - 1 260 1.41 7.1 5.93E-07 71.28 2 23 2.4732 2.9 base - 2.07 4 115 Table 0.8. Fitted values for the model parameters used for the normal film before annealing in Figure 3.2. Thickness 0 x106 0 2 La er - bk y (A) (A 2) (A) 9 x O - O - 1 300 1.41 4 1.14E-05 393.2 2 14 2.4 4 base - 2.07 4 Table 0.9. Fitted values for the model parameters used for the normal film after annealing in Figure 3.2. Thickness p x106 0 2 La er - bk y (A) (A2) (A) 9 x O - 0 - 1 294 1.41 4 8.27E-06 97.9 2 15 2.8 4 base - 2.07 4 Table C.10. Fitted values for the model parameters used for the flipped film before annealing in Figure 3.2. Thickness p x109 0 2 La er . bk y (A) (A2) (A) 9 x 0 - O - 1 16 1.41 4 1.38E-05 35.3 2 293 2.3 4 base - 2.07 4 116 Table C.11. Fitted values for the model parameters used for the film before annealing in Figure 3.4. Thickness p x109 0 2 La er . bk y (A) (A2) (A) 9 x O - O - 1 1 58 1 .41 4 1 .54E-O6 14.7 2 1 7 3.33 4 base - 2.07 4 Table 0.12. Fitted values for the model parameters used for the film after annealing in Figure 3.4. Thickness p x106 0 2 La er - bk y (A) (A2) (A) 9 x 0 - 0 - 1 1 58 1 .41 4 1 .54E-06 6.7 2 20' 3.1 4 base - 2.07 4 117 Table C.13. Fitted values for the model parameters used for the film during annealing in Figure 3.5. ' 6 Layer ”"2239” “(212? (X) bkg x2 0 - 0 - 1 72 5.67 4 2 1 00 5.1 1 3 1 00 4.84 1 4 1 00 4.76 1 5 100 4.7 1 2.80E-06 9.5 6 1 00 4.63 1 7 1 00 4.49 1 8 1 00 4.06 1 9 1 00 3.17 1 base - 2.07 4 118 Table C.14. Fitted values for the model parameters used for the before annealed sample in Figure 3.7. p><106 Layer “322953 (A,2) (X) bkg X2 0 - O - 1 268 5.92 4 1.66E-05 73.2 base - 2.07 4 Table C.15. Fitted values for the model parameters used for the before annealed sample in Figure 3.7. Thickness p x106 0 2 La er . bk y (A) (A 2) (A) 9 x O - O - 1 248 6.41 4 1.28E-O6 54.4 2 40 2.68 4 base - 2.07 4 119 Table 0.16. Fitted values for the model parameters used for the film before annealing in Figure 4.8. Thickness p ><106 o 2 La er - bk y (A) (A 2) (A) 9 x 0 - 0 - 1 20 2.34 4 3.37E-06 14.00 2 241 1.41 4 base - 2.07 4 Table 0.17. Fitted values for the model parameters used for the film after annealing in Figure 4.8. Thickness p x106 0 2 La er . bk y (A) (A 2) (A) 9 x 0 - 0 - 1 215 1 .51 4 6.22E-06 11.15 2 55 1.41 4 base - 2.07 4 120 Table 0.18. Fitted values for the model parameters used for the flipped film before annealing in Figure 4.9. Thickness p x106 0 2 La er - bk y (A) (A2) (A) 9 x O - 0 - 1 20 2.7 4 1.25E-06 28.40 2 312 1.41 4 base - 2.07 4 Table 0.29. Fitted values for the model parameters used for the flipped film after annealing in Figure 4.9. Thickness p x109 0 2 La er _ bk y (A) (A 2) (A) 9 7‘ 0 - 0 - 1 282 1.51 4 1.06E-05 50.99 2 75 1.41 4 base - 2.07 4 121 ReferenccsReferences 1. Licari, J. J., Plastic Coatings for Electronics. McGraw-Hill: New York, 1970. 2. Kazmerski, L. L., Polycrystalline and amorphous thin films and devices. Academic Press: New York, 1980. 3. Xie, FL; Karim, A.; Douglas, J. F.; Han, 0. 0.; Weiss, R. A., Spinodal dewetting of thin polymer films. Phys. Rev. Lett. 1998, 81, (6), 1251-1254. 4. Stange, T. G.; Evans, D. F.; Hendrickson, W. A., Nucleation and growth of defects leading to dewetting of thin polymer films. Langmuir 1997, 13, (16), 4459- 4465. 5. Muller-Buschbaum, P., Dewetting and pattern formation in thin polymer films as investigated in real and reciprocal space. Journal of Physics-Condensed Matter 2003, 15, (36), R1549-R1582. 6. Reiter, G.; Sharma, A.; Casoli, A.; David, M. 0.; Khanna, R; Auroy, P., Thin film instability induced by long-range forces. Langmuir 1999, 15, (7), 2551- 2558. 7. Zhao, H. P.; Wang, Y. J.; Tsui, O. K. 0., Dewetting induced by complete versus nonretarded van der Waals forces. Langmuir 2005, 21, (13), 5817-5824. 8. Royer, J. FL; Gay, Y. J.; Desimone, J. M.; Khan, 8. A., High-pressure rheology of polystyrene melts plasticized with 002: Experimental measurement and predictive scaling relationships. Journal of Polymer Science Part B-Polymer Physics 2000, 38, (23), 3168-3180. 9. Grate, J. W.; McGill, Ft. A., Dewetting Effects on Polymer-Coated Surface- Acoustic-Wave Vapor Sensors. Analytical Chemistry 1995, 67, (21 ), 4015-4019. 10. Carre, A.; Shanahan, M. E. FL, Effect of cross-linking on the dewetting of an elastomeric surface. Joumal of Colloid and Interface Science 1997, 191, ( 1), 141-145. 122 11. Feng, Y.; Karim, A.; Weiss, R. A.; Douglas, J. F.; Han, 0. 0., Control of polystyrene film dewetting through sulfonation and metal complexation. Macromolecules 1998, 31, (2), 484-493. 12. Ashley, K. M.; Meredith, J. 0.; Amis, E.; Raghavan, D.; Karim, A., Combinatorial investigation of dewetting: polystyrene thin films on gradient hydrophilic surfaces. Polymer 2003, 44, (3), 769-772. 13. Yerushalmirozen, R.; Klein, J.; Fetters, L. J., Suppression of Rupture in Thin, Nonwetting Liquid-Films. Science 1994, 263, (5148), 793-795. 14. Barnes, K. A.; Karim, A.; Douglas, J. F.; Nakatani, A. l.; Gruell, H.; Amis, E. J., Suppression of dewetting in nanoparticle-filled polymer films. Macromolecules 2000, 33, (1 1), 4177-4185. 15. Mackay, M. E.; Hong, Y.; Hong, 8.; Russell, T. P.; Hawker, 0. J.; Vestberg, R.; Douglas, J., The influence of dendrimers on the dewetting of thin polystyrene films. Langmuir 2002, 18, 1877-1882. 16. Sharma, 8.; Rafailovich, M. H.; Peiffer, D.; Sokolov, J., Control of dewetting dynamics by adding nanoparticle fillers. Nano Letters 2001, 1, (10), 511-514. 17. Besancon, B. M.; Green, P. F., Polystyrene-based single-walled carbon nanotube nanocomposite thin films: Dynamics of structural instabilities. Macromolecules 2005, 38, (1 ), 1 10-1 15. 18. Krishnan, R. 8.; Mackay, M. E.; Hawker, C. J.; Van Horn, 3., Influence of molecular architecture on the dewetting of thin polystyrene films. Langmuir 2005, 21 , (13), 5770-5776. 19. Luo, H. B.; Gersappe, D., Dewetting dynamics of nanofilled polymer thin films. Macromolecules 2004, 37, (15), 5792-5799. 20. Jacobs, K.; Herminghaus, 8.; Mecke, K. R, Thin liquid polymer films rupture via defects. Langmuir1998, 14, (4), 965-969. 21. Reiter, G., Unstable Thin Polymer-Films - Rupture and Dewetting Processes. Langmuir1993, 9, (5), 1344-1351. 123 22. Thiele, U.; Verlarde, M. G., Dewetting: Film Rupture by Nucleation in the Spinodal Regime. Physical Review Letters 2001 , 87, (1 ), 016104. 23. Kim, H. l.; Mate, 0. M.; Hannibal, K. A.; Perry, 8. 8., How Disjoining Pressure Drives the Dewetting of a Polymer Film on a Silicon Surface. Physical Review Letters 1999, 82, (17), 3496-3499. 24. Sharma, A.; Khanna, R., Pattern Formation in Unstable Thin Liquid Films under the Influence of Antagonistic Short- and Longs-Range Forces. Journal of Chemical Physics 1 999, 1 10, (10), 4929-4936. 25. Karapanagiotis, l.; Evans, D. F.; Gerberich, W. W., Dewetting Dynamics of Thin Polystyrene Films from Sputtered Silicon and Gold Surfaces. Colloids and Surfaces A: Physiochemical and Engineering Aspects 2002, 207, 59-67. 26. Redon, 0.; Brzoska, J. B.; Brochard-Wyart, F., Dewetting and Slippage of Microscopic Polymer Films. Macromolecules 1994, 27, (2), 468-471. 27. Brochard-Wyart, F.; de Gennes, P. G.; Hervert, H.; Redon, 0., Wetting and Slippage of Polymer Melts on Semi-Ideal Surfaces. Langmuir 1994, 10, 1566-1572. 28. Jacobs, K.; Seemann, R.; Schatz, G.; Herminghaus, 8., Growth of Holes in Liquid Films with Partial Slippage. Langmuir1998, 14, 4961-4963. 29. Neto, 0.; Jacobs, K., Dynamics of Hole Growth in Dewetting Polystyrene Films. Physica A 2004, 339, 66-71. 30. Debregeas, G.; Martin, P.; Brochard-Wyart, F., Viscous Bursting of Suspended Films. Physical Review Letters 1995, 75, (21), 3886-3889. 31. Xavier, J. H.; Pu, Y.; Li, 0.; Rafailovich, M. H.; Sokolov, J., Transition of Linear to Exponential Hole Growth Modes in Thin Free-Standing Polymer Films. Macromolecules 2004, 37, 1470-1475. 32. Gu, X.; Raghavan, D.; Douglas, J. F.; Karim, A., Hole-growth instability in the dewetting of evaporating polymer solution films. Journal of Polymer Science Part B-Polymer Physics 2002, 40, (24), 2825-2832. 124 33. Reiter, G., Dewetting of Thin Polymer Films. Phys. Rev. Lett. 1892, 68, (1), 75-78. 34. Karapanagiotis, l.; Evans, D. F.; Gerberich, W. W., Nucleation processes for dewetting initiation of thin polymer films. Langmuir 2001, 17, (11), 3266-3272. 35. Masson, J. L.; Green, P. F., Hole formation in thin polymer films: A tvvo- stage process. Physical Review Letters 2002, 88, (20), 205504. 36. Voronov, A.; Shafranska, 0., Synthesis of chemically grafted polystyrene "brushes" and their influence on the dewetting in thin polystyrene films. Langmuir 2002, 18, (11), 4471-4477. 37. Voronov, A.; Shafranska, 0., Dependence of thin polystyrene films stability on the thickness of grafted polystyrene brushes. Polymer 2003, 44, (1), 277-281. 38. Holmes, M. A.; Mackay, M. E; Giunta, R. K., Nanoparticles for Dewetting Suppression of Thin Polymer Films used in Chemical Sensors. Journal of Nanoparticle Research published on-line July 2006. 39. Lee, S. H.; Yoo, P. J.; Kwon, S. J.; Lee, H. H., Solvent-driven dewetting and rim instability. Joumal of Chemical Physics 2004, 121, (9), 4346-4351. 40. Peng, J.; Xuan, Y.; Wang, H.; Li, B.; Han, Y., Solvent Vapor Induced Dewetting in Diblock Copolymer Thin Films. Polymer 2005, 46, 5767-5772. 41. Zhu, J.; Zhao, J.; Liao, Y.; Jiang, W., Multiscale Dewetting of Triblock Copolymer Thin Film Induced by Solvent Vapor. Journal of Polymer Science Part B: Polymer Physics 2005, 43, 2874-2884. 42. Redon, 0.; Brochardwyart, F.; Rondelez, F., Dynamics of Dewetting. Phys. Rev. Lett. 1991, 66, (6), 715-718. 43. Bird, R. B.; Stewart, W. E.; Lightfoot, E. N., Transport Phenomena. J. Wiley: New York, 1960. 125 44. Rauch, J.; Kohler, W., Collective and thermal diffusion in dilute, semidilute, and concentrated solutions of polystyrene in toluene. Journal of Chemical Physics 2003, 1 19, (22), 11977-11988. 45. Wiedersich, J.; Surovtsev, N. V.; Rossler, E., A comprehensive light scattering study of the glass former toluene. Journal of Chemical Physics 2000, 113, (3), 1143-1153. 46. Ober, R.; Paz, L.; Taupin, 0.; Pincus, P.; Boileau, 8., Study of the Surface-Tension of Polymer-Solutions - Theory and Experiments .1. Good Solvent Conditions. Macromolecules 1983, 16, (1 ), 50-55. 47. Gandhi, K. 8.; Williams, M. 0., Solvent Effects on Viscosity of Moderately Concentrated Polymer Solutions. Joumal of Polymer Science Part C-Polymer Symposium 1971, (35), 21 1-&. 48. McCabe, W.; Smith, J.; Harriott, P., Unit operations of chemical engineering. MoGraw-Hill: New York, 1993. 49. Brandrup, J.; lmmergut, E. H., Polymer Handbook. John Wiley & Sons: New York, 1989. 50. Fox, T. G.; Flory, P. J., Viscosity-Molecular Weight and Viscosity- Temperature Relationships for Polystyrene and Polyisobutylene. Journal of the American Chemical Society 1948, 70, (7), 2384-2395. 51. Mackay, M. E.; Dao, T. T.; Tuteja, A.; Ho, D. L.; Van Horn, B.; Kim, H. C.; Hawker, C. J., Nanoscale effects leading to non-Einstein-like decrease in viscosity. Nature Materials 2003, 2, (11), 762-766. 52. Shull, K. R.; Karis, T. E., Dewetting dynamics for large equilibrium contact angels. Langmuir1994, 10, (1), 334-339. 53. Ober, R.; Paz, L.; Taupin, 0.; Pincus, P.; Boileau, 8., Study of the surface tension of polymer solutions: theory and experiments. Good solvent conditions. Macromolecules 1983, 16, (1 ), 50-55. 54. Brown, H. R.; Russell, T. P., Entanglements at Polymer Surfaces and Interfaces. Macromolecules 1996, 29, (2), 798-800. 126 55. Keddie, J. L.; Jones, R. A. L.; Cory, R. A., Size-Dependent Depression of the Glass-Transition Temperature in Polymer-Films. Europhysics Letters 1994, 27, (1), 59-64. 56. DeMaggio, G. B.; Frieze, W. E.; Gidley, D. W.; Zhu, M.; Hristov, H. A.; Yes, A. F., Interface and Surface Effects on the Glass Transition in Thin Polystyrene Films. Physical Review Letters 1997, 78, (8), 1524. 57. Tsui, O. K. 0.; Russell, T. P.; Hawker, C. J., Effect of Interfacial Interactions on the Glass Transition of Polymer Thin Films. Macromolecules 2001, 34, (16), 5535-5539. 58. Pan, 0.; Winey, K. |.; Hu, H., H.; Composto, R. J., Unstable Polymer Bilayers. 2. The Effect of Film Thickness. Langmuir 1997, 13, 1758-1766. 59. Masson, J. L.; Green, P. F., Viscosity of Entangled Polystyrene Thin Film Melts: Film Thickness Dependence. Physical Review E 2002, 65, 031806. 60. Saulnier, F.; Raphael, E.; de Gennes, P. G., Dewetting of thin-film polymers. Physical Review E 2002, 66, (6), 061607. 61. Elbs, H.; Krausch, G., Ellipsometric determination of Flory-Huggins interaction parameters in solution. Polymer 2004, 45, (23), 7935-7942. 62. Richardson, H.; Lopez-Garcia, l.; Sferrazza, M.; Keddie, J. L., Thickness dependence of structural relaxation in spin-cast, glassy polymer thin films. Physical Review E 2004, 70, (5), 051805. 63. de Gennes, P. G., Comptes Rendus de I'Académie des Sciences 1979, 228B, 219. 64. Henson, D. J.; Mackay, M. E., Effect of Gap on the Viscosity of Monodisperse Polystyrene Melts - Slip Effects. Journal of Rheology 1995, 39, (2), 359-373. 65. Mackay, M. E.; Henson, D. J., The effect of molecular mass and temperature on the slip of polystyrene melts at low stress levels. Journal of Rheology 1998, 42, (6), 1505-1517. 127 66. Migler, K. B.; Hervet, H.; Leger, L., Slip transition of a polymer melt under shear stress. Physical Review Letters 1993, 70, (3), 287. 67. Reiter, G.; Khanna, R., Real-time determination of the slippage length in autophobic polymer dewetting. Phys. Rev. Lett. 2000, 85, (13), 2753-2756. 68. Malay, K. J.; Feder, J.; Jessang, T., Viscous Fingering Fractals in Porous Media. Physical Review Letters 1985, 55, (24), 2688. 69. Allen, E.; Boger, D. V. In The Influence of Rheological Properties on Viscous Fingering in Enhanced Oil Recovery, Proceedings of the Xth International Congress on Rheology, Sydney, Australia, 1988; Uhlherr, P. H. T., Ed. Australian Society of Rheology: Sydney, Australia, 1988; pp 146-148. 70. Sader, J. E.; Chan, D. Y. 0.; Hughes, B. D., Non-Newtonian effects on immiscible viscous fingering in a radial Hale-Shaw cell. Physical Review E 1994, 49, (1), 420. 71. Wilson, 8. D. R., The Taylor-Saffman Problem for a Non-Newtonian Liquid. Journal of Fluid Mechanics 1990, 220, 413-425. 72. Mackay, M. E.; Hong, Y.; Jeong, M.; Hong, 8.; Russell, T. P.; Hawker, 0. J.; Vestberg, B.; Douglas, J. F., Influence of dendrimer additives on the dewetting of thin polystyrene films. Langmuir 2002, 18, (5), 1877-1882. 73. Hosaka, N.; Torikai, N.; Otsuka, H.; Takahara, A., Structure and Dewetting Behavior of Polyhedral Oligomeric Silsesquioxane-Filled Polystyrene Thin Films. Langmuir 2007, 23, (2), 902-907. 74. Wei, B.; Gurr, P. A.; Genzer, J.; Qiao, G. G.; Solomon, D. H.; Spontak, R. J., Dewetting of Star NanogeVHomopolymer Blends from an lmmiscible Homopolymer Substrate. Macromolecules 2004, 37, (21), 7857-7860. 75. Xavier, J. H.; Sharma, 8.; Sec, Y. 8.; Isseroff, R.; Koga, T.; White, H.; Ulman, A.; Shin, K.; Satija, S. K.; Sokolov, J.; Rafailovich, M. H., Effect of Nanoscopic Fillers on Dewetting Dynamics. Macromolecules 2006, 39, (8), 2972- 2980. 128 76. Holmes, M. A.; Mackay, M. E., Dynamics and Stability of Solvent Annealed Polymer Films: Influence of Molecular Weight and Film Thickness. in preparation. 77. Harth, E.; Van Horn, B.; Lee, V. Y.; Germack, D. 8.; Gonzales, 0. P.; Miller, R. D.; Hawker, C. J., .A facile approach to architecturally defined nanoparticles via intramolecular chain collapse. Journal of the American Chemical Society 2002, 124, (29), 8653-8660. 78. Tuteja, A.; Mackay, M. E.; Hawker, C. J.; Van Horn, B.; Ho, D. L., Molecular architecture and rheological characterization of novel intramolecularly crosslinked polystyrene nanoparticles. Journal of Polymer Science Part B- Polymer Physics 2006, 44, (14), 1930-1947. 79. Nelson, A., Co-refinement of multiple-contrast neutron/X-ray reflectivity data using MOTOFIT. Joumal of Applied Crystallography 2006, 39, (2), 273-276. 80. Krishnan, R. 8.; Mackay, M. E.; Duxbury, P. M.; Pastor, A.; Hawker, 0. J.; Van Horn, B.; Asokan, 8.; Wong, M. 8., Self-assembled multilayers of nanocomponents. Nano Letters 2007, 7, (2), 484-489. 81. Bezmel'nitsyn, V. N.; Eletskii, A. V.; Okun, M. V., Fullerenes in solutions. Uspekhi Fizicheskikh Nauk1998, 168, (11), 1195-1220. 82. Seemann, R.; Herminghaus, 8.; Jacobs, K., Dewetting patterns and molecular forces: A reconciliation. Phys. Rev. Lett. 2001, 86, (24), 5534-5537. 83. Seemann, R.; Herminghaus, 8.; Jacobs, K., Gaining control of pattern formation of dewetting liquid films. Journal of Physics-Condensed Matter 2001, 13, (21), 4925-4938. 84. Zolotukhin, l. V.; Yanchenko, L. l.; Belonogov, E. K., Fractal structure of fullerite. Jetp Letters 1998, 67, (9), 720-722. 85. Mukherjee, M.; Singh, A.; Daillant, J.; Menelle, A.; Cousin, F., Effect of solvent-polymer interaction in swelling dynamics of ultrathin polyacrylamide films: A neutron and X-ray reflectivity study. Macromolecules 2007, 40, (4), 1073-1080. 129 86. Singh, A.; Mukherjee, M., Effect of polymer-particle interaction in swelling dynamics of ultrathin nanocomposite films. Macromolecules 2005, 38, (21), 8795-8802. 87. Singh, A.; Mukherjee, M., Swelling dynamics of ultrathin polymer films. Macromolecules 2003, 36, (23), 8728-8731. 88. Yim, H.; Kent, M.; McNamara, W. F.; lvkov, R.; Satija, 8.; Majewski, J., Structure within thin epoxy films revealed by solvent swelling: A neutron reflectivity study. Macromolecules 1999, 32, (23), 7932-7938. 89. Mackay, M. E.; Tuteja, A.; Duxbury, P. M.; Hawker, C. J.; Van Horn, B.; Guan, 2.; Chen, G.; Krishnan, R. 8., General Strategies for Nanoparticle Dispersion. Science 2006, 31 1, (5768), 1740-1743. 90. Fowkes, F. M.; Mccarthy, D. 0.; Mostafa, M. A., Contact Angles and the Equilibrium Spreading Pressures of Liquids on Hydrophobic Solids. Journal of Colloid and Interface Science 1980, 78, (1 ), 200-206. 91. Holmes, M. A.; Mackay, M. E., In preparation. 92. Krishnan, R. 8.; Mackay, M. E.; Duxbury, P. M.; Hawker, C. J.; Asokan, 8.; Wong, M. 8.; Goyette, R; Thiyagarajan, P., Improved polymer thin-film wetting behavior through nanoparticle segregation to interfaces. Submitted. 93. Bucknall, D. G., Neutron reflection studies of polymers. In Modern Techniques for Polymer Characterisation, Pethrick, R. A.; Dawkins, J. V., Eds. Wiley: New York, 1999; pp 109-140. 94. Karim, A.; Arendt, B. H.; Goyette, R.; Huang, Y. Y.; Kleb, R.; Felcher, G. P., An Automated Neutron Reflectometer (Posy-Ii) at the Intense Pulsed Neutron Source. Physica B1991, 173, (1-2), 17-24. 95. Nelson, A. Motofit, Australian Nuclear Science and Technology Organisation. 96. IGOR Pro, WaveMetrics: Lake Oswego, OR. 130 IIIIIIIIIIIIIIIIIII llllllllllllililllllllllll111111111111 3 1293 02956 1028