PLASMA APPLICATION FOR SYNTHESIS OF NOVEL NANOSTRUCTURES AND MECHANICAL PROPERTIES INVESTIGATION By Alborz Izadi A DISSERTATION Michigan State University in partial fulfillment of the requirements Submitted to for the degree of Mechanical Engineering – Doctor of Philosophy 2020 PLASMA APPLICATION FOR SYNTHESIS OF NOVEL NANOSTRUCTURES AND MECHANICAL PROPERTIES INVESTIGATION ABSTRACT By Alborz Izadi Today, nanomaterials are receiving increasing attention, as they exhibit exciting and useful at- tributes that can be employed across applications and disciplines range from electronics to drug delivery. This dissertation focuses on two disparate projects, both related to plasma-synthesized nanocrystals. The two projects highlight the unique properties of nanocrystals, and their sustain- able synthesis using gas phase approaches. These methods apply non-thermal plasma reactors for nanocrystal synthesis. First, we introduce a plasma-based gas phase method for Gold nanoparticle (AuNPs) synthesis. Second, we use plasma-synthesized Silicon nanocrystals (SiNCs) in bilayer and composite structures with a commonly used elastomer, to investigate the mechanical behavior of the structures in a first-of-its kind investigation. There are many applications for AuNPs due to their interesting optoelectronic properties, such as tunable optical absorption and plasmonic resonance behavior. While synthesis and stabilization of colloidal AuNPs is well-established, new synthesis routes can lead to enhanced versatility of applications for AuNPs, particularly if the methods allow avoidance of solution processes or surfactants. In Chapter 2, we introduce a plasma-based synthesis of AuNPs, using a consumable gold wire and a radiofrequency power source. The AuNPs are monodisperse, with an average diameter of 4 nm. While production yield is low, the narrow size distribution of the AuNPs and the avoidance of solution processing in this method are promising for future syntheses of metal NPs based on plasmas. Next, a comprehensive analysis of the mechanical behavior of SiNC/PDMS systems, using plasma-produced SiNCs has been performed. Chapter 3 details our experimental methods com- bined with modeling to estimate the mechanical behavior of thin layers of SiNCs on PDMS, as deposited directly onto the PDMS from non-thermal plasma reactor. For the first time we estimated the mechanical behavior of thin films of SiNCs by using the onset of bifurcations as an indicator of their modulus. Next, reaching towards luminescent nanocomposites for applications in luminescent devices, we investigated the optical and mechanical behavior of blended SiNC/PDMS nanocom- posites. The results from these investigations, reported in Chapter 3 and Chapter 4, have shed light for the first time on the interactions between SiNCs and PDMS both in bilayer and composite struc- tures, pointing to future optoelectronic and opto-mechanical device applications with predictable properties. Finally, in Chapter 5, we share ongoing projects which will be finalized soon, as well as detailing future work surrounding these ideas. We share our parametric study to uncover the various eects of surface functionality, SiNC layer thickness, and PDMS modulus on the resulting SiNC thin film. Concurrently, we probed the formation of wrinkles and cracks on SiNC film surfaces, which were deposited on pre-stretched PDMS using a finite bending configuration, to examine how instabilities on the thin films can be predicted. Copyright by ALBORZ IZADI 2020 To my mother, Shahla, my father, Hossein, and my sister, Dena v ACKNOWLEDGEMENTS I would like to thank all the people who contributed in some way to the work described in this thesis. First of all, I would like to express my sincere gratitude to my advisor, the most wonderful woman I have met in the U.S., professor Rebecca Anthony, for the continuous support of my PhD research, her patience, enthusiasm, and knowledge . When I joined her research group in 2014, I had absolutely no experience in research. Without her guidance and support, finishing this dissertation would not have been possible. In addition, I thank our collaborator, professor Sara Roccabianca. Without her eorts my job would have undoubtedly been more dicult. I am thankful to my committee members, professor Anctil and professor Yeom, whose comments have always been helpful. Every result described in my thesis was accomplished with the help and support of fellow lab mates and collaborators from Plasma and Nanomaterial Laboratory (PNL). I am especially thankful to Kyle O’Shea, who was my closest friend during the last three years of graduate school, Alex Ho, the most reliable lab-mate during my research , Dr. Gabriel Ceriotti, who always had useful ideas and suggestions for me, Geena Du, Brant Toback, Michael Bigelow and many others who are all my best friends. I am very grateful for the modeling data shared with me from Dr. Roccabianc’s lab. I also would like to thank all members of Mechanical Engineering, Chemical Engineering and Material Sciences Departments at MSU, specially Dr. Askeland, professor Naguib and professor Wichman. I thank all the journals which gave me the permission to include copyrighted pictures as part of my dissertation. I am also thankful for the funding sources that allowed me to pursue my graduate school studies including National Science Foundation. I am thankful to all the members of Advanced Microscopy Center Facility at MSU, who helped and mentored me through my sample preparation process. I would like to acknowledge all my friends, professor Babak Borhan, Dr. Reza Loloee, professor Maghrebi, Dr. Mehdi Nasr, Mohammad, Fariborz and Alireza who supported me during my stay at MSU and made my time a lot more fun. Finally, I would like to thank my mom, Shahla, my dad, Hossein, and my sister, Dena, for their constant love and support. My deepest appreciation goes to my parents, who dedicated their own vi time and energy to support me in all aspects of my career and acadamic desires. Their video calls always motivated me to stay strong. I have a wish to always have their back and give my best support to them. I will never forget their sacrifices in tough moments of my journey. Without their support, help and dedication, it would be absolutely impossible for me to pursue graduate studies abroad and be where I am today. Another deep appreciation from the bottom of my heart goes to my sister, Dena, who supported me through moments of hopelessness and played the role of my family during my stay in East Lansing. vii TABLE OF CONTENTS . . . . . . . . . . . . . x xi 1.1 1.2 Gold Nanoparticles . . . . . INTRODUCTION . LIST OF TABLES . LIST OF FIGURES . CHAPTER 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Introduction to Non-thermal Plasma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2.1 History of AuNP synthesis . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2.2 AuNP formation in the gas phase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1 4 5 5 8 1.3.1 SiNCs on Flexible Substrates: Polydimethylsiloxane (PDMS) . . . . . . . . 11 1.3.2 Mechanical properties of SiNCs, deposited as thin films . . . . . . . . . . . 12 CHAPTER 2 AUNP SYNTHESIS: A PLASMA-BASED APPROACH . . . . . . . . . . . 14 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 . . 2.1 2.2 Experimental Details . 2.3 Results and Discussion . 2.4 Conclusion . . 1.3 Silicon Nanoparticles . Introduction . . . . . . . . . . . . . CHAPTER 3 ESTIMATING THE MECHANICAL PROPERTIES OF SINC LAYERS 3.1 3.2 Methods . . . . . . . . . . . . . Introduction . . ON PDMS . . . . . 3.2.1 3.2.2 3.2.3 3.2.4 Uniaxial tensile test . 3.2.5 3.2.6 Modeling bifurcation in elastic layered structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32 PDMS fabrication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33 Plasma synthesis of SiNCs . . . . . . . . . . . . . . . . . . . . . . . . . . 33 Silicon nanocrystal film characterization . . . . . . . . . . . . . . . . . . . 34 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36 . . . . . . . . . . . . . . . . . . . . . . . . . . 37 . . . . . . . . . . . . . . 38 Finite bending . . . . . . . . . . . . . . . . . . . . . . . . . . . 39 . . . . . . . . . . . . . . . . . . . . . . . . 40 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45 3.2.6.1 3.2.6.2 Bifurcation analysis Finite bending experiment . . 3.3 Results and Discussion . 3.4 Conclusion . . . . . CHAPTER 4 OPTICAL AND MECHANICAL BEHAVIOR OF SINC-EMBEDDED . . . . . . Introduction . PDMS SLABS . 4.1 . . . 4.2 Experimental Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49 . Plasma synthesis of SiNCs . . . . . . . . . . . . . . . . . . . . . . . . . . 49 4.2.1 Functionalization of SiNCs . . . . . . . . . . . . . . . . . . . . . . . . . . 50 4.2.2 4.2.3 PDMS preparation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50 4.2.4 Uniaxial tensile testing . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51 viii . . BIBLIOGRAPHY . . . 4.4 Conclusions . . . CHAPTER 5 FUTURE . . . . . . . . . . . . . . . 4.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51 4.3.1 SiNC characterization and nanocomposite formation . . . . . . . . . . . . 51 4.3.2 Nanocomposite characterization . . . . . . . . . . . . . . . . . . . . . . . 53 Photoluminescence . . . . . . . . . . . . . . . . . . . . . . . . . 53 4.3.2.1 4.3.2.2 Microscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53 4.3.2.3 Mechanical properties of the composite . . . . . . . . . . . . . . 57 4.3.2.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62 . . . . . . . . . . . . . 63 5.1 AuNP Application for Silicon NanoRods (SiNRs) Sythesis 5.2 Modulus of SiNC Layers on PDMS: Parametric Investigation . . . . . . . . . . . . 68 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75 . 77 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77 SiNC deposition . . . . . . . . . . . . . . . . . . . . . . . . . . 77 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89 Introduction . 5.3.1.1 5.3.2 Results . 5.3.3 Conclusion . Future work . 5.3.4 . . 5.2.1 . 5.2.2 Experimental work . 5.2.3 Results . 5.2.4 Conclusions . 5.3 Unbending- Another Finite Bending Condition for Mechanical Property Estimation Introduction . 5.3.1 . . . . . . . . . . . . . . . . . . . . . . . . . . ix LIST OF TABLES Table 2.1: Reaction parameters for plasma-based AuNP synthesis . . . . . . . . . . . . . . 18 Table 2.2: Comparing plasma-based and hotwire methods . . . . . . . . . . . . . . . . . . 25 Table 3.1: Critical angles of bifurcation for dierent slenderness ratios, expressed in degrees. Angles are reported for ⇤ = 2, 3 and 5. . . . . . . . . . . . . . . . . . . 42 Table 5.1: List of operation used for making surface instabilities in SiNC plasma reactor . . . . at PNL . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84 x LIST OF FIGURES Figure 1.1: Hot wire synthesis of AuNPs as developed by Boies et al. . . . . . . . . . . . . 7 Figure 2.1: An overview of gold nanoparticle syhthesis . . . . . . . . . . . . . . . . . . . . 15 Figure 2.2: Photograph and schematic of plasma-based AuNP synthesis. Insets show TEM images of AuNPs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 Figure 2.3: Schematic and photograph of hot-wire AuNP synthesis, including inset table of reaction parameters. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19 Figure 2.4: Imaging of AuNPs using transmission electron microscopy (TEM). (a) Bright field and (b) dark field low-magnification images of AuNPs collected on carbon fiber TEM grid; (c) higher-magnification TEM image of single Au NP with [2 0 0] crystallographic plane identified as based on (d) FFT of the same AuNP. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19 . . . . . . . . Figure 2.5: (a)STEM imaging confirms crystalline AuNPs, with inset showing the scan- ning line for EDS analysis. (b) EDS confirms the composition of the AuNP. . . 20 Figure 2.6: AuNP average diameter and standard deviation of various samples from Ta- ble.2.1, as measured from TEM images . . . . . . . . . . . . . . . . . . . . . . 21 Figure 2.7: Figure 2.8: (a) SEM image of the consumable Au wire surface before plasma exposure. (b) SEM image of the Au wire surface after plasma exposure (conditions as for Sample 2 in Table2.1), showing noticeable roughness. (c) Higher magnification of SEM image of the Au wire after the plasma process. . . . . . . . . . . . . . . . . . . . . . . . 22 (a,b) TEM images of AuNPs as synthesized using plasma-based method (Ar: 900 sccm, pressure: 180 Torr) and the hot-wire method (Ar: 900, DC power: 12 W) respectively. (c) Average AuNP diameter and standard deviation for the plasma-based method (circular symbols, Samples 1 and 2 from Fig.2.1) and the hot-wire method (square symbols, Samples 3 and 4 from Fig.2.3) as measured using TEM image analysis. AuNP, gold nanoparticle; DC, direct current; TEM, transmission electron microscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24 . . . . . . Figure 2.9: OES test for Plasma-Based AuNP synthesis, 340-1050 nm spectrometer with Hydrogen(red) and without Hydrogen(blue) . . . . . . . . . . . . . . . . . . . 28 xi Figure 3.1: Steps from plasma deposition to finite bending. SiNCs are deposited through a slit-shaped orifice onto PDMS, which is then bent to observe formation of wrinkles on the SiNC surface. (b) Schematic diagram of reactor and SiNC synthesis. (c) SiNC/PDMS sample illuminated with UV light, demonstrating photoluminescence. (d) Finite bending applied with a lab-designed apparatus. . . . . . . . . . . . . . . . . . . 32 (a) Photograph of plasma reactor. Figure 3.2: SEM and FIB cross-sectioning of the SiNC layer. The thickness is estimated to be (⇠4.50 µm). (a) SEM image showing the surface of the SiNC layer on PDMS, including the FIB-milled area (inset). (b) Higher magnification SEM image of the FIB-milled area demonstrating the thickness of the SiNC layer. . . 35 Figure 3.3: (a) X-ray diraction confirmed formation of SiNCs. XRD for SiNCs shows crystal reflections at 2✓ values 28.3, 47.3 and 56.1 degrees consistent with the (111), (220), and (311) planes of silicon. (b) TEM image confirming the formation of SiNCs: black circles highlight individual SiNCs. The inset shows FFT of the correspondig SiNCs of the TEM image. . . . . . . . . . . . . 36 Figure 3.4: The stress machine loaded with a colored PDMS and dots depicted on the PDMS, dislocation of dots will be recorded by a digital camera . . . . . . . . . 37 Figure 3.5: (a and b)A schematic diagram of proposed aparatus for finite bending mea- surement desgned by Solid Work software and (c) Alluminum cast aparatus designed at PNL . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37 . . Figure 3.6: PDMS sample in the finite bending apparatus. (a) Side view, showing sam- (b) Top view showing ple (with handles) in the finite bending apparatus. formation of bifurcations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38 Figure 3.7: Reference and deformed configurations for the PDMS-SiNC bilayer. The bilayer deforms from the Cartesian coordinate system in the reference config- uration (e1, e2, e3) to a cylindrical coordinate system in the deformed config- uration (er, e✓, ez). Specifically, a plane at constant x0 1 transforms to a plane at constant r (dashed line), and a plane at constant x0 2 transforms to a circular arc at constant ✓ (dashed-dotted line). Since the out of plane deformation is taken to be zero, we have x0 . . . . . . . . . . . . . . . . . . . . . . . . . 39 3 = z. xii Figure 3.8: Critical angles of bifurcation evaluated as a function of ratio of material properties for a PDMS-SiNC bilayer sample. The ratio of the neo-Hookean material parameters is defined as ¯µ0 = µNC 0 . All curves have been calculated for a thickness ratio H included in the range 0.0013–0.0018. Each line represent a dierent value of slenderness ratio, ⇤ = 2, black solid line; ⇤ = 3, black dashed line; ⇤ = 4, dark gray solid line; ⇤ = 5, dark gray dashed line; ⇤ = 6, light gray solid line. The insert represents a schematic of the bilayer and defines the slenderness ratio. 0 /µP . . . . . . . . . . . . . . . . . . . . . 42 Figure 3.9: Plotting points from the experiments on graph obtained by numerical cal- culations. The ratio of the neo-Hookean material parameters is defined as ¯µ0 = µNC 0 . Shown here are points obtained for ⇤ = 2, 3 and 5. Insert shows an instability on the surface of a sample with ⇤ = 2. 0 /µP . . . . . . . . . . . 43 Figure 4.1: SiNC Characterization: XRD (a) shows crystallographic reflections for nano- sized silicon. Gray lines are the actual data and the black trace is the peak fit. TEM imaging (b) indicates spherical nanocrystals and lattice spacing consistent with crystalline silicon. Inset shows FFT of image. . . . . . . . . . . 52 Figure 4.2: SiNCs in solution, (a) photographed in room light and b) under UV excitation Shown are hexane, NF, and F samples from left to right. (c) PL spectroscopy shows much brighter luminescence for surface-functionalized (F) SiNCs as compared to nonfunctionalized (NF) SiNCs in solution. . . . . . . . . . . . . . 53 Figure 4.3: (a) Photographs of SiNC/PDMS nanocomposites in room light and (b) under UV excitation (c) shows the average PL QY of SiNCs in hexane and in the nanocomposites, with error bars reflecting the standard deviation of the measurements. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54 . . . Figure 4.4: (a) For reference during microscopy, we assigned each nanocomposite the same coordinate system, (b) A single CLSM image of a slice is shown and (c) the depth-resolved composite image (scale inset), demonstrates the distri- bution of small clusters within the sample slab. . . . . . . . . . . . . . . . . . 55 Figure 4.5: Representative CLSM images of the SiNC/PDMS nanocomposites . . . . . . . 56 Figure 4.6: SiNC cluster size statistics based on image processing from the CLSM images. 3 separate images were analyzed for each sample type. A key to the data representative is shown as inset. . . . . . . . . . . . . . . . . . . . . . . . . . 56 Figure 4.7: Mechanical properties of the nanocomposites. (a) shows a representative plot of Cauchy stress vs. stretch data for each sample type. From this data, (b) we extracted the neo-Hookean coecient for each sample, showing a distinct reduction for the higher-concentration sample. (c) the rupture stretch was unaected by the SiNC inclusions. . . . . . . . . . . . . . . . . . . . . . . . . 58 xiii Figure 5.1: (a) Schematic Diagram for SiNRs synthesis (b) early stage experimental set up with hot wire method (Courtesy R.J.Anthony) . . . . . . . . . . . . . . . . . 64 Figure 5.2: High flowrate of AuNP-Ar from the first step(600-1200sccm) and low flowrate of SiH4 from the side arm did not mix properly in the second step of the reactor 65 Figure 5.3: Modified hotwire AuNP reactor for sythesis SiNRs . . . . . . . . . . . . . . . . 65 Figure 5.4: (a) Filament formation associated with high plasma density (b) High plasma density resulted in wide size range of Silicon nanoparticles with both crys- talline and amorphous structures. (c) Controlling the SiH4 flowrate and RF power stoped fillament and resulted in narrow size distribution for Silicon nanoparticles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67 . . . . Figure 5.5: Schematic of dual-plasma reactor for AuNPs-SINRs simultaneous synthesis . . 68 Figure 5.6: Schematic diagram for investigating the impact of film thickness, surface functionality of SiNCs via gas phase plasma and shear modulus of the PDMS substrate on mechanical properties of SiNC thin film . . . . . . . . . . . . . . . 71 Figure 5.7: SEM images for PDMS with 3:1 slenderness ratio as defined in chapter 3 with rastering deposition time of 5 mins (Sample 1), 10 mins (Sample 2) and 30 mins (Sample 3) and averaged measured thickness of the film with corresponding standard deviation in comparison with the sample discussed with same slenderness ratio and 20 mins (in red)of rastering deposition time from previous chapter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72 Figure 5.8: The methodology introduced in Chapter 3, employed to estimate neo-Hookean coecient for SiNC layars corresponds to parametric variation. The asterisks indicate samples with significant deviation from value we abrtained in Chapter 3 73 Figure 5.9: Neo-Hookean coecient of all sample introduced in this section.The sam- ples with asterisks, present the data with siginficant deviation from samples reported in Chapter 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74 Figure 5.10: The surface functionalization applied in gas phase during plasma processing and sythesis of SiNCs. The PL and QY measured. This results coupled with mechanical property estimation of corresponding samples presented in this Figure. The results here, confirm that improving the oprical properties and QY do not interfere with mechanical property of the SiNC thin film. . . . . . . 75 Figure 5.11: Sample holder for various stretching ratios (left), sample holder for finite bending around pushing rod (right) . . . . . . . . . . . . . . . . . . . . . . . . 78 Figure 5.12: Above view of sample, relaxed after deposition . . . . . . . . . . . . . . . . . 78 xiv Figure 5.13: SEM image of deposited area with 20 percent pre-stretching condition, Col- umn A3-B3-C3 (left), Columns A2-B2-C2 and A4-B4-C4 (middle), Columns A1-B1-C1 and A5-B5-C5 (right) . . . . . . . . . . . . . . . . . . . . . . . . . 79 Figure 5.14: SEM image of deposited area with 20 percent pre-stretching condition, Col- umn A3-B3-C3 (left), Columns A2-B2-C2 and A4-B4-C4 (middle), Columns A1-B1-C1 and A5-B5-C5 (right) . . . . . . . . . . . . . . . . . . . . . . . . . 79 Figure 5.15: pre-stretched before deposition or during deposition . . . . . . . . . . . . . . . 80 Figure 5.16: Early design for finite bending surface instabilities by wrapping PDMS around the rod, the picture on right is the microscope image of the sample after unwrapped with the scale bar of 10 micrometer . . . . . . . . . . . . . . . . . . 82 Figure 5.17: A new method for pure bending surface instabilities, the mathematical cor- relation and assumptions for incompressible, Neo-Hookean material (left) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83 . . . . . . . . . . . Figure 5.18: A new method for pure bending surface instabilities, the mathematical cor- relation and assumptions for incompressible, Neo-Hookean material (left) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83 . . . . . . . . . . . Figure 5.19: XRD for SiNCs deposited on PDMS, 30 Watt(Left), 25 Watt (right) . . . . . . . 85 Figure 5.20: Lambda () measurement for surface instabilities, 10 round rotations in the reactor(up-left), 20 round rotations(up-right) and 30 round rotations (bottom), numbers are corresponding to the cells that measurement done with specified standard deviation for each averaged size letter corresponding to the cell which wrinkle instabilities observed . . . . . . . . . . . . . . . . . . . . . . . 86 Figure 5.21: A sample of PDMS deposited with SiNCs under pre-stretched condition for 10 round rotation(left), definition of Lambda () and thickness of the SiNCs film(right) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86 . . . . . . Figure 5.22: Atomic Forced Microscopy (AFM), AFM scanned a line across a 20 by 20 micrometer area and provided a 3D model and a cross-sectional view for investigating the amplitude of these surface instabilities . . . . . . . . . . . . . 87 xv CHAPTER 1 INTRODUCTION The focus of this dissertation is synthesis and use of nontoxic nanomaterials for a variety of applications. Nanomaterials such as nanocrystals are critical for a range of novel applications ranging from biological treatments and diagnostic procedures to devices such as light-emitting devices (LEDs), solar photovoltaics, and wearable sensors. There are two main thrusts to the research: one focuses on new gas-phase routes to freestanding, surfactant-free gold nanoparticles, including a novel plasma reactor. The other explores the optical and mechanical behavior of plasma-synthesized silicon nanocrystals. This introductory chapter sets the stage for discussing the subsequent research by providing a background on plasmas, the history and use of gold nanoparticles, and the prior work on studying the mechanical behavior of thin films of silicon nanocrystals. 1.1 Introduction to Non-thermal Plasma Today, many research studies have been dedicated to investigating the synthesis and use of nanocrystal-based materials, including those processed in the liquid phase [83, 122, 164]. While liquid-phase synthesis and/or surface modification has yielded exceptional results for some nanocrystalline materials, there are limitations imposed by solution processing. For example, the temperature of liquid phase synthesis methods are limited by the boiling points of solvents [83]. This can hamper the formation of some nanocrystals due to the required high crystalliza- tion temperatures, though some approaches have been successful for materials such as silicon nanocrystals [36, 63, 64, 189]. Additionally, liquid-phase synthesis and/or surface modification processes necessitate disposal procedures for the solution wastes and by-products which are time- consuming and are not environmentally sustainable. In contrast, gas phase approaches do not require organic solvents during the synthesis process and are capable of achieving nanocrystals from high-temperature-crystallization materials. There are two classes of gas phase synthesis 1 methods, dierentiated according to the thermal equilibrium during the process. Thermal equi- librium syntheses include flame synthesis [83, 158], thermal pyrolysis in furnaces [129], laser pyrolysis [83, 91] and thermal plasmas [142]. In this class of gas phase synthesis, the entire system achieves a relatively high temperature (e.g. gas species, excited species, and synthesized nano- materials). During many thermal-equilibrium processes, the produced nanoparticles have reduced charging and a large degree of agglomeration. Some liquid phase post-processing steps are often be required in this regard. Also, higher temperature thermal plasma increases the chance of broad size distribution [83]. Gas-phase processes that are not in thermal equilibrium (such as non-thermal plasmas) are the alternate class of method for nanoparticle synthesis. In non-thermal equilibrium we have heavy gas species and ions that are close to room temperature, while free plasma electrons (which have negligible mass in comparison) can achieve temperature between 1000 to 50000 K [83, 107]. Collisions between plasma electrons and molecules in gas flow dissociate and ionize gaseous precursors, producing highly reactive components [83]. Non-thermal plasmas have several of advantages including ligand-free, high-purity final products. High melting point materials can be synthesized due to highly exothermic reactions just at the surfaces of the forming nanoparticles. In non-thermal plasmas, synthesis of very small nanoparticles (with sizes below 10 nm and size distributions around 10-20% std. dev.) are possible. Diusion loss to the walls, together with nanoparticle agglomeration, are negligible due to negative charges of surface of the reactor walls as well as at the surfaces of nanoparticles - both of these are the result of the high kinetic energy and collision frequency of electrons with surfaces in the plasma. Additionally, nanoparticle growth occurs over a particle residence time on the ms time scale [3, 83, 107]. These attributes make nonthermal plasmas an interesting substitute for liquid phase synthesis. In general, nonthermal plasmas are classified by low ionization fraction. In radiofrequency (RF) plasmas, such as the ones used in this dissertation, ions are generated by the applied rf field. This field causes the free electrons to accelerate to high speed. The number of free electrons exponentially increases in the plasma and few of these electrons are able to exchange energy with heavier gas species, such as precursor molecules [83]. Most of the energy provided to the plasma 2 also absorbed with electrons since they have higher mobility compared to ions. The densities of electrons and ions remain approximately equal due to the principle of quasineutrality in the plasma, except near the walls (this small region of non-neutrality is called the plasma sheath) [18, 92]. The dierence between speed and temperature of electrons and ions in the plasma is what establishes the condition of non-thermal equilibrium, and is also the cause of the negative charging of surfaces (i.e. walls and nanoparticles) in the plasma [25, 83]. In the non-thermal plasma, nanoparticles will be charged to negative potentials in the order of kBTe/2e. By considering the typical temperature of electrons, individual nanoparticles will have a charge of up to few volts. As an example, a typical particle flowing in the plasma with diameter of around 1 nm carrying a single electron will have a negative potential of -1.4 V [83, 136]. In order to validate a charge balance between electron and ion currents reaching the surface of the nanoparticles, it is essential to note that nanoparticles of dierent sizes will charge to approximately the same steady state potential. Bouchoule et al. [17, 83] proposed a three-steps mechanism for nanoparticle growth in non- thermal plasmas. Although the repulsion between negative charges prevents nanoparticles from agglomeration, the smaller particles at earlier stage of growth (nanoparticles with size less than 1 nm) usually have no charge and their aggolomeration is one of the main mechanisms for nanoparticles growth. First, initial clusters of precursor fragments generate at nucleation stage, forming proto- nanoparticles. Second, since the nanoparticles emerging from the first stage are mostly less than 1nm in size and has no charge, larger particles grow quickly through the coagulation of initial neutral small clusters [83]. This step continues up to point that all of nanoparticles have single charge. Finally, the growth of nanoparticles continues by reaction of precursor on their surface. Non-thermal plasma reactors can operate under reduced pressure or atmospheric pressure. These reactors can include various components and characterization tools including the reaction chamber, vessels for precursor injection [83], gas processing pipelines, analysis ports, and the nanomaterial collection system (often by impaction or diusion). The reaction chamber can be modified over a vast range for flow-through and batch processing requirements. 3 This dissertation research employs plasma reactors for nanomaterials synthesis in two separate projects. A background on alternate methods provides useful context. In the next section we briefly review prior methods for fabrication of noble metal nanoparticles, pertinent to the chapter on gold nanoparticle synthesis using a novel non-thermal plasma reactor. Then we briefly talk about silicon based- nanomaterials for electronic, sensor, and optoelectronic applications, especially nanoparticles produced by non-thermal plasmas. In addition, we introduce poly(dimethylsiloxane) (PDMS) as a stretchable substrate with some unique physical properties that make it suitable for probing the mechanical behavior of silicon nanoparticle layers. Finally, we review some existing mechanical models for estimating the mechanical properties of a thin film of semiconductor layers on a flexible substrate. 1.2 Gold Nanoparticles Gold nanoparticles (AuNPs) are one of the oldest forms of nanostructures, with medical appli- cations dating back to 1618 [15]. Today scientists are especially interested in Localized Surface Plasmon Resonance (LSPR), a phenomenon of unique optical shape-dependent property with quan- tum features, which results from the small size of these structures in less than 10nm. LSPR is an optical property which occurs when a noble metal nanoparticle interacts with a photon[43]. Depending on the metal, particle size, particle shape, and photon energy, the free electrons in the nanoparticle can undergo resonant oscillations and amplify or otherwise modify the electromag- netic field nearby. LSPR has two main consequences. First, the electric field near the particle’s surface is greatly enhanced; second, the nanoparticle’s optical absorption has a maximum at the plasmon resonant frequency (which is dependent on metal, size, and shape). The enhancement decreases quickly with increasing distance from surface of the particle. LSPR occurs at visible wavelengths for the noble metal nanoparticles including AuNPs [21, 71, 90, 98, 101, 147, 198]. The LSPR eects (both for emission and absorption) from AuNPs can be used in many applications including as biological labels, therapeutic cancer treatments, solar photovoltaics, near-field and electron microscopy, and more [47, 70]. 4 In addition to uses centered around their LSPR, metal nanoparticles can be used to catalyze the growth of anisotropic semiconductor nanostructures such as silicon nanorods (SiNR) [63]. These anisotropic structures oer novel and shape/size-dependent optoelectronic properties, and so the catalysis oered by AuNPs and others adds to their attractive properties. 1.2.1 History of AuNP synthesis Among the synthesis methods for AuNPs production, many involve wet chemistry-based ap- proaches, which require ligands and other solution-phase modifiers to the naoparticle surfaces [101, 121]. Solution based syntheses are among the earliest methods for formation of gold nanos- tructures and are still used today[20, 35, 39, 49, 54, 101, 171, 197]. In general, AuNPs in solution phase produced by reduction of chloroauric acid. In these methods the naoparticle growth happens step by step from molecular precursors (also called bottom-up), leading to better control on the shape and size distribution of AuNPs in the final product. In the solution-based routes, the AuNPs end up solubilized in water or other solvents. We endeavored to eliminate the possibility of con- tamination from these liquids and extend the uses of AuNPs by synthesizing the AuNPs in the gas phase. This will open up a new approach for AuNPs synthesis and reduce both product contami- nation as well as reduce the need for solution disposal after-process. Our research builds on prior work to synthesize AuNPs in the gas phase, particularly aiming towards a synthesis method that permits easy integration of the AuNP product into subsequent reaction steps without interruption of process parameters. We review these pre-existing methods below, and in Chapter 2 we introduce our new plasma-based method. 1.2.2 AuNP formation in the gas phase Other work has led to gas-phase, solvent-free synthesis routes to AuNPs. There are several options for synthesis in the gas-phase, including inert gas condensation of Au vapor in a chamber, spark-ignition AuNP generation, sputtering techniques, and hot-wire synthesis. Here we detail the hot-wire synthesis method, which has demonstrated promise in multi-step reaction procedures. 5 [15]. This approach is similar to other gas-phase methods in that it requires a gas flow and a heat source for generating the AuNPs. Since the metal wire is heated via electrical resistance, the average background gas temperature will not increase except in the area locally close to the resistive element. Under the right conditions of gas flow, pressure, and suplied electrical current, the evaporated metal atoms reach supersaturation upon leaving the metal wire surface, and rapidly nucleation of nanoparticles begins. Gas drag then carries the AuNPs downstream for collection or further use. The big advantage of the hot wire method is that according to previous studies, background gas has small eect on property of metal nanoparticles. Peineke et al. described that the hot wire method is suitable for all metals which have vapor pressure of higher than (3.75*10-5 Torr) at the melting point. According to this guideline, gold (with a vapor pressure of 1.5*10-5 Torr) does not meet this requirement. However, as Boies et al. reported [15], gold can be used if the source is an electroplated wire structure. For example, they had success using a Platinum (Pt) (99.5%, trace metal, Aldrich) wire with 0.5 mm in diameter as the core. The gold as the outer shell was deposited via electroplating by the ProPlate Company in Minnesota. Studies have shown that hot wire method with a coaxial gold source can produce AuNPs with less than 10 nm average diameter [15, 16]. Figure1.1 depicts the schematic diagram of hot wire gold nanoparticle formation method as developed by Boies et al. This method is promising as it results in AuNPs entrained into the gas flow – but had only been reported at atmospheric pressure prior to our work. For coupling into a low-pressure reactor for possible multi-step reactions including non-thermal plasmas, we were interested to explore this method as an AuNP catalyst source at lower pressure. We were also interested in other methods for AuNP synthesis at lower pressure. Our work on a reduced-pressure approach to the hot wire synthesis of AuNPs is detailed in Chapter 2, alongside our introduction of a plasma-based synthesis of AuNPs, using a consumable gold wire and a radiofrequency power source. In our plasma method, the AuNPs are mono-disperse, with an average diameter of 4 nm, oering some improved properties as compared to our attempt at a low-pressure hot wire synthesis route. Notabely, this size range is one of the most interesting for optoelectronic and LSPR functionality for AuNPs. As shown in that chatper, although the 6 Figure 1.1: Hot wire synthesis of AuNPs as developed by Boies et al. production yield is low, the narrow size distribution of the AuNPs and the avoidance of solution processing in this method are promising for future syntheses of metal NPs based on plasmas. 7 1.3 Silicon Nanoparticles The other focus of this dissertation is on deployment of silicon nanoparticles onto and within polymer substrates for evaluation of optical and mechanical properties of these bilayered and hybrid macroscale structures. Silicon has been a dominant material in the field of electronics over the last half century or more, due to its semiconducting electronic behavior, optical band gap at 1.1 eV, its natural abundance, its low cost, and its relative nontoxicity. Despite the long history of use in transistors, traditional solar photovoltaics, bulk silicon has poor emission behavior. Due to its indirect band gap, silicon requires that photon emission and absorption processes involve a momentum-balancing phonon [107]. This disadvantage limits the applications of bulk silicon as a potential material for host of optoelectronic devices including emissive applications, although doping, alloying, and combination of bulk silicon into hybrid structures with other direct band gap semiconductors has allowed some device development in the area of optoelectronics. [75]. However, since the late 1980s, new discoveries on the behavior of nanoscale silicon have revealed that optical processes become much more ecient when the silicon becomes quantum- confined, meaning its size becomes small enough that electron-hole combination processes are aected. In particular, when the silicon nanostructure size shrinks to approximately the Bohr exciton radius (around 5 nm), the electron and hole wavefunctions can exceed the nanostructure’s physical size. This widens the band gap, relieves some of the quantum selection rules governing the recombination processes, and improves the eciency of photon processes for the relaxation of the exciton. This is manifested in size-tunable photon energy and increased photolumienscence (PL) eciency. Many studies have been probed the PL and other optical properties of silicon nanocrystals (SiNCs), leading to their application in optoelectronic devices, microelectronics, and solar photovoltaics [5, 26, 32–34, 39, 107, 139, 167]. In the last few decades many advancements on influence of surface conditions on the lumi- In particular, nescence properties of quantum confined SiNCs have emerged [107, 176, 199]. passivating the SiNC surfaces with organic ligands has been demonstrated to improve their PLQY. There are various approaches to this, including solution-phase thermal hydrosilylation, UV- and 8 white-light-initiated hydrosilylation, and gas-phase methods including using furnaces and low- energy plasmas [5, 37, 76, 105, 106, 150, 189]. The wide range of applications for luminescent SiNCs had been limited by the lack of a simple, fast, and inexpensive synthesis approach [107]. Liquid-phase approaches, including synthesis in inverse micelles [187] synthesis in high temperature supercritical solutions,[42] the oxidation of metal silicide,[135] and the reduction of Silicon tetrahalides and other alkylsilicon halides,[8] are mostley time consuming processes. These methods usually have low production rates. Gas-phase methods including the pyrolysis of silane in furnace flow reactors, [94, 107, 127] the decom- position of silane through laser light irradiation (photolysis),[12] and laser pyrolysis using high power infrared lasers [107], all may yield faster processing times. Although these methods have faster production time, are prone to particle agglomeration. In order to deagglomerate the final clusters,further post processing is necessary. These post processing steps might diminish quan- tum confinement properties [107]. However, non-thermal plasma processes as among the most promising, due to the rapid, facile synthesis and tight control over SiNC properties correspond to the easily adjustable reactor parameters such as pressure, reactor dimensions, supplied power, and gas species. SiNCs experience an interesting journey during synthesis in a non-thermal plasma reactor. One of the earliest reports regarding the particle heating in plasmas dealt with micron-sized particles that were injected into argon plasmas [38]. For 0.5-3 µm, they measured particle temperatures that were 75 k above the gas temperature [38]. The propsed model is that the heating of the particles are due to electron-ion recombination and the main cooling event is due to radiation and coduction at the surface of the particles. Despite a low overall gas temperature, SiNCs can reach elevated temperatures,as pointed out by Daugherty et al. and Mangolini et al. [38, 107]. These investigations considered the eects of electron-ion recombination at SiNC surfaces, other reactions between neutral gas species with SiNC surfaces, radiative cooling processes, and gas conduction cooling processes. Results indicate that the SiNC temperatures can spike to several hundred Kelvin higher than the gas temperature. 9 These spikes are responsible for the crystallization of the SiNCs. These works also revealed that bigger nanoparticles require higher temperatures for their crystallization. Significantly, smaller particles will gain higher energy (corresponding to higher spike temperature) due to exothermic recombination events on their surfaces[107]. Kramer et al. [84, 85] studied the dierences between nanoparticle heating and cooling events at atmospheric pressure and low pressure respectively. Since gas conduction is faster in atmospheric pressure, SiNCs produced at atmospheric pressure experienced higher spikes of temperature and fast cooling successively. However, in low pressure environment, cooling is much slower and SiNCs experience higher average temperature after successive heating and cooling events. One of the major attractions of nonthermal plasmas compared to other gas-phase methods is their ability to produce quantum-confined, luminescent SiNCs that are smaller than 5 nm [83]. Studies have shown that decreasing residence time in the reactor produces smaller particles [55, 83]. Synthesis of SiNCs with tunable sizes between 3 nm and 6 nm and residence time of 2 ms and 7 ms, respectively, in a capacitively coupled plasma (CCP) has already been reported more than a decade ago [107]. One important consideration is that the large surface-to-volume ratio for nanoparticles of all types means that ecient luminescence processes require a surface with few defects. For many materials this is accommodated by ligand attachment at the surfaces, a process which is intrinsic to their synthesis in the solution phase. With plasma-produced SiNCs, as-produced hydrogen-coated nanoparticles have a high density of defects and thus a low PL yield. With air exposure, an oxide layer at the shell of these plasma-produced SiNCs enhances the luminescence eciencies. Even more improved luminescence for plasma-produced SiNCs can be achieved via passivation of the SiNCs by surface-functionalization with organic compounds, such as alkene organic ligands. This covalent bonding at the surfaces of the SiNCs helps to achieve solution PL quantum yields (QYs) as high as 60% [79, 105]. Our work on using plasma-produced SiNCs in bilayer and hybrid polymer structures (3) includes surface functionalization using a corresponding procedure, and application of these surface-modified SiNCs. Andthony et al. [7] pointed out that the injection of hydrogen H2 downstream of the synthesis 10 plasma also improves the quantum yield of the resulting SiNCs [83]. The improvement has been confirmed by electron paramagnetic resonance (EPR), with lower surface defects in SiNCs produced using H2 injection. [133]. These prior discoveries built the foundation of our approach to synthesis and surface-modification of SiNCs, as will be discussed in subsequent chapters. 1.3.1 SiNCs on Flexible Substrates: Polydimethylsiloxane (PDMS) In Chapter 3 and Chapter 4, we introduce methods for investigating the SiNCs both on top of and embedded within polydimethysiloxane (PDMS), a commonly-used flexible/stretchable elastomer. We anticipate these structures to find applications in the next generation of gas sensors, solar panel devices beyond traditional architectures, LEDs, and other flexible microelectronics[26]. Previously, luminescent SiNCs have been used exclusively in traditional rigid devices - but flexible, stretchable devices oer exciting opportunities. Recently, the Plasmas and Nanomaterials Laboratory (PNL) performed a study on the aging properties of SiNCs on PDMS [104]. Keeping the synthesis parameters constant, SiNCs were deposited directly out of the plasma reactor into thin-film layers on the PDMS and onto bulk Si wafer substrates and/or glass using inertial impaction through a slit-shaped orifice [3, 5, 68]. This impaction results in a dierent thin film morphology on the rigid substrate as compared to the PDMS. Specifically, we observed the SiNC film to be less dense on the rigid substrate, as observed using electron microscopy [104]. The reason behind this morphological dierence is under investigation, however, is likely attributed to the dierence in elastic properties of the PDMS as compared to the rigid substrate. In chapter 4 we introduce a in situ method for measuring mechanical properties of SiNCs and PDMS as a bilayer system. PDMS belongs to a class of polymeric compounds that are commonly referred to as silicones. PDMS is by far the most widely used organosilicon compound polymer [1]. PDMS also has distinctive flow properties. PDMS is inert, non-toxic and non-flammable. Its applications span a wide range, including in contact lenses and medical devices, additives in beauty products, food- related industries as a foam prevention agent, adhesives, replica molding, micro-nanofluidic and membranes, and more [24]. Moreover, PDMS is a useful functional polymer material in academic 11 experimental research and industrial applications due to its unique properties including reasonable thermal stability, high transparency, low surface free energy and excellent biocompatibility which provides a wide range of other applications for medical device development and health monitoring sensors [24]. PDMS is a viscoelastic polymer. Its viscosity is a function of temperature. During long flow times and high temperatures, it acts like liquids with high viscosity, however, in low temperature it acts like a rubber-like material. Viscoelasticity is a common form of nonlinear elasticity for a wide range of polymers [46]. The elastic behavior of PDMS mostly depends on the concentration of cross-linking agent used during the curing process. PDMS can conform to various types of surfaces during the curing process. Obviously this feature makes PDMS a tunable material for a wide range of experimental applications including, microelectromechanical systems [111, 145]. Mostly, the cross-linked cured version of PDMS as a rubber-like material, has a solidified form and is extensively used for stretched, bent, compressed and other type of stress-based deformation experiments [185]. Generally, PDMS has a low elastic modulus which enables it to be easily deformed and results in the a rubber-like behavior [78, 96]. The shear modulus of PDMS changes with preparation conditions, and consequently varies between the range of 100 kPa to 3 MPa. This topic will be discussed further in chapter 3 and chapter 4. 1.3.2 Mechanical properties of SiNCs, deposited as thin films To date, nanoparticles have been used for variety of applications. However, there is a noticeable lack of knowledge regarding the mechanical behavior of layers of nanoparticles, especially the layer’s ability to bend, stretch and deform when deposited on a flexible substrate. The mechanical behavior upon deformation of a substrate will cause instabilities in deposited thin film structures, including in layers of nanoscale particles. Additionally, when nanoparticles are mixed within polymers, they can aect the nanocomposite’s mechanical behavior. Combining the mechanical properties of a polymer-based materials such as PDMS, with the tunable optical, electronic, and estimated mechanical properties of SiNCs could lead to the introduction of a new class of highly 12 flexible technologies utilizing multi-layer and composite architectures including wearable sensors and foldable LEDs, among others. [10, 11, 156]. In order to be able to practically design devices with such a new level of flexibility, the mechanical properties of the substrate and thin film layers should be investigated individually and together as a system. Large deformations such as uniaxial compression, equibiaxial compression, pure shear, finite bending, and unbending are attracting interest as methods to estimate the mechanical behavior of materials [14, 40, 143, 144, 153]. In addition, uniaxial tensile testing can reveal information about the mechanical behavior of nanocomposites. Chapter 3, Chapter 4, and Chapter 5 discuss our eorts to apply uniaxial tensile testing, bifurcation analysis, and theoretical modeling frameworks to estimate the mechanical properties of nanocomposites based on SiNCs and of SiNC thin films deposited on PDMS directly from the gas phase. In chapter 3 we discuss a novel technique for measuring the mechanical properties of thin film of nanocrystals, especially SiNCs, due to their vast ranges of application and growing market for future device development, deposited on PDMS as a flexible substrate. This method further can be applied for other nanostructure thin films and flexible substrates. In Chapter 4 we investigate the mechanical properties of a composite structure, SiNC clusters embedded in PDMS for luminescence solar concentrator and other flexible optical devices. In Chapter 5, we introduce a parametric study to investigate the dependence of various parameters, including, thickness of the film, modulus of the substrate and surface functionality of SiNCs on the mechanical property of the thin film in a bilayer system. Later in Chapter 5 we also report our study of another class of surface instability, which arises as a result of unbending of SiNC layers as deposited on an intially pre-stretched PDMS system. In summary, these results shed light on the mechancial behavior of SiNCs, dependent on their surface functionality, both in nanocomposite and in bilayer structures. This work contributes to the future applications of SiNCs in flexible, bendable, and stretchable manifestations. 13 CHAPTER 2 AUNP SYNTHESIS: A PLASMA-BASED APPROACH Portions of this work were published in the journal "Plasma Processes and Polymers" in July 2019 and reproduced here. This article was also featured on the front cover of the July 2019 issue of the journal. 14 Figure 2.1: An overview of gold nanoparticle syhthesis 2.1 Introduction Metallic nanostructures have proven to be among the most flexible nanomaterials owing to the synthetic control of their size, shape, composition, as well as the resulting tunable optical properties[72]. Gold nanoparticles (AuNPs) are among the oldest forms of engineered nanostruc- 15 tures, with medical applications dating back to the early 17th century. AuNPs are broadly attractive across disciplines, and are used for applications such as medical theranostics, Surface Enhanced Raman Spectroscopy (SERS), optoelectronic device enhancements, and as seeding sites for other materials such as silicon nanowires[43, 61, 72, 97, 159, 170, 190]. Most synthesis methods of AuNP production involve wet chemistry-based approaches[20, 35, 49, 54, 73, 101, 121, 171, 197]. In solution-based routes, the AuNPs typically end up solubilized in water or other solvents using surfactants, which is useful for some applications – but this can also increase contamination for many post-processing purposes. To eliminate the possibility of contamination in post-synthesis reactions and applications, there is interest in routes for synthesizing the AuNPs in the gas phase. Pre-existing methods in gas phase include inert gas condensation, spark ignition, plasma sputtering, and hot wire synthesis[15, 81, 117, 132, 137, 161]. Recently, several groups have turned to flow-through plasma synthesis of metallic NPs and nanostructured films[28, 29, 86, 188, 197]. The advantages to these plasma-based methods are a relatively narrow size distribution and continuous synthesis. Here we introduce a new method for making AuNPs based on a radiofrequency (rf) plasma together with a consumable Au electrode for synthesis of monodisperse AuNPs with an average diamter of ⇠4nm. We compare this method with a low pressure hot-wire technique for AuNP synthesis, finding that while the hot-wire method has a higher yield of AuNPs, the plasma-based route produces NPs with a smaller size dispersity. This approach adds to the versatility of plasma reactors for nanomaterials synthesis. 2.2 Experimental Details We used a non-thermal radiofrequency (RF) plasma reactor for AuNP synthesis. Figure2.2 shows a schematic view of this reactor. The plasma is initiated via RF power (13.56 MHz) supplied via a matching network to a copper coil around a quartz tube (2.54 cm O.D. and 20 cm length). The grounded electrode is a pure Au (99.99) wire threaded coaxially within the tube, which is also the source of Au for the AuNPs. We flowed Argon (Ar) gas through the reactor at flowrates between 600-1200 standard cubic centimeters per minute (sccm) and used orifice plates downstream of the 16 Figure 2.2: Photograph and schematic of plasma-based AuNP synthesis. Insets show TEM images of AuNPs. reactor to modulate the pressure between 60-220 Torr. The pressure beneath the orifice was <10 Torr. We varied the power to the external coil between 5-15W (nominal), upon which the gas surrounding the Au wire began to glow. To confirm that the Au wire acts as the ground, we attempted to operate the reactor without the coaxial wire – with no success at initiating a plasma. We placed a TEM grid (Lacey Carbon Film Cu 300 mesh from TedPella®) at the exit of orifice for collecting and analyzing the AuNPs. Figure 1 shows a schematic diagram of our process, and Fig2.1 shows a list of samples and corresponding parameters we explored for the plasma method, including pressure, power, and discharge length. We sought to compare our plasma method with an alternate technique, similar to the previously- reported hot-wire AuNP synthesis technique at atmospheric pressure reported by Boies et al. [15], but operated under reduced pressure of 10 Torr. Our hot-wire setup consisted of a glass tube with a Platinum (Pt) wire electroplated with Au (0.5 mm thickness) and connected to a DC power supply. The electroplated Pt-Au wire was threaded via an electrical vacuum coupler feedthrough into a 17 Table 2.1: Reaction parameters for plasma-based AuNP synthesis sealed glass vacuum reactor tube and connected externally to a DC power supply. The glass reactor tube was a three-way heavy-duty quartz glass (2.54 cm O.D. and 12.7 cm length) which was fused to a flexible stainless steel flange to prevent it from breaking during experiments. Ar was used as a background gas for carrying AuNPs towards the reactor exit, with flowrates between 300-1200 sccm and reactor pressure from 2-30 Torr, depending on gas flowrate and orifice size. DC power was supplied to the Pt-Au wire, resistively heating the gold at the surface and causing evaporation. The gold atoms entered the gas phase and condensed to form AuNPs. Figure 2 shows a schematic view of this reactor. 2.3 Results and Discussion We measured the diameter of the plasma-produced AuNPs with Transmission Electron Mi- croscopy (TEM) and Scanning TEM (STEM). TEM images showed individual AuNPs deposited onto the grid at the exit of the reactor (Fig.2.4a-d). Both bright field and dark field images (Fig2.4) show that the collection grid is covered with individual AuNPs. A higher-magnification image shows lattice spacing, with the [2 0 0] plane identified (Fig.2.4c). We performed Fast Fourier Transform (FFT) analysis of an individual AuNP to measure the crystal lattice spacing (Fig.2.4d). 18 Figure 2.3: Schematic and photograph of hot-wire AuNP synthesis, including inset table of reaction parameters. Figure 2.4: Imaging of AuNPs using transmission electron microscopy (TEM). (a) Bright field and (b) dark field low-magnification images of AuNPs collected on carbon fiber TEM grid; (c) higher-magnification TEM image of single Au NP with [2 0 0] crystallographic plane identified as based on (d) FFT of the same AuNP. We performed STEM combined with Energy Dispersive X-ray Spectroscopy (EDS) to further confirm the presence of Au and to measure possible contamination – which we did not observe 19 Figure 2.5: (a)STEM imaging confirms crystalline AuNPs, with inset showing the scanning line for EDS analysis. (b) EDS confirms the composition of the AuNP. Fig.2.5[188, 190, 193]. The EDS peaks aside from Au are from Copper (Cu) and Carbon (C), both likely from the carbon-coated copper TEM grid we used. We did measure some trace Oxygen (O), which is likely the result of normal atmospheric contamination during sample loading. We used high resolution TEM (HRTEM) to measure the average diameter of the AuNPs. Figure2.6 shows a chart of AuNP diameter (error bars reflect standard deviation) that we found for various reactor conditions as listed in Figure 2.1. The AuNPs have narrow size distribution with average diameter below 5 nm, regardless of the reactor conditions we tested (including pressure, RF power, and reactor volume). While the parameters we tested did not yield a major change in AuNP diameter, the small NP size is within a promising range for use as plasmonic resonators and in other applications[45, 60, 175]. 20 Figure 2.6: AuNP average diameter and standard deviation of various samples from Table.2.1, as measured from TEM images To understand the physical process during AuNP synthesis, we used SEM to image the gold wire before and after the plasma synthesis under the conditions for Sample 2 (900 sccm Ar, 10 W supplied RF power). Figure.2.7 shows the SEM images. Roughness at surface of the Au wire after the plasma indicates physical destruction via the plasma process. 21 Figure 2.7: (a) SEM image of the consumable Au wire surface before plasma exposure. (b) SEM image of the Au wire surface after plasma exposure (conditions as for Sample 2 in Table2.1), showing noticeable roughness. (c) Higher magnification of SEM image of the Au wire after the plasma process. We hypothesize that the cause of AuNP production is related to the increased wire temperature during the plasma process. During the process, the wire glows as though it is heated, and when we increased the power to the plasma we saw eventual melting of the wire. By inserting a thermocouple lead into the plasma during the reaction, we were able to measure the gas temperature within a few seconds of extinguishing the plasma (RF interference prohibited direct measurement during 22 the reaction), with temperatures around 400K. This is far from the melting poing of gold (1337 K), consistent with the nonthermal nature of this plasma. However, resistive heating from the supplied RF as well as electron/ion recombination at the wire could lead to increased wire temperature. Due to the higher pressure of our reaction (100-300 Torr) operated at RF we do not suspect ion sputtering to be a major cause of AuNP formation, since any ions produced are likely to have low energies. While the plasma-based method was promising regarding the AuNP size distribution, the mass yields were low – in fact while collection of NPs on a TEM grid was facile, we were unable to collect enough AuNPs for an accurate weight measurement. For comparison, we also performed AuNP synthesis using the hot-wire method under low-pressure conditions (1-10 Torr), as described in the Experimental Details section. Here we compare these two methods. In the hot wire method, we applied a narrow range of DC power to the Au-Pt wire (10-15 W) and the system pressure was lower than plasma-based method (several Torr versus 100-200 Torr for the plasma-based method). The rate of production was lower for the plasma-based method, as confirmed with the TEM results from each method under the same collection time. The hot-wire method produced AuNPs with a broad size distribution and larger average diameter (see Fig.2.8). By comparison, the plasma-based method yields AuNPs with a smaller average diameter and narrower standard deviation, which is attractive as many applications for these NPs rely on size-dependent properties. 23 Figure 2.8: (a,b) TEM images of AuNPs as synthesized using plasma-based method (Ar: 900 sccm, pressure: 180 Torr) and the hot-wire method (Ar: 900, DC power: 12 W) respectively. (c) Average AuNP diameter and standard deviation for the plasma-based method (circular symbols, Samples 1 and 2 from Fig.2.1) and the hot-wire method (square symbols, Samples 3 and 4 from Fig.2.3) as measured using TEM image analysis. AuNP, gold nanoparticle; DC, direct current; TEM, transmission electron microscopy The lack of size control for the plasma method deserves comment. We explored a wide range of plasma powers and gas flowrates, and did not see a pronounced change in average AuNP diameter. In line with our hypothesis that the plasma process for AuNP synthesis is thermally driven, we did not see any size change as a function of power. At low powers, the wire was not heated suciently to lead to AuNP production. However, the thin wire equilibrated in temperature rapidly – so once we supplied enough power for AuNP synthesis, increasing the power led to the wire melting, not to increased AuNP vapor and size: the available power range was small due to this, and did not 24 allow for size control. We also did not see a flowrate dependence. In other flow-through plasma synthesis reactors, the size of the resulting nanoparticles is dictated by the residence time of the nucleates/particles in the plasma[28, 58, 87, 107, 197]. However, in our process, the pressure and flowrate were not decoupled (due to using orifice-based pressure control) and any increase in flowrate caused also an increase in pressure, reducing the eect of flowrate on the residence time. Furthermore, the long residence times for our process (due to the higher pressure) meant that we were not dynamically altering AuNP formation by changing the gas flowrate. When we tried to reduce the pressure and thus the residence time, we were not able to synthesize AuNPs. We also calculated the residence time (RT) of the AuNPs in the reactor. This calculation is based on Equation 2.1 below. The RT (described below) roughly estimates the time each particle travels in the plasma and thus can give an idea of how long the particle has to grow. This equation accounts for fluid velocity and steady flow conditions through the orifice. This equation shows that pressure and flowrate have a direct eect on a particle’s RT. These results show a loose correlation between the size of the AuNPs and their RT, as we dicussed above in this section. Table 2.2 shows calculated RT for particles characterized in Table 2.1. ResidenceTime(RT) = (((⇡ ⇤ R2)⇤ L ⇤ ((P/Patm))))/((F(sccm)) ⇤ 60) (2.1) Table 2.2: Comparing plasma-based and hotwire methods Despite the success at synthesizing AuNPs using the hot-wire method and some control over size via process parameters, we were discouraged by material costs (namely, the Pt core wire), 25 experimental constraints, and diculties associated with the reactor geometry, as well as the wide size distribution of the AuNPs. The plasma technique is a more direct, lower-cost method for AuNP synthesis and gas-entrainment, although the production yield is also low enough to prompt future studies on how to improve on this method. Here, we presented a gas phase method which is based on an RF plasma to synthesis AuNPs. We also produced AuNPs by hotwire method under vacuum and compared these two methods. For either method, it is critical to understand the conditions during AuNP growth so that the method can be extended to other noble metals, and to aid future processes based on specific AuNP properties (for example, Si nanorod growth based on supersaturation of a eutectic alloy of Au and Si using the AuNPs as seeds). Reaching toward future experiments using these combined gas-phase processes, we performed preliminary experiments to analyze the AuNP temperature during formation, the gas temperature during the plasma process, and employed methods to analyze the gas composition with future use in estimating the electron temperature in the plasma. Those results are summarized in this section. Based on results presented earlier in this Chapter, future experiments on anisotropic semicon- ductor nanostructure growth in the PNL will use the plasma-based AuNP synthesis route. As a preliminary step towards this objective, we performed experiments to determine the exact nature of AuNP production in this method as we briefly discussed earlier [83, 84]. In one experiment, we attempted to able to measure the ion density in the plasma using an electrical probe which is installed in the plasma reactor. This theory has been applied in experiment as well [19, 84, 109]. A function generator is connected to a variable frequency power supply through a capacitor. The probe is powered at RF frequencies (< 13.56 MHz) and chopped – during the “signal-o” phase, the capacitor is discharged by the plasma ion flux to the probe. The rate of discharge is proportional to the ion density (which is equal to the electron density under the assumption of quasineutrality in the plasma) [19, 83, 84, 134]. Our experiments with electrical probe measurements are ongoing, due to complications with the lower-frequency power supply which led to many false starts. While our probe measurements are on hold until we can obtain more robust and reliable equipment, we 26 plan to use this method in the future as a means to estimate the plasma characteristics during AuNP synthesis. We also used Optical Emission Spectroscopy (OES) to investigate dierent excited species and their relative intensities in the plasma system which helps to calculate species concentrations and can give an idea of the electron temperature during the reaction. In OES measurements, a spectrometer collects all emitted light from the plasma (ultraviolet and visible light). The lines are indicative of various atomic transitions from excited-state species in the plasma. The ultimate objective in this study was to then use a computer code to estimate the concentrations of various species and the electron temperature during the process, based on comparing these characteristic line intensities. Figure 2.9 shows a sample of OES from the Ar plasma with the gold central ground electrode, including a trace addition of H2 gas for use in calculating electron temperature[19, 84, 109]. The data in fact was inconclusive. We hoped to see a line from Au, if in fact we had generated any excited-state Au atoms during our process. However, either the process was not energetic enough to excite the Au atoms/clusters that coagulated during AuNP synthesis, or else the densities were too low to observe. Characteristic lines from Au are expected at 406.511 nm, 460.750 nm, 479.273 nm, 627.817 nm and 751.073 nm (from NIST Atomic Spectra Database Lines Data website). Additionally, we hoped to use the trace H2 that we added to the plasma as a means to estimate the electron temperature, but we were also not able to clearly see evidence of the H2 optical emission, making that calculation impossible. It is interesting that the two spectra with and without H2 (red and blue, respectively) are qualitatively dierent from one another - and so the addition of H2 even at low concentration (1 sccm) was sucient to shift the emission from the Ar plasma. In Figure 2.9 we used a 340-1050 nm Ocean Optics USB spectrometer to run the AuNPs plasma reactor with using H2. 2.4 Conclusion We presented a new gas phase method which is based on an RF plasma to produce AuNPs in a more sustainable path. For comparison, we also produced AuNPs by hot wire method at low 27 Figure 2.9: OES test for Plasma-Based AuNP synthesis, 340-1050 nm spectrometer with Hydro- gen(red) and without Hydrogen(blue) pressure. This new plasma-based approach results in a smaller average size and smaller overall distribution of sizes per sample, as compared to the hot wire method. Among the challenges presented by this method are a low production yield and invariant particle diameter, based on the parameters we tested, although the average AuNP diameter of 4 nm is suitable for many applications. Future work will explore methods to increase the production.yield as well as control the AuNP diameter using the plasma method, along with eorts to employ this facile technique to synthesis of other metallic NPs. 28 ESTIMATING THE MECHANICAL PROPERTIES OF SINC LAYERS ON PDMS CHAPTER 3 Portions of this work have been published in the journal "Nanoscale" in March 2019 and are reproduced here. 29 3.1 Introduction As flexible devices become more prevalent, it is increasingly important to measure and predict the mechanical properties of active device layers, including thin films [9–11]. Nanoscale com- ponents, such as nanocrystals, can be incorporated into mechanically flexible devices.[10, 152] Silicon in its nanocrystal form has semiconducting and photoluminescence properties which make it a promising material for technologies like microelectronics[103, 140, 163], photonics[50, 103], solar-photovoltaics[95, 166], memory devices[146], biosensors, and light emitting devices[6, 7], as well as other applications involving polymer matrices.[115, 173] Using silicon nanocrystals in these applications is particularly attractive considering the natural abundance and non-toxicity of silicon. In most of the aforementioned applications, silicon nanocrystals (SiNCs) are deposited on rigid substrates. Depositing a SiNC film onto stretchable or flexible substrates opens up possibilities for new device development and modifications[179]. While the opportunity of designing these devices is exciting, it also highlights the need for an innovative way to estimate mechanical properties for these systems. Substantial work has been done by various groups on studying the mechanical properties of nanocrystalline materials, both in the elastic [57, 191] and plastic [62, 172, 180, 181, 184] regimes. Some works also discuss the significant dierences in mechanical properties of nanocrystalline material from their bulk counterparts [57, 62, 172, 181, 182, 184, 191, 196]. These testing techniques, while powerful in estimating the material properties for a large variety of nanocrystalline materials, inherently require the destruction of the samples or the plastic deformation of the material. This is in conflict with the scope of this work, which is focusing on non-destructive techniques to estimate elastic behavior of thin layers of nanocrystals on deformable substrates. These combined systems limit in-situ measurement of the layers’ mechanical properties such as elastic modulus. This necessitates finding an appropriate method to understand and predict the response of not just the nanocrystal layers during deformation, but of the whole elastomer-nanocrystal layer system. Indeed, in-situ evaluation of mechanical behavior is crucial for thin layers, for which interactions between the substrate and the thin film can be as important as the properties of the thin film itself. To 30 our knowledge, such in-situ evaluation of the mechanical properties of SiNC layers on elastomers has not been previously reported. Prior in situ studies on nanoscale materials have employed microscopy during uniaxial tensile testing to observe deformation under applied loads [181, 183, 184]. Barring the specialty equipment necessary for these experiments, a promising option that has been used in the last two decades to estimate material properties of thin films is to measure the onset of instability patterns on the surface of the thin film due to dierent loading conditions [30]. This technique has the advantage of being simple, inexpensive, and non-destructive; furthermore it can be applied to any type of thin layer deposited on a deformable substrate. Because of its flexibility, this method has been employed to estimate mechanical properties of several isolated thin films [157, 168, 178] or multi- layered structures [77, 102, 125, 162]. All of these studies focused on applying homogeneous deformations, namely uniaxial tension or compression, to the system. However, other types of non- homogeneous loading conditions, such as bending, can also lead to the formation of instabilities. In the past, several studies have focused on developing a theoretical framework to describe the onset of instabilities on the surface of rectangular blocks subject to finite bending[31, 40, 41, 44, 51, 52, 143, 144, 153, 169]. The framework developed in these works focused only on finite bending deformation for a homogeneous block. These early predictions have also been recently confirmed by experimental work [51, 52]. In the last decade, a few studies focused on predicting bifurcations during finite bending of multi-laminated structures, specifically in bilayered systems, both theoretically and experimentally[143, 144]. A recent study on the optical and structural properties of SiNCs on polydimethylsiloxane (PDMS) performed at PNL and demonstrated the formation of instabilities on a nanocrystal film deposited on a prestretched elastomer [104]. Combining the onset of instabilities with the predic- tions from the theoretical modeling framework described, we evaluate the mechanical properties of the SiNC thin film deposited on an elastomeric substrate. This theoretical framework can further be applied to dierent combinations of nanocrystals and elastomers, for in situ measurement of mechanical properties of such bilayered systems. 31 Figure 3.1: Steps from plasma deposition to finite bending. SiNCs are deposited through a slit- shaped orifice onto PDMS, which is then bent to observe formation of wrinkles on the SiNC surface. (a) Photograph of plasma reactor. (b) Schematic diagram of reactor and SiNC synthesis. (c) SiNC/PDMS sample illuminated with UV light, demonstrating photoluminescence. (d) Finite bending applied with a lab-designed apparatus. 3.2 Methods Here we take a novel approach to evaluate the mechanical properties of thin films of nanocrystals on an elastomeric substrate. A PDMS-SiNC bilayer was used as a model to present the framework. First, we inertially impacted a uniform film of SiNCs on PDMS. Then, we applied a finite bending deformation to the system until instabilities formed on the surface coated with the SiNCs, noting the critical angle at which the instabilities appeared. Finally, we applied a theoretical modeling framework to describe the bilayer system. See Figure. 3.1 for an overview of our process. 32 3.2.1 PDMS fabrication To prepare the PDMS as substrates for this study, a Dow Corning heat-cured prepolymer PDMS 184-kit Sylgard Elastomer with curing agent catalyst was used with a 10:1 weight ratio of base to curing agent catalyst. Changing this mass ratio will result in PDMS substrates in dierent stiness. In other words, curing agent ratio will determine the mechanical properties such as shear modulus of the PDMS polymer after curing process. This topic will be discusssed in Chapter 5. This mixture was stirred vigorously with a mechanical stirrer until air bubbles formed, and placed under vacuum for 10-15 minutes to remove the trapped air. The mixture was then weighed and poured into a petri dish of 85 mm diameter and was cured for 120 minutes at 60oC by placing on a hot plate to form PDMS with an average thickness 2.5 mm. When PDMS cured, the thickness at the edge of the container is alittle bit higher, due to surface tenssion of fluid during the curing process. After curing, the PDMS was cut in rectangular sections according to our requirements and then attached to two metallic handles with Gorilla® Super Glue as shown in Figure.3.1b and d. 3.2.2 Plasma synthesis of SiNCs We produced the SiNCs using a nonthermal plasma reactor as has been reported previously[76, 80, 83, 104, 107]. This reactor, as presented in Figure 3.1,consists of a pyrex tube through which we flow 94 sccm (standard cubic centimeters per minute) of Argon (Ar) and 16 sccm of silane (SiH4, 5% in Ar), as well as 50 sccm of H2 through the sidearm for surface passivation. Radiofrequency (RF) power at 13.56 MHz was supplied to dual ring electrodes encircling the reactor, and the pressure was kept constant in the range of 3.75 to 3.85 Torr using a slit-shaped orifice. Obviously changing the size of the orifice, will change the operating pressure of this reactor. The stando distance between the orifice and the substrate was ⇠4 mm. In this study, we keep the stando distance constant to presevre same impaction for SiNCs deposited on the flexible substrate. The downstream pressure was ⇠0.500 Torr. We also mesured the pressure in the middle of the reactor, using a side arm in the same height as for the side arm desgined for Hydrogen line. This middle pressure was ⇠1.250 Torr. The nanocrystals were deposited on thin films by inertial impaction directly out of the reactor onto 33 the substrate[6, 48, 69, 124]. One of the assets of this synthesis method is the room-temperature process and ability to directly deposit SiNCs without post-processing steps or solvents, even on temperature-sensitive substrates such as PDMS. The PDMS samples were made (as described in Sec. 3.2.1), attached to a pushrod, and placed under the orifice downstream of the reactor. The SiNCs were deposited by rastering the substrate back and forth beneath the orifice to achieve a uniform film of SiNCs on the PDMS. Figure 3.1b shows a work-flow schematic of the reactor and deposition of SiNCs on the PDMS. The rastering duration was 20 minutes with each back and forth movement lasting 3 seconds, resulting in a SiNC layer thickness of ⇠4.5 µm. 3.2.3 Silicon nanocrystal film characterization The thickness of the sample was measured by scanning electron microscope (SEM) imaging of a cross-section of the sample generated using a focused ion beam (FIB). Figure.3.2a shows an SEM image of the sample, and Figure.3.2b shows the FIB-generated cross section for thickness estimation. To decrease charging eects and improve the image quality of the SiNC film, a thin layer of Platinum (Pt) (⇠6 nm) was deposited via Ar plasma sputtering prior to imaging.To perform a reasonable FIB cross sectioning, we need to roughly estimate the thickness of the film prior to milling and final imaging. As we mentioned in previous section, we started our FIB milling with an estimation of 4.5 µm. The speed of FIB cross sectioning depends on the thin film structure and its thickness. The FIB image presented in Figure 3.2 took around 10 mins. Another challenge in this thickness measurement was due to similarity of contrast of PDMS and SiNCs. In other word, due to the existing Silicon in the structure of PDMS and SINCs, the contrast at the interface of PDMS, as the substrate and SiNC layers, as the thin film, is negligible and need high resolutiom for confirming proper estimation of the interface as presented in Figure 3.2. X-ray diraction (XRD), shown in Figure.3.3a, confirms the formation of crystalline nanoparticles. Transmission electron microscopy (TEM) was performed on the SiNCs using a JEOL 2200 instrument, showing individual NCs (Figure.3.3b). The TEM sample was prepared by tapping the TEM grid on the surface of the sample to collect some of the SiNCs. The size of the SiNCs was measured using 34 TEM image analysis, giving an average SiNC diameter of ⇠3.7 nm. Figure 3.2: SEM and FIB cross-sectioning of the SiNC layer. The thickness is estimated to be (⇠4.50 µm). (a) SEM image showing the surface of the SiNC layer on PDMS, including the FIB- milled area (inset). (b) Higher magnification SEM image of the FIB-milled area demonstrating the thickness of the SiNC layer. 35 Figure 3.3: (a) X-ray diraction confirmed formation of SiNCs. XRD for SiNCs shows crystal reflections at 2✓ values 28.3, 47.3 and 56.1 degrees consistent with the (111), (220), and (311) planes of silicon. (b) TEM image confirming the formation of SiNCs: black circles highlight individual SiNCs. The inset shows FFT of the correspondig SiNCs of the TEM image. 3.2.4 Uniaxial tensile test We applied a standard uniaxial tensile test to evaluate the mechanical properties of the PDMS. A rectangular sample of PDMS was mounted on a custom built uniaxial tensile test machine and four black dots were applied on it as fiducial markers. Figure 3.4 shows a PDMS mounted to the load cell for performing unaxial tensile test. The sample was then stretched cyclically for 5 cycles from a configuration with a 10g pre-load ( reference configuration) until a 20% deformation was achieved. The sample was stretched between two clamps, one of which was attached to a load cell to measure the force on line with the applied deformation. The strain rate used during the tensile test was 0.1 mm/s. The force from the load cell was converted to engineering stress by calculating the average cross-section area of the sample in the reference configuration. A Hitachi KP-M2AN CCD camera was used to capture and feed the image of the fiducial markers during uniaxial stretch into LabVIEW. 36 Figure 3.4: The stress machine loaded with a colored PDMS and dots depicted on the PDMS, dislocation of dots will be recorded by a digital camera Figure 3.5: (a and b)A schematic diagram of proposed aparatus for finite bending measurement desgned by Solid Work software and (c) Alluminum cast aparatus designed at PNL 3.2.5 Finite bending experiment We designed a finite bending apparatus similar to the one used previously by Roccabianca et al.(2010)[143]. A schematic diagram of this aparatus shown in Figure 3.5. Two mainpa- rameters in this design are slenderness ratio (⇤ = l0/h0) which is the ratio of width to thickness of the sample in cross section, and thickness ratio (H = hNC 0 ) which is the ratio of thickness of the SiNC layer by that of the PDMS before finite bending (see Fig. 3.6a for reference). The third dimension is taken to be 8-10 times larger than l0 to achieve a plane-strain condition on the cross section.In other words, due to the size ratio of third dimension to the other two dimensions of the substrate,the third dimension does not have an impact on the formation of surface instabilities. Here (and in all following equations) the superscript ‘P’ refers to the PDMS and ‘NC’ refers to the SiNCs. 0 /hP 37 Figure 3.6: PDMS sample in the finite bending apparatus. (a) Side view, showing sample (with handles) in the finite bending apparatus. (b) Top view showing formation of bifurcations. The bilayer PDMS/SiNC system was put in the finite bending apparatus (see Fig.3.6) and progressively bent until the SiNC film underwent bifurcation, at which point the critical angle was recorded. Due to higher critical angles of bifurcation observed for samples with higher slenderness ratios, deformation was applied by hand instead of using the apparatus. The onset of bifurcations was detected by direct visual inspection, and the critical angle was measured using analysis of photographs of the side view of the sample. Each measurement was performed three times to ensure accurate and reproducible readings. 3.2.6 Modeling bifurcation in elastic layered structures The following sections pertaining to theoretical modeling of the behavior of the bilayer system were largely executed by M. Sinha and Dr. S. Roccabianca, our collaborators on this project. For completeness they are included here. Roccabianca et al. (2010), formulated a theoretical framework to study incremental bifrucations due to finite bending of an elastic incompressible multilayered block[143]. This theory was developed for an N-layered system for Mooney-Rivlin type of material. The authors also performed some experiments in bilayered structures that confirmed their theoretical predictions. Here, we employ a similar theoretical modeling framework for the bilayered system to estimate the ratio of mechanical properties of SiNC thin film to that of PDMS. 38 3.2.6.1 Finite bending The finite bending deformation is prescribed such that the reference configuration is rectangular and described in a Cartesian coordinate system, and the deformed configuration is a sector of ring described in a cylindrical coordinate system (see Fig. 3.7). Figure 3.7: Reference and deformed configurations for the PDMS-SiNC bilayer. The bilayer de- forms from the Cartesian coordinate system in the reference configuration (e1, e2, e3) to a cylindrical coordinate system in the deformed configuration (er, e✓, ez). Specifically, a plane at constant x0 1 transforms to a plane at constant r (dashed line), and a plane at constant x0 2 transforms to a circular arc at constant ✓ (dashed-dotted line). Since the out of plane deformation is taken to be zero, we have x0 3 = z. Briefly, the finite bending deformation can be described, within both the PDMS and the SiNC layers, by the following deformation gradient F[143] F = l0 2 ¯✓r er ⌦ e0 1 + 2 ¯✓r l0 39 e✓ ⌦ e0 2 + ez ⌦ e0 3, (3.1) where l0 is the overall width of the sample and 2 ¯✓ is the angular deformation of the sample. We consider the volume of the system to be conserved and the two layers to be perfectly bonded. Therefore, we can apply the interface condition r NC i are the outer radius of the SiNCs layer and the inner radius of the PDMS, respectively. This allows us to write, i , where r NC o o = rP and rP r NC o i hP 0 , hNC,↵ = rP hNC 0 , hP,↵, (3.2) where hm 0 and hm represent the thickness of each layer before and after bending (m = NC, P), and ↵ = 2 ¯✓/l0. To predict all geometrical characteristics of the deformed state, we consider all the layers to be equilibrated. Briefly, we define the Cauchy stress in each layer as, T = ⇡I + F dW dC FT, (3.3) where C = FTF and I are the right Cauchy-Green deformation tensor and the identity tensor, respectively, and W and ⇡ are the strain energy function and the Lagrange multiplier for each layer, respectively. We have employed the neo-Hookean strain energy function to describe both layers, µ0 defined as W = 2 (trC 3). Note that we are defining two constitutive equations, one for the PDMS layer as a function of the material parameter µP 0 , and one for the SiNC layer as a function of the material parameter µNC . 0 For the layers to be equilibrated, the equilibrium equation in absence of body forces is required to be satisfied within each layer. Furthermore, boundary conditions of zero traction are applied at the innermost and outermost surfaces, as well as an interface condition of perfect bonding between the two layers. 3.2.6.2 Bifurcation analysis The loss of uniqueness of the plane-strain incremental boundary value problem has been investi- gated for multi-layered structures in previously published work[143]. Here we have adapted the formulation to fit our application for a bilayered system of PDMS and SiNCs. Briefly, for each layer, we seek to solve the incremental counterpart of the equilibrium, div(⌃) = 0, where ⌃ = €SFT is the 40 incremental first Piola-Kirchho stress tensor, and the first Piola-Kirchho stress and the Cauchy stress tensors are related by S = T F-T for incompressible materials. Here, and in the following, we use a superimposed dot to represent incremental quantities. The linearized constitutive equation for each layer is given by ⌃ = €⇡I + CL (3.4) where C is the fourth-order tensor of instantaneous elastic moduli and L = grad u is the gradient of incremental displacement u(x). We consider bifurcations to be represented within each layer by an incremental displacement field in a separable-variables form, as described by Roccabianca et al. (2010). For incompressible isotropic elastic materials, one can write C as a function of two incremental moduli, namely µ and µ⇤. For neo-Hookean materials subject to finite bending in plane strain condition, the elastic moduli can be written as µ = µ⇤ = µ0 2 h(↵r)2 + (↵r)2i (3.5) Finally, the incremental boundary value problem is completed by a set of boundary and interface conditions. Specifically, zero normal and shear incremental stresses at the innermost and outermost layer, zero incremental shear stress and incremental radial displacement at the boundary ✓ = ± ¯✓, and continuity of incremental stresses and displacements across the interface. To calculate the angle that satisfies the bifurcation condition, ✓cr, we employed the compound matrix method, as described in a previously published study [144]. 3.3 Results and Discussion Experimentally, the critical angles of bifurcation were measured for nine samples by performing the finite bending experiment described in Sec. 3.2.5. The values, reported in Table 3.1, have been obtained for three experimental groups, characterized by a slenderness ratio ⇤ = 2, 3 and 5, each group formed by a total of three samples (see Fig. 3.8). 41 Figure 3.8: Critical angles of bifurcation evaluated as a function of ratio of material properties for a PDMS-SiNC bilayer sample. The ratio of the neo-Hookean material parameters is defined as ¯µ0 = µNC 0 . All curves have been calculated for a thickness ratio H included in the range 0.0013–0.0018. Each line represent a dierent value of slenderness ratio, ⇤ = 2, black solid line; ⇤ = 3, black dashed line; ⇤ = 4, dark gray solid line; ⇤ = 5, dark gray dashed line; ⇤ = 6, light gray solid line. The insert represents a schematic of the bilayer and defines the slenderness ratio. 0 /µP Slenderness ratio, ⇤ Critical angle, ✓cr Material parameters ratio, ¯µ0 2 3 5 47.33 ± 3.21 71.33 ± 1.53 130.67 ± 3.21 1.90 ± 0.13 1.85 ± 0.05 1.68 ± 0.04 Table 3.1: Critical angles of bifurcation for dierent slenderness ratios, expressed in degrees. Angles are reported for ⇤ = 2, 3 and 5. Employing the model described in Sec. 3.2.6, we evaluated the critical angle of bifurcation for the bilayered structure, shown in Fig. 3.8 and Fig. 3.9. Specifically, after fixing the overall geometry of the sample, namely the thickness ratio H and the slenderness ratio ⇤, we calculated the value of the angle representing the onset of bifurcation for each value of the neo-Hookean material parameters ratio, defined as ¯µ0 = µNC In both Figures, we considered ¯µ0 to be included 0 /µP 0 . 42 Figure 3.9: Plotting points from the experiments on graph obtained by numerical calculations. The ratio of the neo-Hookean material parameters is defined as ¯µ0 = µNC 0 . Shown here are points obtained for ⇤ = 2, 3 and 5. Insert shows an instability on the surface of a sample with ⇤ = 2. 0 /µP between 0 and 100, however a pilot study showed that the value of the critical angles plateaus for values of ¯µ0 included between 102 and 105 (data not shown). Furthermore, the same pilot study showed that the critical angle of bifurcation is independent of the thickness ratio H, if H is included in the range 0.0013 – 0.0018. The experimental value measured in this study is included within that range, specifically H = 0.0017 on average for the nine samples. Figure 3.8 shows the variation of the critical angle of bifurcation for varying values of the slenderness ratio, specifically for ⇤ included in the range 2 to 6. For a fixed value of ¯µ0, which is represented by a vertical line in the graph, an increase in the slenderness ratio results in a later appearance of the bifurcation, namely a higher value for the critical angle. This trend is consistent with previously published results[143, 144]. Figure3.9 shows how we employed the onset of bifurcation to estimate the mechanical properties 43 of the SiNC layer. Briefly, for each value of slenderness ratio considered in the experiments, we calculated the correspondent theoretical sets of bifurcation angles. Then, the critical angle measured experimentally was employed to uniquely identify the corresponding ¯µ0. The experimental value of the critical angle fixed the abscissa of the point in the ✓cr – ¯µ0 plane, represented in Fig. 3.9. Then, we imposed that the experimental point belonged on the bifurcation threshold curve calculated for the corresponding slenderness ratio, which in turns fixed the ordinate of the point which is ¯µ0. The value estimated for each experimental group, namely for each value of slenderness ratio considered, is reported in Table 3.1. A vertical dashed line in Fig. 3.9 represents the average value of ¯µ0 for all the experimental data, which is 1.81 ± 0.12. We then estimated the value of µP 0 = 190 kPa by fitting the data collected from the uniaxial tensile test, as described in Sec. 3.2.4. This gave us a value for the neo-Hookean material coecient for the SiNC layer of µNC 0 345 ± 23 kPa. The neo-Hookean material constant is analogous to the shear modulus for low values of applied strain. This finding is a clear indicator that the mechanical properties of this materials system are determined by the interactions between SiNCs within the NC layer and between the NC layer and the PDMS. Direct measurements of the mechanical properties of individual silicon nanocrystals via microscope-guided nanoindentation have yielded elastic moduli around 170 GPa,[53] several orders of magnitude higher than our measurement of the shear modulus for the SiNC layer on PDMS (assuming incompressible behavior, shear modulus is ⇠1/3 of the elastic modulus). Others have produced elastic/Young’s modulus estimates for single SiNCs between 70-170 GPa.[82, 119, 149] This range of values (⇠20-60 GPa for shear modulus) is also consistent with mechanical measurements of thin ribbons of crystalline silicon on PDMS [82]. By contrast, our value of 345 kPa for shear modulus is far lower, and is consistent with the higher porosity of our nanopowder layer as compared to bulk silicon. For example, Gross et al. showed that increasing porosity of powder hydroxyapatite coatings leads to decreasing elastic modulus - further, regardless of porosity, the coatings exhibited values of elastic modulus more than 35x smaller than the modulus of single- crystal hydroxyapatite. [59, 192] Similarly, the Young’s modulus of aluminum foams is between 101 and 103 the value of Young’s modulus of solid aluminum, depending on porosity.[2] Finally, 44 previous results have shown that increased porosity significantly decreases the Young’s modulus of polymer thin films deposited on PDMS, when measured by employing mechanical instabilities[155]. The porosity of our layers was dicult to ascertain with high accuracy. However, a preliminary density measurement provided us with a value near the limit for close-packed randomly oriented rigid spheres (volume fraction 0.64). While we did not probe the eects of porosity in this proof-of-concept investigation, the porosity of the SiNC layer is tunable by adjusting reactor properties,[48, 69] and future work will be devoted to determining the relationship between layer porosity and mechanical properties. This versatile method is broadly applicable to extract the mechanical properties of nanoparticle layers on elastomer substrates regardless of material, layer thickness, and density. Our results also oer a unique opportunity in the future to investigate how the NC surface functionality (organic ligand coating, inorganic shell, etc.) influences the mechanical behavior of the NC films. 3.4 Conclusion In conclusion, we have used a new method based on the onset of bifurcation to estimate the mechanical properties of thin layers of nanocrystals on PDMS by coupling the experiment and theoretical modeling. In general this method, opening up new possibilities for estimating, and controlling the mechanical behavior of nanoparticle layers on stretchable or bendable device technologies, leveraging both the unique optoelectronic properties of nanoparticles together with the mechanical versatility oered by these films. Finally, the results obtained from this experiment are exciting since previous attempts in in- vestigating instability formation due to finite bending in a multilayered structure were limited to structures with elastomeric layers only. Here we have reliably shown formation of instabilities in thin SiNC films deposited on PDMS as a flexible substrate. This brings opportunities to study and develop constitutive models which are better suited to such constituents, in addition to other experimental methods that can perform non-destructive and in-situ measurements of this class of multilayer systems. This versatile method is broadly applicable to extract the mechanical proper- 45 ties of nanoparticle layers on elastomer substrates. In Chapter 5 we apply same methodology to investigate a parametric study regarding the influence of a few properties of the SiNC layer and the PDMS and how they aect the mechanical properties of the SiNC thin films. Also this method, regardless of material, layer thickness, and density can be applied for a range of nanomaterial films and substrates. 46 OPTICAL AND MECHANICAL BEHAVIOR OF SINC-EMBEDDED PDMS SLABS CHAPTER 4 47 4.1 Introduction Luminescent nanocrystal inclusions in polymer hosts have received considerable interest re- cently, in particular due to their promise in novel optoelectronic device applications such as pho- tovoltaics based on luminescent solar concentrators (LSCs), electrochromic windows, and more [23, 65, 89, 115, 116, 141, 165, 174, 186]. The combination elastomeric behavior of polymers together with the tunable optical absorption and emission from these nanocomposites also opens avenues to stretchable and flexible manifestations of these same applications, as well as wearable electronics such as biosensors for health monitoring systems, and more. While many groups have studied the optical behavior of these nanocomposites, their mechanical behavior is not well understood, although a few recent studies have made advances in uncovering the eects of the nanocrystal inclusions [23, 66]. Modifications to the mechanical behavior of the polymers upon inclusion of the nanocrystals is especially important for elastomeric polymers with intended use in flexible or stretchable applications. The inclusions can change the overall mechanical properties of the composites, depending on a range of factors such as the nanocrystal concentration, dispersion within the host polymer, the nanocrystal surface properties, and more. Going beyond informing the deformability of the composites, tests on the mechanical properties of the composites can also provide information about the interaction between the inclusions and the polymer, and the ways in which applied stretching or bending can aect the other physical properties of each. A particularly compelling materials system is silicon nanocrystals included in polydimethyl- siloxane (PDMS). Silicon nanocrystals (SiNCs) are made from an abundant and nontoxic material, and exhibit exciting tunable optical properties such as photoluminescence[3, 5, 27, 36, 107, 113, 114, 154, 160, 194]. PDMS is a commonly-used silicone-based elastomer used in biomedical, industrial, microfluidic, and optoelectronic applications [27, 88, 99, 110, 112, 123, 131, 148, 151, 177, 195]. As a host for nanocrystals, PDMS oers easy fabrication, tunable mechanical behavior, biocom- patibility, and high translucency in ultraviolet through near infrared spectral range [93]. Recently SiNCs embedded in polymers such as polymethyl(methacrylate) (PMMA) and PDMS have been proposed as agents for LSCs, solar cells, UV-blocking layers, and more due to the NCs’ absorbance 48 in the ultraviolet (UV) – blue spectral range and ecient emission at red-near infrared (NIR) wave- lengths. Understanding the influence of the SiNCs on the mechanical behavior of the composite NC/polymer system as a whole is important for designing future applications based on this materials system, as well as other NC/polymer combinations. A recent report by Chen et al.indicated a modified rupture stretch and mechanical properties (tensile strength and Young’s modulus) for SiNC-infused PDMS, depending on the surface func- tionality of the SiNCs. That group also demonstrated the UV-blocking ability of the SiNC/PDMS composite system, highlighting the multifunctionality of these materials. We sought to further characterize the influence of surface termination on the mechanical properties of SiNC/PDMS com- posites by using uniaxial tensile testing to evaluate the neo-Hookean coecient of the composite system, depending on both nanoparticle concentration within the nanocomposite and nanoparticle surface functionality as modified using a solution-phase thermal hydrosilylation technique. The neo-Hookean coecient is analogous to the shear modulus of a material, at low values of applied strain. Our results indicate that, while all nanocomposites exhibited photoluminescence (PL), the SiNC concentration and surface functionality play a profound role in the mechanical properties of the composite. 4.2 Experimental Work 4.2.1 Plasma synthesis of SiNCs SiNCs were synthesized with a non-thermal low-pressure plasma reactor as described previously [7, 13, 22, 56, 104, 107, 118, 120, 126]. The reactor consists of a pyrex tube with dimensions of 0.25” OD at the reaction zone, expanding to 1” OD in the plasma euent. Dual ring electrodes bearing radiofrequency (RF) power at 13.56 MHz encircled the tube, with power delivered at 30W via a lab-built impedance matching network. Argon (Ar, 96 standard cubic centimeters per minute, sccm) and silane (SiH4, 5% in Ar, 16 sccm) are flown through the reactor at a pressure of 3.6-3.8 Torr as controlled using an orifice near the exhaust. Hydrogen gas (H2, 50 sccm) is injected just below the onset of the expansion region in order to provide more thorough surface passivation 49 [7]. SiNCs are nucleated and grown during the reaction and collected at the reactor exhaust onto stainless steel mesh filters. The filters were weighed before and after the reaction to measure the weight of SiNCs collected, which was on average 3.2mg for the 12 minutes collection time used for each batch. A section of the mesh filter was removed for FTIR analysis, before the remaining mesh filter loaded with SiNCs was placed into a test tube, capped with a rubber septum, and flushed with N2 to ensure an inert atmosphere. 4.2.2 Functionalization of SiNCs Thermal hydrosilylation was performed for all F samples using a nitrogen-purged Schlenk line [106]. The SiNCs were mixed with a dried and degassed solution of 5:1 v/v mesitylene and 1-dodecene (Aldrich) and sonicated for 5 minutes. The resulting cloudy yellow solution was transferred via cannula to a round-bottom flask with a refluxer and heated in a sandbath heater to 205-210°C for 2-hours. Following this process, the SiNC solution became clear indicating surface stabilization with the dodecyl ligands. The resulting solution was dried under N2/vacuum for 24 hours at a temperature of 60°C to evaporate the residual solvent mixture. 12 mL hexane was added to the round bottom flask and sonicated for 5 minutes to ensure proper redispersion. Next, 2 mL of the solution was placed in a vial for PL spectroscopy and the remaining 10 mL was ready for nanocomposite formation. 4.2.3 PDMS preparation We made the PDMS using a previously documented method [110]. A Dow Corning heat-cured 184 Sylgard Elastomer PDMS kit was used with a 10:1 w/w ratio of base to curing agent. The SiNCs with known mass and concentration in hexane were added to the PDMS pre-polymer solution. The mixture was stirred vigorously and then placed in a vacuum desiccator to remove any air from the mixture and most of the hexane evaporated during this process. The resulting composite mixture was poured into an aluminum mold with a rectangular shape and defined dimensions. The 50 PDMS-SiNCs composite was cured on a hot plate at 60 °C, typically reaching final cure in 3 hours. Once the samples were cured they removed using tweezers. 4.2.4 Uniaxial tensile testing We performed a standard uniaxial tensile test as reported previously to evaluate the mechanical properties of the plain as well as the nanocrystal infused PDMS [154] A rectangular sample was mounted on a custom built uniaxial tensile test machine and four black dots were applied on it as fiducial markers. The sample was then stretched cyclically for 5 cycles from a reference configuration, until a stretch of 1.2 was achieved [154]. This stretch is equivalent to a tensile strain of 20%. The sample was stretched between two clamps, one of which was attached to a force transducer to measure the force on line with the applied deformation; simultaneously, the deformation was measure by tracking the fiducial markers using a CCD camera (Hitachi KP- M2AN). The strain rate used during the tensile test was 1 mm/s. The force was converted to Cauchy stress by evaluating the average cross-section area of the sample in the reference configuration, and the change in cross section for the deformed configuration. The samples were then stretched in the same setup at a strain rate of 1 mm/s until rupture. 4.3 Results 4.3.1 SiNC characterization and nanocomposite formation SiNCs were produced using a nonthermal plasma technique, as reported previously and describeed in the Experimental Details section below [5, 104, 107, 154]. We characterized the SiNCs using x- ray diraction (XRD) and transmission electron microscopy (TEM) (Fig. 4.1). The average size of a single SiNC was estimated using TEM image analysis to be⇠4 nm in diameter and confirmed to be around⇠3.7 nm using Scherrer broadening analysis of the XRD peaks [56, 154]. In formation of the nanocomposites, we considered two separate surface treatments: non-functionalized SiNCs (NF, which have a native hydride coating from the plasma synthesis process), and functionalized SiNCs (F, which have alkylated surfaces following a thermal hydrosilylation process with 1-dodecene as 51 Figure 4.1: SiNC Characterization: XRD (a) shows crystallographic reflections for nanosized silicon. Gray lines are the actual data and the black trace is the peak fit. TEM imaging (b) indicates spherical nanocrystals and lattice spacing consistent with crystalline silicon. Inset shows FFT of image. functionalizing ligand and then dried). Photographs of the NF and F samples are shown in Figure 4.2a-b under room light and ultraviolet (UV) excitation, demonstrating visible red luminescence. PL spectra from solutions of NF SiNCs and F SiNCs are presented in Figure 4.2. Solution PL quantum yields from these samples were 1.6% and 54%, respectively, as measured using an Ocean Optics, Inc. USB2000 spectrometer and integrating sphere and excited using a 395nm LED excitation source. The NF and F samples were mixed into a small volume of hexane (12 mL, of which 10 mL was used for fabricating the nanocomposites) and then added to the PDMS pre-polymer solution and the curing agent – 1:10 w/w ratio for all samples – then vigorously mixed using a mechanical rotary stirrer for 1 minute. Air bubbles and residual hexane were removed using vacuum desiccation and the PDMS was poured into casts and cured at 60°C (more details in Section 4). We examined two dierent concentrations of SiNCs/PDMS – 0.03 wt% and 0.12 wt%. Henceforth we will refer to the lower and higher concentration samples with the prefix L- and H-. Specifically we characterized 4 sample sets and one control: L-NF, L-F, H-NF, H-F, and PDMS (no nanocrystal inclusions). 52 Figure 4.2: SiNCs in solution, (a) photographed in room light and b) under UV excitation Shown are hexane, NF, and F samples from left to right. (c) PL spectroscopy shows much brighter luminescence for surface-functionalized (F) SiNCs as compared to nonfunctionalized (NF) SiNCs in solution. 4.3.2 Nanocomposite characterization 4.3.2.1 Photoluminescence Photographs of representative control PDMS and nanocomposite slabs in room light and under ultraviolet (UV) excitation are shown in Figure 4.3a-b. We also measured PL quantum yields from the nanocomposites, which are plotted in Figure 4.3c. These PL QYs underscore the high optical quality of the nanocomposites, at 1.2% and 16% on average for H-NF and H-F samples. 4.3.2.2 Microscopy Based on PL and visual inspection, it was not evident how the SiNCs were dispersed within the PDMS. To investigate the dispersion of SiNCs in the nanocomposite, we first assigned a specified spatial Cartesian coordinate system (Figure 4.4a) to each slab, sliced the slabs to obtain thin cross-sections, and then performed microstructural analysis using a Nikon AIRsi Confocal Laser Scanning Microscope (CLSM), in which thin slices can be imaged with white light as well as 53 Figure 4.3: (a) Photographs of SiNC/PDMS nanocomposites in room light and (b) under UV excitation (c) shows the average PL QY of SiNCs in hexane and in the nanocomposites, with error bars reflecting the standard deviation of the measurements. imaged at selected optical wavelengths using a laser excitation at 405 nm. The CLSM allows visual detection of the SiNCs and clusters of SiNCs, which are otherwise dicult to discriminate from the PDMS matrix using traditional optical or electron microscopy. The confocal nature of the instrument allows depth resolution of the luminescent features as well, so that one can view a composite 2-D image of 80 imaging planes (in X-Z, slice depth along the Y direction, Figure 4.4b, or a 3-D depth-resolved map of the dispersion (Fig.4.4c). For this 3-D map, the position of each SiNC cluster is represented in its physical location within the map as well as using color-coding for depth. Based on this 3-D representation we found that, regardless of surface functionality, the SiNCs are clustered together to some degree. This optical measurement technique is resolution- limited, and so we were not able to visualize the degree of unclustered dispersion of the NCs within the sample, but the results are important as they demonstrate the dierences in clustering behavior of the samples. Figure 4.5 shows representative composite images for each sample type, all acquired under identical microscope settings. The depth of each image is ⇠60µm, and the representative images in 54 Figure 4.4: (a) For reference during microscopy, we assigned each nanocomposite the same coordinate system, (b) A single CLSM image of a slice is shown and (c) the depth-resolved composite image (scale inset), demonstrates the distribution of small clusters within the sample slab. Figure 4.5 are a composite of 80 separate depth scans along that 60µm thickness. Visual inspection of the CLSM images indicates an increase in number of SiNC clusters at higher concentrations (HF and HNF) as compared to the lower concentration samples, which is not surprising. To quantify any dierences between the samples, we analyzed the cluster size for each sample type using a particle counting/volume algorithm via the ImageJ processing platform. We analyzed 3-4 individual images from dierent individual samples for each sample type using this algorithm to build a comprehensive picture of the SiNC dispersion characteristics. The results are plotted in Figure 4.6 using a box plot to display the five-number summary of the data. First, we note that the smallest cluster size for our data was resolution-limited, and so the lower bounds of each sample set are nearly identical. These lower bounds represent not the true smallest cluster sizes but instead the smallest size that is resolvable using the CLSM instrument and the processing algorithm. Aside from the increase in number of clusters, the data suggests that 55 Figure 4.5: Representative CLSM images of the SiNC/PDMS nanocomposites Figure 4.6: SiNC cluster size statistics based on image processing from the CLSM images. 3 separate images were analyzed for each sample type. A key to the data representative is shown as inset. the average cluster size and median cluster size do not significantly depend on the concentration of SiNCs in the polymer, at least over the concentration range we investigated. This indicates that 56 for the SiNC concentrations that we studied, the mixing and curing procedure we used is sucient to ensure relatively uniform dispersions. In other words, the distance between clusters within the polymer host at our “high” concentration of SiNCs is still so large that the cluster formation and dispersion is largely unaected. It is worth noting, however, that the larger particle concentrations sometimes required slightly modified PDMS curing procedures, namely extended time on the hot plate. At the 0.12 wt% concentration, which is modest by nanocomposite standards, the SiNC clusters still seemed to interfere with the curing process to some degree, occasionally leading to longer required curing times. Next we examine the eect of surface functionality of the SiNCs on the dispersions. One notable dierence between NF and F samples is the average cluster size. Visual inspection of the CLSM images indicates that the clusters for the NF samples were larger than for the F samples. The median cluster sizes as measureable using CLSM were nearly equivalent, between F and NF samples, but the spread of data (both the interquartile ranges and the outlying points) was larger for NF samples at both low and high concentrations. The mean cluster size of the SiNCs in the PDMS, represented by the X symbols in the plot, is also larger for the NF samples as compared to the F samples at both concentrations, although the discrepancy appears smaller at the lower concentration. This finding is consistent with the fact that surface functionalization causes a greater separation between individual SiNCs as well as improved integration with the monomer chains before polymerization. 4.3.2.3 Mechanical properties of the composite To test the eect of the concentration and surface functionality of the SiNC inclusions on the nanocomposite mechanical behavior, we used uniaxial tensile testing and a material model to estimate a characteristic mechanical parameter, and we also quantified rupture stretch. The resulting stress-stretch data (Figure 4.7a) can be analyzed to extract the neo-Hookean coecient (µ, with the dimension of a stress) based on the following equation for Cauchy stress (): = µ ⇤( 2 ( 1/)) 57 (4.1) Figure 4.7: Mechanical properties of the nanocomposites. (a) shows a representative plot of Cauchy stress vs. stretch data for each sample type. From this data, (b) we extracted the neo-Hookean coecient for each sample, showing a distinct reduction for the higher-concentration sample. (c) the rupture stretch was unaected by the SiNC inclusions. where is the applied stretch. Figure 4.7b shows the estimated neo-Hookean coecient of each sample, with standard deviations plotted as error bars. The rupture stretch, shown in Figure 4.7c, has a value of 2 for all samples including the control PDMS. However, the results for µ (Figure 4.7b) tell a dierent story: while the µ values for the L-NF and L-F samples are within the statistical variation of the control PDMS, the H-NF and H-F samples both exhibit a reduced µ, correlating to a reduced stiness. This reduction is most prominent for the H-NF samples. 4.3.2.4 Discussion These results have significant meaning. From an optical standpoint, inclusion of surface-alkylated SiNCs in PDMS via this procedure leads to brightly emissive nanocomposites with PL QYs around 16%, pointing to the promise of this materials system going forwards. The non-functionalized SiNCs still result in emissive nanocomposites, but the PL QY of these is quite low (⇠1-2%) and is dicult to measure at the lower concentration we studied. For optically emissive nanocomposites, using the surface-functionalized SiNCs is critical. The CLSM images make it clear that clustering of SiNCs within the PDMS is prevalent across sample type. Preventing this clustering eect using modified nanocomposite fabrication procedures or by tuning the surface of the SiNCs by using other organic ligands with alternate terminal groups are promising routes to explore, for even 58 higher-eciency photoluminescence from the nanocomposites. For example,Hill et al. revealed that the ligand terminal group greatly influenced the dispersion of SiNCs in a PMMA host [65]. This calls for a comprehensive investigation of the interaction between the ligand termination and the resulting dispersion of SiNCs in PDMS. Mechanically, our first conclusion is that inclusion of SiNCs at very low concentration (0.03 wt%) in the PDMS does not appear to influence either the rupture stretch or the modulus of the PDMS nanocomposites. At low concentration, the SiNC cluster sizes within the nanocomposites were widely dispersed, and despite a change in average cluster size between NF and F samples, the low concentration did not influence the overall mechanical behavior of the nanocomposites (although it did lead to bright PL in the case of the LF samples). Second, at higher concentrations both the NF and F samples demonstrated a significant reduction in the neo-Hookean coecient µ and the reduction is more prominent for the NF samples. When compared with the nonexistent change in µ for the lower concentration samples, this indicates that the modulus of the nanocomposites depends both on the surface functionality and on the SiNC concentration in the nanocomposite. We were cognizant of the fact that mixing the SiNCs into the PDMS prepolymer solution might be one cause for the large cluster sizes and the reduction in modulus. To test this hypothesis we also used a shear mixer (FlackTek SpeedMixer at 3000 rpm, 1 minute) for mixing H-F samples into PDMS. Our uniaxial tensile testing results and CLSM imaging from the samples prepared using shear mixing do not indicate any significant deviation from the samples prepared as described above. This points to solubility as the major cause for cluster formation and modified µ for the nanocomposites, rather than incomplete mixing. The reason behind the reduction in the neo-Hookean coecient for the high-concentration composites is cause for further consideration. Previously, we estimated µ for non-functionalized SiNC layers deposited on PDMS and found a value of 345 kPa, approximately twice the value for PDMS. While the parameters influencing this value are undoubtedly complicated, we can at least use it as an approximation for the modulus of the SiNC inclusions in the composites. Our CLSM imaging confirmed that a large fraction of SiNCs are not individually dispersed within the PDMS matrix but instead are grouped together in clusters 59 up to several hundreds of nm in diameter and thus have a high amount of self-same interaction rather than exclusively interacting with the matrix, and so considering their modulus within the nanocomposite as being similar to the modulus of SiNC layers is a reasonable approximation for the purposes of our analysis. Here, we can base our discussion around that value of µ as estimated previously. Our first approach to estimating the theoretical modulus of the composite employed the standard theory of mixtures, in which the µ values for the host and inclusion are treated as unchanging. The volume fraction of the mixture is used to weight the two µ values, as µCOMP = µPDMS ⇤ PDMS + µNC ⇤ NC (4.2) where, µ(COMP),µ(PDMS) and µ(NC) are the moduli of the composite, PDMS and SiNC respec- tively, and (PDMS)and (NC)are the volume fractions of PDMS and SiNC (calculated for each sample type using ImageJ). Based on the previous equation, the composite is expected to exhibit a modulus of 190.59 kPa, which is nearly identical to the modulus of PDMS (190 kPa) on its own. From the theory standpoint, this is logical due to the very low volume fraction of SiNCs (calculated theoretically at 0.1% for the high concentration samples). However, our measured value of µ in both the functionalized and non-functionalized samples was lower than the modulus of PDMS on its own. Even using the more sophisticated Mori-Tanaka approach [13, 120], the theoretical mixture modulus was near the value for PDMS. Recognizing this discrepancy, we also considered the possibility that the µ of the PDMS matrix itself may be altered by the presence of the SiNCs at higher concentration. The clusters of SiNCs are large enough to conceivably interfere with cross-linking between polymer chains, in eect countering the concentration of cross-linking agent and reducing the modulus of the PDMS [22, 126]. Considering this, we can work backwards from our measured µ for the composites, assuming the SiNC modulus as estimated in our prior work, and allow µPDMS to change. Using this approach, we calculated that the PDMS modulus for the H-F samples was ⇠40% lower than the PDMS without SiNCs, and for the H-NF samples it was even more dramatically altered, ⇠70% lower than for PDMS alone. Based on established trends for the mechanical properties of PDMS as a function 60 of cross-linking agent concentration, this would mean that by including SiNCs (F and NF) in the PDMS at a concentration of <1% causes an eective change in the cross-linking ability of PDMS equivalent to reducing the curing agent:prepolymer ratio to 1:12 and ⇠1:18, respectively, for the H-F and H-NF samples [118]. 4.4 Conclusions We measured the luminescence and neo-Hookean coecients of SiNC/PDMS nanocomposites based on concentration and surface functionality. The PL QY from the nanocomposites is high at 16% for the samples made with surface-functionalized SiNCs, and still measurable at 1-2% for the samples made with non-functionalized SiNCs at higher concentration. Our experiments indicate that the mechanical behavior of luminescent SiNC/PDMS nanocomposites depends on both the nanocrystal concentration within the PDMS as well as the surface functionality of the SiNCs. At low concentration (0.03 wt%) the SiNC inclusions do not influence the rupture stretch nor the neo-Hookean coecient of the nanocomposites, regardless of surface functionality. However, at higher concentration (0.1 wt%) there is a statistically significant reduction in the neo-Hookean coecient. This is in contrast to the report by Chen et al. [23], who found similar PL results for their SiNC/PDMS nanocomposites but reported a reduction in Young’s modulus (⇠three times µ at low applied strain and assuming a Poisson’s ratio of 0.5) and the tensile strength of hydrogenated (NF) samples but an increase in both for decyl-capped SiNCs in PDMS. We do not expect that the dierence between decyl- and dodecyl-capped SiNCs should have such a profound impact on the mechanical properties of the nanocomposites, neither at low nor at higher concentrations. Further, the PL results from our study largely concur with their findings. Thus, these conflicting results point to the importance of further characterization and an improved understanding of the exact mechanisms of interaction between nanocrystals, their surface functionality, and manufacturing procedures during fabrication of nanocomposites using elastomers such as PDMS. 61 CHAPTER 5 FUTURE 62 5.1 AuNP Application for Silicon NanoRods (SiNRs) Sythesis Throughout our experiments on synthesis of AuNPs, our overarching objective was to fabricate SiNRs with aspect ratio (AR) (Length/ Diameter) of 2-7. If successful, this work would introduce a brand-new plasma-based flow-through approach to continuous synthesis of SiNRs. Similar anisotropic SiNR growth in a fluid suspension was achieved by Heitsch et al, which used a ligand- assisted solution-phase method, based on Au-seeded solution-liquid-solid (S-L-S) growth [63]. Our focus is on translating this approach from solution to the plasma, which can avoid solvent use and could lead to wide tunability of the SiNR properties such as structure, AR, and surface in a continuous nanomanufacturing method. For plasma-based SiNR growth, we sought to build on existing methods and use seeded growth based on AuNPs. In Chapter 2, we reviewed dierent methods for fabrication of AuNPs and introduced our plasma-based method which provides a new opportunity for fabricating SiNRs based on multi-stage reactor design. Figure 5.1 shows a schematic diagram for the proposed multi- stage reactor for producing AuNP-assisted SiNRs. The ultimate goal of this reactor is continuously making gold nanoparticles and then using them as the seed catalysts for growing SiNRs and continuously, using the plasma as energy source. For our preliminary trials, we constructed several iterations of this reactor. The concept was based on multi-stage reactors, with the first stage used for AuNP synthesis, and a subsequent plasma stage for growth of SiNRs using a supersaturated eutectic mixture method. This reactor consists of gas cylinders connected to quarter inch tubing coming to MFCs and tubes, non-thermal RF plasma, electrical circuit for matching the power and stainless steel tubing for gas lines [107]. AuNPs are produced in the first stage using the hot-wire or plasma-based approach. After producing the AuNPs, in the second step these particles enter the second stage RF plasma (operated at 13.56 MHz and with supplied powers ranging from 5 to 40 W), tuned with a lab-constructed impedance matching networks including an adjustable capacitor, inductor coil, and air gap. SiH4 was provided by a SiH4/Ar mixture cylinder (1% SiH4-99% Ar) as injected into the reactor from a 0.5-inch O.D. sidearm at flowrates between 5 and 50 sccm. The eutectic alloy 63 temperature for the Au/Si alloy is 363°C, and so we sought to explore the parameters necessary to reach this temperature by adjusting reactor parameters such as power, gas flowrate, overall pressure, and reactant partial pressure. Figure 5.1: (a) Schematic Diagram for SiNRs synthesis (b) early stage experimental set up with hot wire method (Courtesy R.J.Anthony) We modified the early stage design shown in Figure 5.1 by adding the Silane from an elevated side arm between the two stages. The early design added silane in the middle of the second stage, as presented in Figure 5.1, depicted in Figure 5.2. After using the reactor for ⇠ 20 mins, a thick film deposition of Silicon formed in the side arm and on the main reactor tube, preferentially on the injection side. We hypothesized that eective mixing of the AuNP-Ar flow from the first step and Silane from the second stage did not take place, causing the SiH4 to remain spatially segregated, decomposing into film growth near the sidearm rather than mixing with the bulk flow. Figure 5.3 shows the modified reactor, including the side arm for adding the Silane between two steps. This modification greatly reduced the gas mixture homogenization problem, and eliminated the preferential film growth near the side arm. We collected the reaction products at the bottom of the second stage in the reactor, and performed TEM characterization to investigate possible SiNR 64 Figure 5.2: High flowrate of AuNP-Ar from the first step(600-1200sccm) and low flowrate of SiH4 from the side arm did not mix properly in the second step of the reactor Figure 5.3: Modified hotwire AuNP reactor for sythesis SiNRs growth. Surprisingly the microscopy showed a widely size-dispersed set of silicon nanoparticles with both crystalline and amorphous structures. We had expected to see some mix of AuNPs, whether the SiNR growth had happened or not, as we were certainly injecting AuNPs into the 65 second stage. EDS provided some proof for AuNP from the first step collected after the second step of adding Silane without turning on the RF plasma in second stage. We focused on the form of the plasma, which included a filamentary structure (constricted region of high plasma density) during the second reaction. Other reports have indicated that, in argon/silane plasmas, the plasma density and temperature spikes reached by the nanoparticles in the plasma are related [3, 84, 109]. We hypothesized that the constricted plasma filament caused sucient power density to in fact coalesce the smaller SiNCs typically produced under the same pressure/flow/power conditions but with a diuse plasma. While Si nanoparticles exhibit size-related melting point depression [67, 108], the larger crystalline Si nanoparticles we observed had diameters >10 nm, indicating that they reached temperatures in excess of 1300 K. Not only this is higher than the eutectic alloy temperature we were seeking to reach, it also exceeds the melting temperature of gold. Our hypothesis was that by passing through the filamentary region, the AuNPs were re-evaporated (explaining their disappearance before collection on the TEM grid) and the SiNCs were coagulated and recrystallized. The solution we proposed was to eliminate the filament using gas flowrates. We were able to stop the filament in the reactor by controlling the gas species flowrate inputs to the second stage (in particular, by keeping the SiH4 flowrate <5 sccm) and by controlling the power to the plasma (Rf power limited to the range of 5-15 Watts). Photographs of this second- stage plasma with filament present (a) and the diuse plasma (c) achieved using limited power and reduced SiH4 flowrate are presented in Figure 5.4. Unfortunately, though, we probed a wide range of parameters within the confines of avoiding the filamentary plasma, we never observed anisotropic growth from this reactor configuration. 66 Figure 5.4: (a) Filament formation associated with high plasma density (b) High plasma density resulted in wide size range of Silicon nanoparticles with both crystalline and amorphous structures. (c) Controlling the SiH4 flowrate and RF power stoped fillament and resulted in narrow size distribution for Silicon nanoparticles As presented previously, the plasma-based method works with wider ranges of experimental parameters, compared to the hotwire method. Thus, we next attempted to combine plasma- synthesized AuNPs and SiNR growth. Figure 5.5 shows a schematic picture of the projected dual-plasma reactor. Future work will involve exploring the full range of capabilities of this dual-stage reactor. 67 Figure 5.5: Schematic of dual-plasma reactor for AuNPs-SINRs simultaneous synthesis 5.2 Modulus of SiNC Layers on PDMS: Parametric Investigation 5.2.1 Introduction In-situ investigation of mechanical properties of nanomaterials in form of thin films is one of the challenging obstacles in advancement of nanotechnology and nanomaterial fields. There are many studies dedicated to plastic properties of individual nanoparticles in crystalline and amorphous forms [100]. Nanoindentation experiments and other physical probe-based techniques can be used to evaluate single-nanocrystal deformation in elastic and plastic regime. These techniques generally require especial microscopy techniques like tunneling electron microscopy. However, as these approaches are mostly irreversible and/or destructive, there is a significant interest in non- destructive methods for evaluating properties of thin films, especially for layers of nanocrystals on elastomeric substrates. The future of electronics strongly relies on advancement in mechanical flexibility, sustainability and biocompatibility of electro-optical sensors, health monitoring systems, 68 photovoltaic devices, displays, and functional coatings. We recently introduced a novel in-situ approach for estimating mechanical properties of luminescent SiNCs deposited on flexible substrates by measuring the onset of instabilities in the SiNC layers on PDMS under finite bending deformation as reported in Chapter 3. This study revealed a neo-Hookean coecient (µ, analogous to shear modulus at low stress/strain) of a 4.5 µm thick SiNC film to be 345 ± 23 kPa, which is significantly smaller than the young modulus of a single SiNC at the value of 170 GPa. There are many parameters that could influence the mechanical behavior of the SiNC layers on PDMS, including layer thickness, surface functionality, and mechanical properties of the PDMS substrate. In order to reveal some of these interrelationships, we investigated these parameters regarding their influence over the mechanical properties of SiNC layers on PDMS. Our results indicate that the layer of SiNCs and PDMS parameters exert competing influences on the mechanical properties of the SiNC layers. For example, we found that reducing the stiness (reducing µ) for PDMS leads to a decrease in µ for the SiNC layer. However, the relationship is not monotonic. Unraveling the connection of these parameters on the physical, mechanical, and optical properties of the SiNCs will allow advancement for predictive engineering of nanomaterial thin films, for stretchable and flexible devices. 5.2.2 Experimental work We prepared the PDMS in lab using a Sylgard prepolymer PDMS 184-kit and curing agent catalyst produced by Dow Corning. The PDMS mechanical properties depend on the prepolymer PDMS kit-184/curing agent catalyst ratio and curing conditions during PDMS preparation. We produced a range of substrate with same geometrical shape and dimensions (defined earlier in Chapter 3 as slenderness ratio). For experiments designed around modifying the SiNC layer, the prepoly- mer:curing agent ratio was 10:1, as before. For experiments to test the dependence of SiNC layer modulus on PDMS modulus, we altered this ratio to produce PDMS slabs with lower moduli. With a higher fraction of curing agent, the modulus of the resulting cured PDMS increases. [78]. We attempted to increase the modulus for this study by using a 7:1 ratio of prepolymer to curing agent, 69 but our experiments revealed that the critical angle required for bifurcation onset was too large to measure practically. Instead we reduced the amount of curing agent, using ratios of 10:1, 12:1, and 20:1 for this parametric study. We used the same method for evaluating µ for the substrates as discussed earlier in Chapter 3 and Chapter 4. For performing this parametric study, we chose a 3/1 slenderness ratio for our substrate. As we mentioned earlier there are many advantages for gas phase processing. Similar to Chapter 3 and Chapter 4, we synthesized SiNCs using a low-pressure non-thermal plasma. The deposition of nanocrystals occurs directly out of the described reactor into thin-film layers using inertial impaction through a slit-shaped orifice onto a substrate placed beneath[68, 124]. This reactor enables us to control thickness of the film, pressure, sample-orifice distance and forwarded RF power to the glass tube. As we discussed in Chapter 3 there is a stand o distance between orifice outlet and the top surface of PDMS substrate which is around 4-5 mm. This distance plays a significant role on the thickness and porosity of the thin film during deposition via back and forth rastering motion of the substrate. For instance, decreasing the distance would result in a thin film with lower porosity (higher density) and increasing the distance on the other hand, would result in higher porosity (lower density) [5, 68]. This is a consequence of the dynamics of SiNCs exiting the reactor from orifice outlet. Since we fixed the value for stand o distance and the frequency of rastering motion for substrate (3 sec for each back and forth motion), the only parameter which impacts the thickness of the thin film is the number of rastering cyclces. In other words, changing the time of rastering from 20 mins as discussed in Chapter 3 to 10 mins and 5 mins results in a thinner and thinnest film in our results, respectively. We also produced samples with higher rastering time, 30 mins, as well. We used SEM analysis to evaluate the thickness of the layers depending on rastering time. As we discussed the surface functionalization of the SiNCs will improve their PL properties, but changing the surfaces may also influence the mechanical behavior of the layer as a whole. We used a facile method to surface functionalize the SiNCs in the gas phase during the plasma process 70 as previously studied by Anthony et al. [4]. Here, we used 1-dodecene for surface functionalization. We also measured the PL and QY for the produced film and results presented in Fig 5.10. We also used SEM to compare the surface morphology of non-surface functionalized and surface functionalized thin films under similar conditions. In a summary of our experiments, Figure 5.6 presents a work flow of this parametric study. Figure 5.6: Schematic diagram for investigating the impact of film thickness, surface functionality of SiNCs via gas phase plasma and shear modulus of the PDMS substrate on mechanical properties of SiNC thin film 5.2.3 Results Film Thickness: We used a mechanical diamond glass cutter pen to mark our samples for thickness measurements. Figure 5.7 depicts the cross sectioning of samples produced via non-thermal plasma with various rastering deposition times. Regarding the the SEM image, which consisted from samples with 5 mins , 10 mins and 30 mins rastering time, the thickness of the film increased linearly with respect to the time of rastering deposition in plasma reactor. We performed a linear 71 regression and the R2 value was 0.9918. We also attempted a exponential fit, for which the R2 value was 0.9602. We expect that the deposition rate is linear, but it is possible that at low thicknesses there is a deviation from this behavior. Future work will investigate the sub-micron film thicknesses to evaluate if the rate remains constant even very close to the substrate, where packing eects may influence the porosity of the SiNC layer. In Figure 5.7 sample in red represented the results from Chapter 3 for comparison. Figure 5.7: SEM images for PDMS with 3:1 slenderness ratio as defined in chapter 3 with rastering deposition time of 5 mins (Sample 1), 10 mins (Sample 2) and 30 mins (Sample 3) and averaged measured thickness of the film with corresponding standard deviation in comparison with the sample discussed with same slenderness ratio and 20 mins (in red)of rastering deposition time from previous chapter Modulus of SiNC Layers (Estimated Values): We used the same methodology employing the onset of bifurcation to estimate the neo-Hookean coecient for the SiNC layers subject to the parametric variation as noted above. The results are plotted in Figure 5.8. The asterisks on the bars indicate samples which had a statistically significant deviation from the value we obtained for the samples originally examined and reported in Chapter 3. 72 Figure 5.8: The methodology introduced in Chapter 3, employed to estimate neo-Hookean coef- ficient for SiNC layars corresponds to parametric variation. The asterisks indicate samples with significant deviation from value we abrtained in Chapter 3 The first observation is that the thickness of the SiNC layers does influence the neo-Hookean coecient, with thinner layers exhibiting increased µ as compared to the thicker layers. However, increasing the thickness beyond the 4-5 µm, which was our standard, does not appear to alter the µ of the SiNC layers. From this, our hypothesis is that the soft PDMS substrate creates microstructural changes or densifications of the SiNCs, altering their mechanical behavior close to the substrate - yet once the layer reaches a certain critical thickness, the eect diminishes as the subsequent layers are deposited and they are less able to respond to the substrate’s mechanical properties during deposition. Figure 5.9 presents the calculated mechanical property (neo-Hookean coecient )of the SiNC thin film deposited on PDMS. The asterisks show the samples with neo-Hookian value significantly deviate from samples reported in Chapter 3. Next, we can observe that the modulus of the PDMS also changes the resulting SiNC modulus, 73 Figure 5.9: Neo-Hookean coecient of all sample introduced in this section.The samples with asterisks, present the data with siginficant deviation from samples reported in Chapter 3 with decreasing PDMS modulus creating a decreased µ for the SiNCs. Our hypothesis to this regard is still very nascent and this research is ongoing, but we can draw some ideas from what we observed during the thickness measurements. The thinnest layers of SiNCs on substrates made with the standard PDMS recipe (10:1 prepolymer:curing agent) have a higher modulus (are stier) as compared to the thicker layers. From this it is reasonable to expect that near the substrate there are eects from the impact of the NCs onto the PDMS, but as the layer becomes thicker these substrate eects are diminished. When we decrease the modulus of the PDMS by altering the recipe, we possibly alter the damping eect of the SiNCs impacting upon the substrate. We know that the thickness of the SiNC layer does not vary with changing PDMS modulus, but it’s possible that the propagation of the substrate eect is reduced for the substrates with lower modulus. Our ongoing experiments will check this possibility by studying very thin layers of SiNCs (rastered for  5 mins) to test the limits of the substrate eect in terms of how far into the SiNC film is plays a role in the mechanical behavior. 74 Notable is that the surface functionalization plays no statistically significant role in the modulus of the SiNC layers (see Figure5.9). This is especially important regarding the fact that the SiNCs with in-flight surface functionalization display much higher PL quantum yields as compared to those made without surface functionalization (See Figure. 5.10). We can conclude that, for brightly emissive layers of SiNCs on PDMS, one can perform surface functionalization without disrupting the layer’s mechanical behavior. Figure 5.10: The surface functionalization applied in gas phase during plasma processing and sythesis of SiNCs. The PL and QY measured. This results coupled with mechanical property estimation of corresponding samples presented in this Figure. The results here, confirm that improving the oprical properties and QY do not interfere with mechanical property of the SiNC thin film. 5.2.4 Conclusions This study is nearly complete. In summary, we have investigated some of the parameters of SiNC layer properties and PDMS properties. We discovered that reducing the thickness of the SiNC layer increases the modulus of the layer, and that in-flight surface modification greatly improves 75 the PLQY but does not appear to significantly alter the modulus of the SiNCs. However, reducing the PDMS modulus by altering the recipe for fabrication causes a significant reduction in the SiNC layer modulus. Our ongoing experiments and modeling will seek to explain these parametric eects and recommend future experiments. 76 5.3 Unbending- Another Finite Bending Condition for Mechanical Property Estimation 5.3.1 Introduction Prior experiments involved the deformation of SiNC layers deposited on PDMS in a relaxed condition, and the bifurcations we observed upon bending the system as a whole. Here, instead we focus on experimental observations and physical circumstances which result in surface instabilities (wrinkling) of SiNCs layers deposited onto PDMS substrates in a prestretched condition, with future applications in tunable gratings and new manifestations of LEDs and other devices. Prior work has shown the emergence of wrinkling patterns in these systems [104], but here we have performed experiments to understand the conditions of the SiNC layers and the PDMS under which these wrinkling patterns are formed. 5.3.1.1 SiNC deposition In this section, we discuss the various deposition schemes for placing the SiNCs layers onto the PDMS. The goal is to deposit a uniform layer of SiNC onto the PDMS as thin film, under varying pre-stretched conditions. We are interested in studying how the optical and physical properties of the SiNCs are altered when various degrees of pre-strain are applied to the PDMS. We prepared a stage (Figure 5.11) which can be used to obtain any stretching ratio between 0- 100%, and which is compatible with holes in both ends to be mounted onto a pushrod in the plasma reactor. This apparatus helps us to use PDMS as a substrate since we apply coaxial stress-strain to substrate. One end is fixed, and the other end is moving with a corresponding twisting screw. We achieve uniform deposition by rastering the substrate as mounted on the stage underneath the orifice. The rastering is a back and forward movement which helps to keep the thickness uniform throughout the substrate. While this allows for easy control over the applied pre-strain in the long direction and uniform deposition, it also leads to a pronounced bi-axial deformation due to the Poisson’s eect of the PDMS, meaning that during the applied linear pre-strain there is also a 77 compression across the perpendicular edge of the PDMS. This compression is measured to be ⇠ 10 percent at the waist of compression in the middle section (Figure 5.12). Figure 5.11: Sample holder for various stretching ratios (left), sample holder for finite bending around pushing rod (right) Figure 5.12: Above view of sample, relaxed after deposition 5.3.2 Results As we discussed in previous section about loading the substrate in the reactor, here, we present the results for both methods. In the first method, the PDMS sample was loaded on the sample holder and pre-stretched between 5% to 40% beyond the initial length (relaxed). After the SiNCs were deposited uniformly onto the PDMS in the reactor via inertial impaction, we used the screw on the stage to return the substrate to its original length steadily. During this process, SiNCs (which were experiencing no tensile force or compression on the pre-stretched PDMS) start to experience compression and this makes surface instabilities on SiNC layers. These wrinkles are observed by mounting the stage-sample setup in a vacuum chamber of SEM. The corresponding wrinkles for 20% applied pre-strain are depicted in Figure 5.13. Their axis can be categorized based on the area in which they formed, depending on the tensile or compressive forces that they experience during relaxation of the PDMS. Again, we see this combination of forces due to the Poisson’s ratio of 78 the PDMS, leading to the SiNC layer experiencing compression near the clamps and tension near the waist of the pre-stretched PDMS. We categorized these responses to pre-stretched forces into 3 regimes based on position on the pre-stretched substrate: pure compression, pure tension, and a combination. In order to study the whole deposited area, we divided each deposited area of sample to 15 cells, including 3 rows (A-B-C) and 5 columns (1to5). The corresponding deposited area with cells shown in Figure 5.14 and Figure 5.12. Since the PDMS applied pre-strain was 20%, the Figure 5.13: SEM image of deposited area with 20 percent pre-stretching condition, Column A3- B3-C3 (left), Columns A2-B2-C2 and A4-B4-C4 (middle), Columns A1-B1-C1 and A5-B5-C5 (right) Figure 5.14: SEM image of deposited area with 20 percent pre-stretching condition, Column A3- B3-C3 (left), Columns A2-B2-C2 and A4-B4-C4 (middle), Columns A1-B1-C1 and A5-B5-C5 (right) shape of the substrate during the SiNC deposition is not rectangular. Moreover, this new geometry will result in an interesting force response that include wrinkles on the surface of the SiNCs. The SEM images from dierent regions are shown in Figure 5.13. We can categorize the corresponding surface instabilities into 3 distinctive types: each occurs in specific cell areas depicted in Figure 5.14. As we can see, in the area near clamps of both side, where the deposited area changes 79 its thickness to some sucient amount we observe vertical Lines. These areas include columns A1-B1-C1 and A5-B5-C5. However, in the column A2-B2-C2 and A4-B4-C4 we observe transient or inclined directions with respect to the X-Y Cartesian coordinate introduced in Figure 5.12. In the center, including middle column (A3-B3-C3) we observe horizontal wrinkles. Since these mechanical instabilities are dierent in terms of direction they appear throughout the deposited area, we justify the observation from mechanical point of view. Figure 5.15 shows a diagram of the forces on the PDMS during stretching and corresponding tension/compression experienced by the PDMS during stretching. In order to understand the force distribution (tensile vs. compressive), we represented the PDMS in a top view schematic before deposition (in pre-stretch condition) and after deposition (in relaxed condition). The stage is set to move in one direction and it designed to stretch the PDMS from 100% (Zero sterteched) to 200% (double stretched), depending on the experiment design and initial PDMS dimensions. The Poisson Ratio for PDMS is ⇠ 0.5 [138]. This means that the strain in longitudinal direction is twice the strain in compression along the middle section of the sample. So, the PDMS area near the clamps Figure 5.15: pre-stretched before deposition or during deposition experience tension, however, one clamp is fixed in position. During deposition SiNCs these areas are deposited on a stretched surface but experience zero force while the SiNC layers are formed. After deposition, when the PDMS is relaxed back to its original length, SiNCs will experience compression near the edges (columns 1 and 5 in our areal mapping scheme) and they start to form surface instabilities as shown in the Figure 5.13 (right). However, the middle section of PDMS (column 3) is under compression during the deposition. By returning to relaxed condition, the SiNC layer in the middle section would experience tension. Ultimately, in this middle section the 80 SiNC layers form instabilities perpendicular to the ones near edges. Between the edge regions and the middle region (columns 2 and 4), the SiNCs experience a hybrid of these responses. In order to limit the scope of this study, we focused on to the formation of instabilities near edges, which result from the SiNC layers experiencing pure compression after the PDMS is relaxed. As a part of manufacturing design, we had to provide a design to able reproduce this (near edges) kind of surface instabilities which in order to categorize them in a similar formation direction since we apply one-dimension force to the substrate. Due to the nature of the PDMS, it is impossible to eliminate the tensile force experienced by the SiNCs when the PDMS is pre-stretched using the stretching stage we described hre so far. Thus, we have to provide a new experimental set up for loading the PDMS: our solution is to use finite bending of the PDMS. This approach will mostly eliminate the Poisson’s ratio eects on the region in the middle of the PDMS. Since the PDMS wraps around the pushrod, the upper surface (SiNC-coated surface) experiences a strain in the circumferential direction and there would be no dimensional change corresponding Poisson’s eect for substrate. We explored finite-bending as a way to minimize the bi-axial eect as observed for the stretching stage described above. To achieve this, we cast the PDMS into aluminum molds of calculated size, attached handles to the PDMS during the process with a commercial painting foam (Foam Board), and wrapped it around the pushrod, gripping the handles together at the base using a staple. Then, SiNCs are deposited onto the PDMS (which is deformed via this finite bending), and when the handles are released, the PDMS returns to its relaxed state. Figure 5.11 shows both setups. Figure 5.16 shows a picture of a prospective sample going under finite bending elastic deformation. We wrapped PDMS around the pushing rod and deposited SiNCs on the wrapped substrate. In order to guarantee an evenly deposited film on substrate we have to clamp the PDMS (Figure 5.16). The microscope picture shown in right side of the Figure 5.16 , is captured by Optical-Electrical Microscope (Keyennce) under aforementioned conditions. This picture confirms the formation of wrinkles with the new design for substrate. However, since we applied some force to clamp the PDMS all around the push-rod, there would be extra force we applied for wrapping around the rod. 81 As the next step, we had to find exact substrate dimensions that allow the substrate to wrap around the rod without the need of additional clamping or stretching. Figure 5.16: Early design for finite bending surface instabilities by wrapping PDMS around the rod, the picture on right is the microscope image of the sample after unwrapped with the scale bar of 10 micrometer To measure the exact size, we need the sample to be wrapped around the pushrod with no tension. We used a mathematical calculation to find the width to thickness (w/t) ratio for a corresponding pre-strain (here it is 20%). So then with that ratio, we are able to wrap the substrate around the pushrod with no extra force which makes it feasible to have surface instabilities based on pure finite bending. (Figure 5.17). Based on the Constitutive Laws and Neo-Hookean assumptions given in Figure 5.17, we are able to apply the corresponding w/t ratio to make an aluminum cast for fabricating PDMS with handles. Figure 5.17 shows the corresponding result for this set up, beside a picture of the setup, including the aluminum cast and flexible handles for sample to be wrapped around the push-rod and securely fixed in place during deposition. This helps us to make sure that during the deposition the rotation of substrate with the pushrod for uniform deposition will not change the position and distance of substrate from the orifice which is an important factor for formation of uniform and continuous SiNCs film on the substrate. The thickness and width of the PDMS sample, in the reference configuration, is taken in such a way that there are no tractions on the surface when it is rolled on a cylindrical rod of a given outer radius. When the system undergoes the unbending deformation the thin SiNC layer deposited on the outside surface of the PDMS experiences a compressive force. This is similar to the compressive force faced by the top surface of the SiNC in the method involving finite bending. Therefore, like in the finite bending experiments, we expect the formation of instabilities on the surface of thin film above a critical 82 compressive stress. This value of compressive stress depends upon the geometric configuration and mechanical properties of the system. We now able to produce surface instabilities based on pure Figure 5.17: A new method for pure bending surface instabilities, the mathematical correlation and assumptions for incompressible, Neo-Hookean material (left) bending, and the microscope image also confirms this result. Figure 5.18 shows a set of wrinkles produced by pure bending eect. These wrinkles spanned over dierent areas in the sample including rows A1-A2-A3-A4-A5 and C1-C2-C3-C4-C5. The surface instabilities will be studied Figure 5.18: A new method for pure bending surface instabilities, the mathematical correlation and assumptions for incompressible, Neo-Hookean material (left) by their amplitude and the size. Since the mechanical properties leading to instability formation are dependent on the bilayer system of PDMS and SiNCs, this study provides an opportunity to correlate the luminescence and mechanical properties of the SiNC layer based on the surface wrinkles. Our early calculation is based on images from an optical microscope which helps us measure the size of the wrinkles based on where the wrinkles are formed. Table 5.1 provides a list 83 of operation conditions used for making surface instabilities in SiNC plasma reactor. The time of Table 5.1: List of operation used for making surface instabilities in SiNC plasma reactor at PNL deposition can turn into number of rotations of sample around the rod underneath the orifice. To calculate this number, we measured each rotation time under the orifice in the plasma deposition reactor which is ⇠ 3 seconds. This will help us to compare the thickness of the film quantitatively. The forward power is the power sent out to the plasma reactor and can be optimized with matching box. It is the operator’s duty to check the appearance (shape, intensity and length) of the plasma for set of recorded parameters. The power range stated in the table is used to form SiNCs with high crystallinity. The crystallinity can be checked by XRD. To verify the crystallinity, we performed a brief power study using the same flowrates and pressure, however, the power received was 30 and 25 respectively. Figure 5.19 shows the XRD patterns from these samples, indicating a higher degree of crystallinity for 30 Watt power than 25 Watt. The clear peaks from the [111], [220] and [311] crystallographic planes at 28.460°, 47.510°and 55.910°observed respectively [130]. This can be further verified by Raman Spectral Analysis [74, 107, 128]. This has been proven with other studies made for SiNCs [3]. 84 Figure 5.19: XRD for SiNCs deposited on PDMS, 30 Watt(Left), 25 Watt (right) Figure 5.20 shows a peak-to-peak distance measurement for these surface instabilities as based on the optical microscopy. The microscope image shown in Figure 30 is one of the corresponding samples used for this measurement. Lamda () is the length of peak to peak separation of two wrinkles next to each other. This changes from⇠ 1 micron to more than 20 microns, corresponding to the dierent cell areas we defined for each PDMS sample (see Fig. 22). The aforementioned system includes 15 equal size cells and these wrinkles repeated with dierent size in various cells. Moreover, since each rotation under orifice takes ⇠ 3 seconds, then 10 rounds equal to 30 seconds (⇠ 100nm), 20 rounds equal to 60 seconds (⇠ 200nm) and 30 rounds equal to 90 seconds (⇠ 300nm). These thickness calculations are based on experimental image samples captured by SEM. Increasing the time or number of rotation (while we keep same speed for rotation) will result in thicker films on the substrates. Regarding the previous SEM thickness measurements Figure 5.20 includes average size with corresponding standard deviation. Since we have 3 samples with varying thicknesses we can measure the of the wrinkles for each sample. All calculations are based on optical microscope images. 85 Figure 5.20: Lambda () measurement for surface instabilities, 10 round rotations in the reactor(up- left), 20 round rotations(up-right) and 30 round rotations (bottom), numbers are corresponding to the cells that measurement done with specified standard deviation for each averaged size letter corresponding to the cell which wrinkle instabilities observed Figure 5.21: A sample of PDMS deposited with SiNCs under pre-stretched condition for 10 round rotation(left), definition of Lambda () and thickness of the SiNCs film(right) The measured based on the schematic diagram shown in Figure 5.21. As demonstrated in Figure 5.20, is not changing much for most of the samples, however, near the edges, including(A1- A5 and C1-C5) the standard deviation is bigger, which means the size of changing from few microns to 10 microns. To able to track the surface instability regimes and their amplitude changes on the surface we have to observe the cross-section image of formed wrinkle patterns by microscopy. Cross-sectional 86 imaging for these soft substrates has been an issue for many instruments and techniques including surface and thickness measurement by spectroscopic ellipsometry and Focused Ion Beam (FIB) cross-sectioning with SEM. Atomic Force microscopy (AFM), however, is a physical and optical method which provides new information about surface morphology in micrometer scale. This method has its own limitation including the maximum scanned size per each cycle. This method has been used for the sample shown in the result section for 20 rounds and the image shown in Figure 5.22. In this figure, a 3D model is constructed based on surface topology data from the AFM tip over a 20x20 micrometer area. This 3D model is conjugated with a cross section estimation to help us evaluate the amplitude of these surface instabilities. In future experiments, we will study the influence of SiNC size and the regime of wrinkles. Ultimately, we intend to study multilayer film composed from SiNCs with dierent sizes to study varied emission wavelengths. Figure 5.22: Atomic Forced Microscopy (AFM), AFM scanned a line across a 20 by 20 micrometer area and provided a 3D model and a cross-sectional view for investigating the amplitude of these surface instabilities 5.3.3 Conclusion In this section, we introduced another class of finite bending problem. Surface instabilities formed on thin film of SiNCs deposited on PDMS and their corresponding patterns categorized based on tenssion and compression deformation of the substrate. In order to control the formation of surface instabilities, we intorduced finite bending (unbending) for prestretched substrate (PDMS) undergoing plasma deposition process. In this configuration, final bilayer system unwrapped and compressed the thin film of SiNCs on top. Surface instabilities formed and their size and presence 87 redorded for dierent thicknesses of the thin film. Our study showed these surface instabilities change to cracks and fractures on the thin film by increasing thickness of the thin film. The size of these instabilities will be correlated to the mechanical property ratio of the thin film to the one of the substrate. These results will be coupled by a theoretical moding to estimate mehanical properties of the SiNCs deposited on PDMS as thin film. 5.3.4 Future work A theoretical model will be applied to the system introduced in this section for mechanical property estimation. This evaluation will be compared with results presented in Chapter 3 and parametric study introduced earlier in this chapter. This eeort will provide a new technique for estimating the mechanical property measurement of the thin films of SiNCs in an in-situ and non destructive fashion. 88 BIBLIOGRAPHY 89 BIBLIOGRAPHY [1] [2] [3] [4] [5] [6] [7] [8] [9] 1983. Joint assessment of commodity chemicals : JACC. Joint assessment of commodity chemicals : JACC . Andrews, E, W Sanders & L J Gibson. 1999. Compressive and tensile behaviour of aluminum foams. Materials Science and Engineering A 270(2). 113–124. doi:10.1016/S0921-5093(99) 00170-7. Anthony, Rebecca & Uwe Kortshagen. 2009. Photoluminescence quantum yields of amor- phous and crystalline silicon nanoparticles (July). 1–6. doi:10.1103/PhysRevB.80.115407. Anthony, Rebecca J, Kai-Yuan Cheng, Zachary C Holman, Russell J Holmes & Uwe R Kortshagen. 2012. An all-gas-phase approach for the fabrication of silicon nanocrystal light-emitting devices. Nano letters 12(6). 2822–2825. Anthony, Rebecca J., Kai Yuan Cheng, Zachary C. Holman, Russell J. Holmes & Uwe R. Kortshagen. 2012. An all-gas-phase approach for the fabrication of silicon nanocrystal light-emitting devices. Nano Letters 12(6). 2822–2825. doi:10.1021/nl300164z. Anthony, Rebecca J, Kai-Yuan Cheng, Zachary C Holman, Russell J Holmes & Uwe R Kortshagen. 2012. An all-gas-phase approach for the fabrication of silicon nanocrystal light-emitting devices. Nano letters 12(6). 2822–2825. Anthony, Rebecca J, David J Rowe, Matthias Stein, Jihua Yang & Uwe Kortshagen. 2011. Photoluminescence: Routes to Achieving High Quantum Yield Luminescence from Gas- Phase-Produced Silicon Nanocrystals (Adv. Funct. Mater. 21/2011). Advanced Functional Materials 21(21). 4041. doi:doi:10.1002/adfm.201190092. Baldwin, Richard K, Katherine A Pettigrew, Jayne C Garno, Phillip P Power, Gang-yu Liu & Susan M Kauzlarich. 2002. Room temperature solution synthesis of alkyl-capped tetrahedral shaped silicon nanocrystals. Journal of the American Chemical Society 124(7). 1150–1151. Bandodkar, Amay J., Rogelio Nuñez-Flores, Wenzhao Jia & Joseph Wang. 2015. All- printed stretchable electrochemical devices. Advanced Materials 27(19). 3060–3065. doi: 10.1002/adma.201500768. [10] Bao, Sungjun, Gunuk Wang, Byungjin Cho, Yonghun Kim, Sunghoon Song, Yongsung Ji, Myung Han Yoon & Takhee Lee. 2012. Flexible molecular-scale electronic devices. Nature Nanotechnology 7(7). 438–442. doi:10.1038/nnano.2012.81. [11] Bao, Zhenan & Xiaodong Chen. 2016. Flexible and Stretchable Devices. Advanced Materials 28(22). 4177–4179. doi:10.1002/adma.201601422. [12] Batson, PE & JR Heath. 1993. Electron energy loss spectroscopy of single silicon nanocrys- tals: the conduction band. Physical review letters 71(6). 911. 90 [13] Benveniste, Yakov. 1987. A new approach to the application of Mori-Tanaka’s theory in composite materials. Mechanics of materials 6(2). 147–157. [14] Biot, Maurice A. 1965. Mechanics of incremental deformations. [15] Boies, Adam M, Pingyan Lei, Steven Calder, Weon Gyu Shin & Steven L Girshick. 2011. Hot-Wire Synthesis of Gold Nanoparticles. Aerosol Science and Technology 45(5). 654–663. doi:10.1080/02786826.2010.551145. [16] Boies, Adam M, Jerey T Roberts, Steven L Girshick, Bin Zhang, Toshitaka Nakamura & Amane Mochizuki. 2009. Sio2 coating of silver nanoparticles by photoinduced chemical vapor deposition. Nanotechnology 20(29). 295604. [17] Bouchoule, A & L Boufendi. 1993. Particulate formation and dusty plasma behaviour in argon-silane rf discharge. Plasma Sources Science and Technology 2(3). 204. [18] Boulos, Maher I, Pierre Fauchais & Emil Pfender. 2013. Thermal plasmas: fundamentals and applications. Springer Science & Business Media. [19] Braithwaite, N St J, J P Booth & G Cunge. 1996. A novel electrostatic probe method for ion flux measurements. Plasma Sources Science and Technology 5(4). 677. [20] Brust, Mathias, Merryl Walker, Donald Bethell, David J Schirin & Robin Whyman. 1994. Synthesis of thiol-derivatised gold nanoparticles in a two-phase Liquid-Liquid sys- tem. Journal of the Chemical Society, Chemical Communications (7). 801–802. doi: 10.1039/C39940000801. [21] Burrows, Nathan D, Ariane M Vartanian, Nardine S Abadeer, Elissa M Grzincic, Lisa M Jacob, Wayne Lin, Ji Li, Jordan M Dennison, Joshua G Hinman & Catherine J Murphy. 2016. Anisotropic Nanoparticles and Anisotropic Surface Chemistry. The Journal of Physical Chemistry Letters 7(4). 632–641. doi:10.1021/acs.jpclett.5b02205. [22] Chagnon, Grégory, Marie Rebouah & Denis Favier. 2015. Hyperelastic energy densities for soft biological tissues: a review. Journal of Elasticity 120(2). 129–160. [23] Chen, Dongzhi, Wei Sun, Chenxi Qian, Annabelle P Y Wong, Laura M Reyes & Georey A Ozin. 2017. UV-Blocking Photoluminescent Silicon Nanocrystal/Polydimethylsiloxane Composites. Advanced Optical Materials 5(10). 1700237. [24] Chen, Dongzhi, Wei Sun, Chenxi Qian, Annabelle PY Wong, Laura M Reyes & Geof- frey A Ozin. 2017. Uv-blocking photoluminescent silicon nanocrystal/polydimethylsiloxane composites. Advanced Optical Materials 5(10). 1700237. [25] Chen, Francis F. 1984. Introduction to plasma physics and controlled fusion, vol. 1. Springer. [26] Cheng, Kai-Yuan, Rebecca Anthony, Uwe R Kortshagen & Russell J Holmes. 2011. High- eciency silicon nanocrystal light-emitting devices. Nano letters 11(5). 1952–1956. 91 [27] Cheng, Xiaoyu, Stuart B Lowe, Peter J Reece & J Justin Gooding. 2014. Colloidal silicon quantum dots: from preparation to the modification of self-assembled monolayers (SAMs) for bio-applications. Chemical Society Reviews 43(8). 2680–2700. [28] Chiang, Wei-Hung, Matthew Cochey, R Craig Virnelson & R Mohan Sankaran. 2007. Nonlithographic fabrication of surface-enhanced Raman scattering substrates using a rastered atmospheric-pressure microplasma source. Applied Physics Letters 91(2). 21501. doi: 10.1063/1.2757152. [29] Chiang, Wei-Hung, Carolyn Richmonds & R Mohan Sankaran. 2010. Continuous-flow, atmospheric-pressure microplasmas: a versatile source for metal nanoparticle synthesis in the gas or liquid phase. Plasma Sources Science and Technology 19(3). 34011. [30] Chung, Jun Young, Adam J Nolte & Christopher M Staord. 2011. Surface wrinkling: a versatile platform for measuring thin-film properties. Advanced Materials 23(3). 349–368. [31] Coman, Ciprian D & Michel Destrade. 2008. Asymptotic results for bifurcations in pure bending of rubber blocks. Quarterly Journal of Mechanics and Applied Mathematics 61(3). 395–414. doi:10.1093/qjmam/hbn009. [32] Conibeer, Gavin, Martin Green, Richard Corkish, Young Cho, Eun Chel Cho, Chu Wei Jiang, Thipwan Fangsuwannarak, Edwin Pink, Yidan Huang, Tom Puzzer, Thorsten Trupke, Bryce Richards, Avi Shalav & Kuo lung Lin. 2006. Silicon nanostructures for third generation photovoltaic solar cells. Thin Solid Films 511-512. 654–662. doi:10.1016/j.tsf.2005.12.119. [33] Cullis, A. G. & L. T. Canham. 1991. Visible light emission due to quantum size eects in highly porous crystalline silicon. Nature 353(6342). 335–338. doi:10.1038/353335a0. [34] Cullis, A G & L T Canham. 1991. Visible light emission due to quantum size eects in highly porous crystalline silicon. Nature 353(6342). 335. [35] Daniel, Marie-Christine & Didier Astruc. 2004. Gold Nanoparticles: Assembly, Supramolecular Chemistry, Quantum-Size-Related Properties, and Applications toward Biology, Catalysis, and Nanotechnology. Chemical Reviews 104(1). 293–346. doi: 10.1021/cr030698+. [36] Dasog, Mita, Julian Kehrle, Bernhard Rieger & Jonathan G C Veinot. 2016. Silicon nanocrys- tals and silicon-polymer hybrids: synthesis, surface engineering, and applications. Ange- wandte Chemie International Edition 55(7). 2322–2339. [37] Dasog, Mita & Jonathan G. C. Veinot. 2014. Tuning silicon quantum dot luminescence via surface groups. Physica Status Solidi (B) 251(11). 2216–2220. doi:10.1002/pssb.201400026. [38] Daugherty, JE & DB Graves. 1993. Particulate temperature in radio frequency glow dis- charges. Journal of Vacuum Science & Technology A: Vacuum, Surfaces, and Films 11(4). 1126–1131. [39] Delerue, C, M Lannoo & G Allan. 1993. Theory of the luminescence of porous silicon. Journal of luminescence 57(1-6). 249–256. 92 [40] Destrade, M, M D Gilchrist, J A Motherway & J G Murphy. 2010. Bimodular rubber buckles early in bending. Mechanics of Materials 42(4). 469–476. doi:10.1016/j.mechmat.2009.11. 018. [41] Destrade, M, R W Ogden, I Sgura & L Vergori. 2014. Straightening: existence, uniqueness and stability. Proceedings of the Royal Society A 470. 20130709. doi:10.1098/rspa.2013. 0709. [42] Ding, Zhifeng, Bernadette M Quinn, Santosh K Haram, Lindsay E Pell, Brian A Korgel & Allen J Bard. 2002. Electrochemistry and electrogenerated chemiluminescence from silicon nanocrystal quantum dots. Science 296(5571). 1293–1297. [43] Dreaden, Erik C, Alaaldin M Alkilany, Xiaohua Huang, Catherine J Murphy & Mostafa A El-Sayed. 2012. The golden age: gold nanoparticles for biomedicine. Chemical Society Reviews 41(7). 2740–2779. doi:10.1039/C1CS15237H. [44] Dryburgh, G & R W Ogden. 1999. Bifurcation of an elastic surface-coated incompressible isotropic elastic block subject to bending. Zeitschrift fur Angewandte Mathematik und Physik 50(5). 822–838. doi:10.1007/s000330050180. [45] Eustis, Susie & Mostafa A El-Sayed. 2006. Why gold nanoparticles are more precious than pretty gold: noble metal surface plasmon resonance and its enhancement of the radiative and nonradiative properties of nanocrystals of dierent shapes. Chemical Society Reviews 35(3). 209–217. [46] Fallon, Erin, RC Fedderson, James Kurtzleben, Maurice A Lee, Susan Rowchette-Crawley & A Reader’s. 2014. Department of liberal arts. Course Booklet 2014-15 126. 90. [47] Fetni, Raouf, Regen Drouin, Nicole Lemieux, Paul-Emil Messier & Claude-Lise Richer. 1991. Simultaneous visualization of chromosome bands and hybridization signal using colloidal-gold labeling in electron microscopy. Proceedings of the National Academy of Sciences 88(23). 10916–10920. [48] Firth, Peter & Zachary C Holman. 2018. Aerosol Impaction-Driven Assembly System for the Production of Uniform Nanoparticle Thin Films with Independently Tunable Thickness and Porosity. ACS Applied Nano Materials 1(8). 4351–4357. [49] FRENS, G. 1973. Controlled Nucleation for the Regulation of the Particle Size in Monodis- perse Gold Suspensions. Nature Physical Science 241. 20. [50] Fujita, S & N Sugiyama. 1999. Visible light-emitting devices with Schottky contacts on an ultrathin amorphous silicon layer containing silicon nanocrystals. Applied physics letters 74(2). 308–310. [51] Gent, A N. 2005. Elastic instabilities in rubber. International Journal of Non-Linear Mechanics 40(2-3). 165–175. doi:10.1016/j.ijnonlinmec.2004.05.006. [52] Gent, A N & I S Cho. 1999. Surface Instabilities in Compressed or Bent Rubber Blocks. Rubber Chemistry and Technology 72(2). 253–262. doi:10.5254/1.3538798. 93 [53] Gerberich, W W, W M Mook, C R Perrey, C B Carter, M I Baskes, R Mukherjee, A Gidwani, J Heberlein, P H McMurry & S L Girshick. 2003. Superhard silicon nanospheres. Journal of the Mechanics and Physics of Solids 51(6). 979–992. doi:10.1016/S0022-5096(03)00018-8. [54] Giersig, Michael & Paul Mulvaney. 1993. Preparation of ordered colloid monolayers by electrophoretic deposition. Langmuir 9(12). 3408–3413. doi:10.1021/la00036a014. [55] Gorla, CR, S Liang, GS Tompa, WE Mayo & Yicheng Lu. 1997. Silicon and germanium nanoparticle formation in an inductively coupled plasma reactor. Journal of Vacuum Science & Technology A: Vacuum, Surfaces, and Films 15(3). 860–864. [56] Graetz, Jason, C C Ahn, Rachid Yazami & Brent Fultz. 2003. Highly reversible lithium storage in nanostructured silicon. Electrochemical and Solid-State Letters 6(9). A194–A197. [57] Greer, Julia R, Warren C Oliver & William D Nix. 2005. Size dependence of mechanical properties of gold at the micron scale in the absence of strain gradients. Acta Materialia 53(6). 1821–1830. doi:10.1016/j.actamat.2004.12.031. [58] Gresback, Ryan, Zachary Holman & Uwe Kortshagen. 2007. Nonthermal plasma synthesis of size-controlled, monodisperse, freestanding germanium nanocrystals. Applied Physics Letters 91(9). 93119. [59] Gross, Karlis A & Saeed Saber-Samandari. 2009. Revealing mechanical properties of a suspension plasma sprayed coating with nanoindentation. Surface and Coatings Technology 203(20-21). 2995–2999. doi:10.1016/j.surfcoat.2009.03.007. [60] Gupta, Rakesh & Beena Rai. 2017. Eect of size and surface charge of gold nanoparticles on their skin permeability: a molecular dynamics study. Scientific Reports 7. 45292. [61] Halas, Naomi J, Surbhi Lal, Wei-Shun Chang, Stephan Link & Peter Nordlander. 2011. Plasmons in Strongly Coupled Metallic Nanostructures. Chemical Reviews 111(6). 3913– 3961. doi:10.1021/cr200061k. [62] Han, By Xiaodong, Kun Zheng, Yuefei Zhang, Xiaona Zhang, Ze Zhang & Zhong Lin Wang. 2007. Low-Temperature In Situ Large-Strain Plasticity of Silicon Nanowires ** 2112–2118. doi:10.1002/adma.200602705. [63] Heitsch, Andrew T, Colin M Hessel, Vahid A Akhavan & Brian A Korgel. 2009. Colloidal Silicon Nanorod Synthesis. Nano Letters 9(8). 3042–3047. doi:10.1021/nl901520t. [64] Hessel, Colin M, Eric J Henderson & Jonathan G C Veinot. 2006. Hydrogen silsesquioxane: a molecular precursor for nanocrystalline Si- SiO2 composites and freestanding hydride- surface-terminated silicon nanoparticles. Chemistry of materials 18(26). 6139–6146. [65] Hill, Samantha K E, Ryan Connell, Colin Peterson, Jon Hollinger, Marc A Hillmyer, Uwe Kortshagen & Vivian E Ferry. 2019. Silicon Quantum Dot–Poly(methyl methacrylate) Nanocomposites with Reduced Light Scattering for Luminescent Solar Concentrators. ACS Photonics 6(1). 170–180. doi:10.1021/acsphotonics.8b01346. 94 [66] Hill, Samantha KE, Ryan Connell, Colin Peterson, Jon Hollinger, Marc A Hillmyer, Uwe Kortshagen & Vivian E Ferry. 2018. Silicon quantum dot–poly (methyl methacrylate) nanocomposites with reduced light scattering for luminescent solar concentrators. ACS Photonics 6(1). 170–180. [67] Hirasawa, Makoto, Takaaki Orii & Takafumi Seto. 2006. Size-dependent crystallization of si nanoparticles. Applied physics letters 88(9). 093119. [68] Holman, Z. C. & U. R. Kortshagen. 2010. A flexible method for depositing dense nanocrystal thin films: Impaction of germanium nanocrystals. Nanotechnology 21(33). doi:10.1088/ 0957-4484/21/33/335302. [69] Holman, Zachary C & Uwe R Kortshagen. 2010. A flexible method for depositing dense impaction of germanium nanocrystals. Nanotechnology 21(33). nanocrystal thin films: 335302. [70] Horisberger, Marc & Jacqueline Rosset. 1977. Colloidal gold, a useful marker for transmis- sion and scanning electron microscopy. Journal of Histochemistry & Cytochemistry 25(4). 295–305. [71] Hsu, Su-Wen, Kathy On & Andrea R Tao. 2011. Localized Surface Plasmon Resonances of Anisotropic Semiconductor Nanocrystals. Journal of the American Chemical Society 133(47). 19072–19075. doi:10.1021/ja2089876. [72] Huang, Xiaohua, Prashant K Jain, Ivan H El-Sayed & Mostafa A El-Sayed. 2007. Gold nanoparticles: interesting optical properties and recent applications in cancer diagnostics and therapy. Nanomedicine 2(5). 681–693. doi:10.2217/17435889.2.5.681. [76] [73] Huang, Yunjie, Di Li & Jinghong Li. 2004. -Cyclodextrin controlled assembling nanostruc- tures from gold nanoparticles to gold nanowires. Chemical Physics Letters 389(1). 14–18. doi:https://doi.org/10.1016/j.cplett.2004.03.019. Iqbal, Z & S Veprek. 1982. Raman scattering from hydrogenated microcrystalline and amorphous silicon. Journal of Physics C: Solid State Physics 15(2). 377. Iyer, Subramanian S & Y-H Xie. 1993. Light emission from silicon. Science 260(5104). 40–46. Jariwala, Bhavin N, Oliver S Dewey, Paul Stradins, Cristian V Ciobanu & Sumit Agarwal. 2011. In situ gas-phase hydrosilylation of plasma-synthesized silicon nanocrystals. ACS applied materials & interfaces 3(8). 3033–3041. Jiang, Chaoyang, Srikanth Singamaneni, Emily Merrick & Vladimir V Tsukruk. 2006. Com- plex Buckling Instability Patterns of Nanomembranes with Encapsulated Gold Nanoparticle Arrays. Nano Letters 6(10). 2254–2259. doi:10.1021/nl061630n. Johnston, ID, DK McCluskey, CKL Tan & MC Tracey. 2014. Mechanical characterization of bulk sylgard 184 for microfluidics and microengineering. Journal of Micromechanics and Microengineering 24(3). 035017. [77] [74] [75] [78] 95 [79] [80] Jurbergs, David, Elena Rogojina, Lorenzo Mangolini & Uwe Kortshagen. 2006. Silicon nanocrystals with ensemble quantum yields exceeding 60%. Applied Physics Letters 88(23). 233116. Jurbergs, David, Elena Rogojina, Lorenzo Mangolini & Uwe Kortshagen. 2006. Silicon nanocrystals with ensemble quantum yields exceeding 60%. Applied Physics Letters 88(23). 233116. doi:10.1063/1.2210788. [81] Khan, Arshad, Pampa Modak, Manish Joshi, P Khandare, Amruta Koli, Alka Gupta, S Anand & B K Sapra. 2014. Generation of high-concentration nanoparticles using glowing wire tech- nique. Journal of Nanoparticle Research 16(12). 2776. doi:10.1007/s11051-014-2776-5. [82] Kilymis, D, C Gérard, J Amodeo, U V Waghmare & L Pizzagalli. 2018. Uniaxial compression of silicon nanoparticles: An atomistic study on the shape and size eects. Acta Materialia 158. 155–166. [83] Kortshagen, Uwe R, R Mohan Sankaran, Rui N Pereira, Steven L Girshick, Jeslin J Wu & Eray S Aydil. 2016. Nonthermal plasma synthesis of nanocrystals: fundamental principles, materials, and applications. Chemical reviews 116(18). 11061–11127. [84] Kramer, N. J., R. J. Anthony, M. Mamunuru, E. S. Aydil & U. R. Kortshagen. 2014. Plasma- induced crystallization of silicon nanoparticles. Journal of Physics D: Applied Physics 47(7). doi:10.1088/0022-3727/47/7/075202. [85] Kramer, NJ, ES Aydil & UR Kortshagen. 2015. Requirements for plasma synthesis of nanocrystals at atmospheric pressures. Journal of Physics D: Applied Physics 48(3). 035205. [86] Kumar, Ajay, Seungkoo Kang, Carlos Larriba-Andaluz, Hui Ouyang, Christopher J Hogan & R Mohan Sankaran. 2014. Ligand-free Ni nanocluster formation at atmospheric pressure via rapid quenching in a microplasma process. Nanotechnology 25(38). 385601. [87] Kumar, Ajay, Pin Ann Lin, Albert Xue, Boyi Hao, Yoke Khin Yap & R Mohan Sankaran. 2013. Formation of nanodiamonds at near-ambient conditions via microplasma dissociation of ethanol vapour. Nature communications 4. 2618. [88] Li, Cheng-Hui, Chao Wang, Christoph Keplinger, Jing-Lin Zuo, Lihua Jin, Yang Sun, Peng Zheng, Yi Cao, Franziska Lissel & Christian Linder. 2016. A highly stretchable autonomous self-healing elastomer. Nature chemistry 8(6). 618. [89] Li, Hongbo, Kaifeng Wu, Jaehoon Lim, Hyung-Jun Song & Victor I Klimov. 2016. Doctor- blade deposition of quantum dots onto standard window glass for low-loss large-area lumi- nescent solar concentrators. Nature Energy 1(12). 16157. [90] Li, N, P Zhao & D Astruc. 2014. Anisotropic gold nanoparticles: synthesis, properties, applications, and toxicity. Angew Chem Int Ed Engl 53(7). 1756–1789. doi:10.1002/anie. 201300441. [91] Li, Xuegeng, Yuanqing He, Suddha S Talukdar & Mark T Swihart. 2003. Process for preparing macroscopic quantities of brightly photoluminescent silicon nanoparticles with emission spanning the visible spectrum. Langmuir 19(20). 8490–8496. 96 [92] Lieberman, Michael A & Alan J Lichtenberg. 2005. Principles of plasma discharges and materials processing. John Wiley & Sons. [93] Lin, Yung-Tsan & Jung-Hua Chou. 2016. Fabricating translucent polydimethylsiloxane (PDMS) super-hydrophobic surface greenly by facile water-dissolved fillers. Journal of Ad- hesion Science and Technology 30(12). 1310–1318. doi:10.1080/01694243.2016.1145314. [94] Littau, KA, PJ Szajowski, AJ Muller, AR Kortan & LE Brus. 1993. A luminescent sili- con nanocrystal colloid via a high-temperature aerosol reaction. The Journal of Physical Chemistry 97(6). 1224–1230. [95] Liu, Chin-Yi, Zachary C Holman & Uwe R Kortshagen. 2008. Hybrid solar cells from P3HT and silicon nanocrystals. Nano letters 9(1). 449–452. [96] Liu, Miao, Jianren Sun, Ying Sun, Christopher Bock & Quanfang Chen. 2009. Thickness- dependent mechanical properties of polydimethylsiloxane membranes. Journal of microme- chanics and microengineering 19(3). 035028. [97] Liu, Nanguo, Bradley S Prall & Victor I Klimov. 2006. Hybrid Gold/Silica/Nanocrystal- Quantum-Dot Superstructures: Synthesis and Analysis of Semiconductor-Metal Inter- actions. doi: 10.1021/ja0660296. the American Chemical Society 128(48). 15362–15363. Journal of [98] Liu, Xin & Mark T Swihart. 2014. Heavily-doped colloidal semiconductor and metal oxide nanocrystals: an emerging new class of plasmonic nanomaterials. Chemical Society Reviews 43(11). 3908–3920. [99] Liu, Yingyi, Yi Sun, Hao Yan, Xiyu Liu, Wei Zhang, Zhuo Wang & Xingyu Jiang. 2012. Electrospun fiber template for replica molding of microtopographical neural growth guid- ance. Small 8(5). 676–681. [100] Lockwood, AJ & BJ Inkson. 2008. In situ tem nanoindentation and deformation of si- nanoparticle clusters. Journal of Physics D: Applied Physics 42(3). 035410. [101] Lohse, Samuel E & Catherine J Murphy. 2013. The Quest for Shape Control: A History of Gold Nanorod Synthesis. Chemistry of materials 25(8). 1250–1261. doi: 10.1021/cm303708p. [102] Lu, Conghua, Ingo Dönch, Marc Nolte & Andreas Fery. 2006. Au Nanoparticle-based Multilayer Ultrathin Films with Covalently Linked Nanostructures: Spraying Layer-by-layer Assembly and Mechanical Property Characterization. Chemistry of Materials 18(26). 6204– 6210. doi:10.1021/cm061759y. [103] Maier-flaig, Florian, Julia Rinck, Moritz Stephan, Tobias Bocksrocker, Michael Bruns, Christian Ku, Annie K Powell, A Ozin & Uli Lemmer. 2013. Multicolor Silicon Light- Emitting Diodes ( SiLEDs ) 1–6. doi:10.1021/nl3038689. [104] Mandal, Rajib & Rebecca J. Anthony. 2016. Aging of Silicon Nanocrystals on Elastomer Substrates: Photoluminescence Eects. ACS Applied Materials and Interfaces 8(51). 35479– 35484. doi:10.1021/acsami.6b10155. 97 [105] Mangolini, L, D Jurbergs, E Rogojina & U Kortshagen. 2006. Plasma synthesis and liquid- phase surface passivation of brightly luminescent si nanocrystals. Journal of luminescence 121(2). 327–334. [106] Mangolini, L, D Jurbergs, E Rogojina & U Kortshagen. 2006. Plasma synthesis and liquid- phase surface passivation of brightly luminescent Si nanocrystals. Journal of Luminescence 121(2). 327–334. [107] Mangolini, L, E Thimsen & U Kortshagen. 2005. High-Yield Plasma Synthesis of Lumines- cent Silicon Nanocrystals. Nano Letters 5(4). 655–659. doi:10.1021/nl050066y. [108] Mangolini, Lorenzo. 2013. Synthesis, properties, and applications of silicon nanocrys- tals. Journal of Vacuum Science & Technology B, Nanotechnology and Microelectronics: Materials, Processing, Measurement, and Phenomena 31(2). 020801. [109] Mangolini, Lorenzo & Uwe Kortshagen. 2009. Selective nanoparticle heating: another form of nonequilibrium in dusty plasmas. Physical review E 79(2). 026405. [110] Mark, James E. 2004. Some interesting things about polysiloxanes. Accounts of chemical research 37(12). 946–953. [111] McDonald, J. Cooper, David C. Duy, Janelle R. Anderson, Daniel T. Chiu, Hongkai Wu, Olivier J. A. Schueller & George M. Whitesides. 2000. Fabrication of microfluidic systems in poly(dimethylsiloxane). ELECTROPHORESIS 21(1). 27–40. doi:10.1002/(SICI) 1522-2683(20000101)21:1<27::AID-ELPS27>3.0.CO;2-C. [112] McDonald, J Cooper & George M Whitesides. 2002. Poly(dimethylsiloxane) as a Material for Fabricating Microfluidic Devices. Accounts of chemical research 35(7). 491–499. doi: 10.1021/ar010110q. [113] McVey, Benjamin F P & Richard D Tilley. 2014. Solution synthesis, optical properties, and bioimaging applications of silicon nanocrystals. Accounts of chemical research 47(10). 3045–3051. [114] Meier, Cedrik, Andreas Gondorf, Stephan Lüttjohann, Axel Lorke & Hartmut Wiggers. 2007. Silicon nanoparticles: Absorption, emission, and the nature of the electronic bandgap. Journal of applied physics 101(10). 103112. [115] Meinardi, Francesco, Samantha Ehrenberg, Lorena Dhamo, Francesco Carulli, Michele Mauri, Francesco Bruni, Roberto Simonutti, Uwe Kortshagen & Sergio Brovelli. 2017. Highly ecient luminescent solar concentrators based on earth-abundant indirect-bandgap silicon quantum dots. Nature Photonics 11(3). 177. [116] Meinardi, Francesco, Hunter McDaniel, Francesco Carulli, Annalisa Colombo, Kirill A Velizhanin, Nikolay S Makarov, Roberto Simonutti, Victor I Klimov & Sergio Brovelli. 2015. Highly ecient large-area colourless luminescent solar concentrators using heavy- metal-free colloidal quantum dots. Nature Nanotechnology 10(10). 878. 98 [117] Meuller, Bengt O, Maria E Messing, David L J Engberg, Anna M Jansson, Linda I M Johansson, Susanne M Norlén, Nina Tureson & Knut Deppert. 2012. Review of Spark Dis- charge Generators for Production of Nanoparticle Aerosols. Aerosol Science and Technology 46(11). 1256–1270. doi:10.1080/02786826.2012.705448. [118] MichelleáGrandin, H. 2007. Micro-well arrays for 3D shape control and high resolution analysis of single cells. Lab on a Chip 7(8). 1074–1077. [119] Mook, W. M., J. D. Nowak, C. R. Perrey, C. B. Carter, R. Mukherjee, S. L. Girshick, P. H. McMurry & W. W. Gerberich. 2007. Compressive stress eects on nanoparticle modulus and fracture. Physical Review B - Condensed Matter and Materials Physics 75(21). 1–10. doi:10.1103/PhysRevB.75.214112. [120] Mori, T & K Tanaka. 1973. Average stress in matrix and average elastic energy of materials with misfitting inclusions. Acta Metallurgica 21(5). 571–574. doi:https://doi.org/10.1016/ 0001-6160(73)90064-3. [121] Murphy, Catherine J, Tapan K Sau, Anand M Gole, Christopher J Orendor, Jinxin Gao, Linfeng Gou, Simona E Hunyadi & Tan Li. 2005. Anisotropic Metal Nanoparticles : Synthesis , Assembly , and Optical Applications 13857–13870. doi:10.1021/jp0516846. [122] Murray, Christopher B, a CR Kagan & MG Bawendi. 2000. Synthesis and characterization of monodisperse nanocrystals and close-packed nanocrystal assemblies. Annual review of materials science 30(1). 545–610. [123] Mussard, Windy, Nasreddine Kebir, Irène Kriegel, Marc Estève & Vincent Semetey. 2011. Facile and Ecient Control of Bioadhesion on Poly(dimethylsiloxane) by Using a Biomimetic Approach. Angewandte Chemie International Edition 50(46). 10871–10874. doi:10.1002/anie.201101029. [124] Nava, Giorgio, Francesco Fumagalli, Stefanie Neutzner & Fabio Di Fonzo. 2018. Large area porous 1D photonic crystals comprising silicon hierarchical nanostructures grown by plasma- assisted, nanoparticle jet deposition. Nanotechnology 29(46). doi:10.1088/1361-6528/ aade21. [125] Nolte, Adam J, Michael F Rubner & Robert E Cohen. 2005. Determining the Young’s Modu- lus of Polyelectrolyte Multilayer Films via Stress-Induced Mechanical Buckling Instabilities. Macromolecules 38(13). 5367–5370. doi:10.1021/ma0507950. [126] Odegard, G M, T C Clancy & T S Gates. 2005. Modeling of the mechanical properties of nanoparticle/polymer composites. Polymer 46(2). 553–562. [127] Onischuk, AA, AI Levykin, VP Strunin, MA Ushakova, RI Samoilova, KK Sabelfeld & VN Panfilov. 2000. Aerosol formation under heterogeneous/homogeneous thermal decom- position of silane: experiment and numerical modeling. Journal of aerosol science 31(8). 879–906. 99 [128] Ossadnik, Ch, S Vepek & I Gregora. 1999. Applicability of Raman scattering for the characterization of nanocrystalline silicon. Thin Solid Films 337(1). 148–151. doi:https: //doi.org/10.1016/S0040-6090(98)01175-4. [129] Ostraat, Michele L, Jan W De Blauwe, Martin L Green, L Douglas Bell, Harry A Atwater & Richard C Flagan. 2001. Ultraclean two-stage aerosol reactor for production of oxide- passivated silicon nanoparticles for novel memory devices. Journal of The Electrochemical Society 148(5). G265–G270. [130] Ozgul, Yasar-Inceoglu, Lopez Thomas, Farshihagro Ebrahim & Mangolini Lorenzo. 2012. Silicon nanocrystal production through non-thermal plasma synthesis: a comparative study between silicon tetrachloride and silane precursors. Nanotechnology 23(25). 255604. [131] Park, Junyong, Shuodao Wang, Ming Li, Changui Ahn, Jerome K Hyun, Dong Seok Kim, Do Kyung Kim, John A Rogers, Yonggang Huang & Seokwoo Jeon. 2012. Three-dimensional nanonetworks for giant stretchability in dielectrics and conductors. Nature communications 3. 916. [132] Peineke, C, M B Attoui & A Schmidt-Ott. 2006. Using a glowing wire generator for production of charged, uniformly sized nanoparticles at high concentrations. Journal of Aerosol Science 37(12). 1651–1661. [133] Pereira, RN, DJ Rowe, RJ Anthony & U Kortshagen. 2012. Freestanding silicon nanocrystals with extremely low defect content. Physical Review B 86(8). 085449. [134] Petcu, M C, A C Bronneberg, A Sarkar, M A Blauw, M Creatore & M C M van de Sanden. 2008. A capacitive probe with shaped probe bias for ion flux measurements in depositing plasmas. Review of Scientific Instruments 79(11). 115104. doi:10.1063/1.3020709. [135] Pettigrew, Katherine A, Qi Liu, Philip P Power & Susan M Kauzlarich. 2003. Solution synthesis of alkyl-and alkyl/alkoxy-capped silicon nanoparticles via oxidation of mg2si. Chemistry of materials 15(21). 4005–4011. [136] Plasmas, Dusty. 1999. Physics, chemistry, and technological impacts in plasma processing, edited by a. bouchoule. [137] Pratontep, S, S J Carroll, C Xirouchaki, M Streun & R E Palmer. 2005. Size-selected cluster beam source based on radio frequency magnetron plasma sputtering and gas condensation. Review of Scientific Instruments 76(4). 45103. [138] Pritchard, Robyn H, Pascal Lava, Dimitri Debruyne & Eugene M Terentjev. 2013. Precise determination of the Poisson ratio in soft materials with 2D digital image correlation. Soft Matter 9(26). 6037–6045. doi:10.1039/C3SM50901J. [139] Prokes, S M & O J Glembocki. 1993. Light emission properties of porous silicon. Materials chemistry and physics 35(1). 1–10. [140] Proot, J P, C Delerue & G Allan. 1992. Electronic structure and optical properties of silicon crystallites: Application to porous silicon. Applied Physics Letters 61(16). 1948–1950. 100 [141] Rafiee, Mehran, Subhash Chandra, Hind Ahmed & Sarah J McCormack. 2019. An overview of various configurations of Luminescent Solar Concentrators for photovoltaic applications. Optical Materials 91. 212–227. doi:https://doi.org/10.1016/j.optmat.2019.01.007. [142] Rao, N, S Girshick, J Heberlein, P McMurry, S Jones, D Hansen & B Micheel. 1995. Nanoparticle formation using a plasma expansion process. Plasma Chemistry and Plasma Processing 15(4). 581–606. [143] Roccabianca, S, M Gei & D Bigoni. 2010. Plane strain bifurcations of elastic layered structures subject to finite bending: Theory versus experiments. IMA Journal of Applied Mathematics (Institute of Mathematics and Its Applications) 75(4). 525–548. doi:10.1093/ imamat/hxq020. [144] Roccabianca, Sara, Davide Bigoni & Massimiliano Gei. 2011. Long wavelength bifurcations and multiple neutral axes of elastic layered structures subject to finite bending. Journal of Mechanics of Materials and Structures 6(1). 511–527. [145] Rogers, John A & Ralph G Nuzzo. 2005. Recent progress in soft lithography. Materials today 8(2). 50–56. [146] Rose, M J, J Hajto, P G Lecomber, S M Gage, W K Choi, A J Snell & A E Owen. 1989. Amorphous silicon analogue memory devices. Journal of Non-Crystalline Solids 115(1-3). 168–170. doi:10.1016/0022-3093(89)90394-3. [147] Rycenga, Matthew, Claire M Cobley, Jie Zeng, Weiyang Li, Christine H Moran, Qiang Zhang, Dong Qin & Younan Xia. 2011. Controlling the synthesis and assembly of silver nanostructures for plasmonic applications. Chemical reviews 111(6). 3669–3712. [148] Sackmann, Eric K, Anna L Fulton & David J Beebe. 2014. The present and future role of microfluidics in biomedical research. Nature 507. 181+. [149] Sadeghian, H, C K Yang, J F L Goosen, E Van Der Drift, A Bossche, P J French & F Van Keulen. 2009. Characterizing size-dependent eective elastic modulus of silicon nanocantilevers using electrostatic pull-in instability. Applied Physics Letters 94(22). 1–4. doi:10.1063/1.3148774. [150] Sangghaleh, Fatemeh, Ilya Sychugov, Zhenyu Yang, Jonathan G C Veinot & Jan Linnros. 2015. Near-Unity Internal Quantum E ciency of Luminescent Silicon Nanocrystals with Ligand Passivation. ACS Nano 9(7). doi:10.1021/acsnano.5b01717. [151] Sekitani, Tsuyoshi, Hiroyoshi Nakajima, Hiroki Maeda, Takanori Fukushima, Takuzo Stretchable active-matrix organic light- Aida, Kenji Hata & Takao Someya. 2009. emitting diode display using printable elastic conductors. Nature Materials 8. 494. doi: 10.1038/nmat2459https://www.nature.com/articles/nmat2459#supplementary-information. [152] Sekitani, Tsuyoshi, Tomoyuki Yokota, Ute Zschieschang, Hagen Klauk, Siegfried Bauer, Ken Takeuchi, Makoto Takamiya, Takayasu Sakurai & Takao Someya. 2009. Organic Nonvolatile Memory Transistors for Flexible Sensor Arrays. Science 326(December). 1516–1520. 101 [153] Sigaeva, Taisiya, Robert Mangan, Luigi Vergori, Michel Destrade & Les Sudak. 2018. Wrinkles and creases in the bending, unbending and eversion of soft sectors. Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences 474. 2212. doi: 10.1098/rspa.2017.0827. [154] Sinha, Mayank, Alborz Izadi, Rebecca Joy Anthony & Sara Roccabianca. 2019. A novel approach to finding mechanical properties of nanocrystal layers. Nanoscale . [155] Staord, Christopher M., Christopher Harrison, Kathryn L. Beers, Alamgir Karim, Eric J. Amis, Mark R. Vanlandingham, Ho Cheol Kim, Willi Volksen, Robert D. Miller & Eva E. Simonyi. 2004. A buckling-based metrology for measuring the elastic moduli of polymeric thin films. Nature Materials 3(8). 545–550. doi:10.1038/nmat1175. [156] Staord, Christopher M, Bryan D Vogt, Christopher Harrison, Duangrut Julthongpiput & Rui Huang. 2006. Elastic moduli of ultrathin amorphous polymer films. Macromolecules 39(15). 5095–5099. [157] Staord, Christopher M, Bryan D Vogt, Christopher Harrison, Duangrut Julthongpiput & Rui Huang. 2006. Elastic Moduli of Ultrathin Amorphous Polymer Films. Macromolecules 39(15). 5095–5099. doi:10.1021/ma060790i. [158] Strobel, Reto & Sotiris E Pratsinis. 2007. Flame aerosol synthesis of smart nanostructured materials. Journal of Materials Chemistry 17(45). 4743–4756. [159] Suchomel, Petr, Libor Kvitek, Robert Prucek, Ales Panacek, Avik Halder, Stefan Va- jda & Radek Zboril. 2018. Simple size-controlled synthesis of Au nanoparticles and their size-dependent catalytic activity. doi:10.1038/ s41598-018-22976-5. Scientific Reports 8(1). 4589. [160] Sun, Wei, Chenxi Qian, Xiao Sherri Cui, Liwei Wang, Muan Wei, Gilberto Casillas, Amr S Helmy & Georey A Ozin. 2016. Silicon monoxide–a convenient precursor for large scale synthesis of near infrared emitting monodisperse silicon nanocrystals. Nanoscale 8(6). 3678–3684. [161] Tabrizi, Nooshin Salman, M Ullmann, V A Vons, U Lafont & A Schmidt-Ott. 2009. Gener- ation of nanoparticles by spark discharge. Journal of Nanoparticle Research 11(2). 315. [162] Tahk, Dongha, Hong H Lee & Dahl-Young Khang. 2009. Elastic Moduli of Organic Electronic Materials by the Buckling Method. Macromolecules 42(18). 7079–7083. doi: 10.1021/ma900137k. [163] Takagahara, Toshihide & Kyozaburo Takeda. 1992. Theory of the quantum confinement eect on excitons in quantum dots of indirect-gap materials. Physical Review B 46(23). 15578. [164] Talapin, Dmitri V, Jong-Soo Lee, Maksym V Kovalenko & Elena V Shevchenko. 2009. Prospects of colloidal nanocrystals for electronic and optoelectronic applications. Chemical reviews 110(1). 389–458. 102 [165] Thakur, Vijay Kumar, Guoqiang Ding, Jan Ma, Pooi See Lee & Xuehong Lu. 2012. Hy- brid materials and polymer electrolytes for electrochromic device applications. Advanced materials 24(30). 4071–4096. [166] Tian, Bozhi, Xiaolin Zheng, Thomas J Kempa, Ying Fang, Nanfang Yu, Guihua Yu, Jinlin Huang & Charles M Lieber. 2007. Coaxial silicon nanowires as solar cells and nanoelectronic power sources 449(October). 885–890. doi:10.1038/nature06181. [167] Timmerman, D., I. Izeddin, P. Stallinga, I. N. Yassievich & T. Gregorkiewicz. 2008. Space- separated quantum cutting with silicon nanocrystals for photovoltaic applications. Nature Photonics 2(2). 105–109. doi:10.1038/nphoton.2007.279. [168] Torres, Jessica M, Christopher M Staord & Bryan D Vogt. 2009. Elastic Modulus of Amorphous Polymer Thin Films: Relationship to the Glass Transition Temperature. ACS Nano 3(9). 2677–2685. doi:10.1021/nn9006847. [169] Triantafyllidis, N. 1980. Bifurcation phenomena in pure bending. Journal of the Mechanics and Physics of Solids 28(3-4). 221–245. doi:10.1016/0022-5096(80)90005-8. [170] Tsakalakos, L., J. Balch, J. Fronheiser, B. A. Korevaar, O. Sulima & J. Rand. 2007. Silicon nanowire solar cells. Applied Physics Letters 91(23). 233117. doi:10.1063/1.2821113. [171] Turkevich, John, Peter Cooper Stevenson & James Hillier. 1951. A study of the nucleation and growth processes in the synthesis of colloidal gold. Discussions of the Faraday Society 11(0). 55–75. doi:10.1039/DF9511100055. [172] Uchic, Michael D, Dennis M Dimiduk, Jerey N Florando & William D Nix. 2004. Sample dimensions influence strength and crystal plasticity. Science 305(5686). 986–989. doi: 10.1126/science.1098993. [173] Van Sickle, A R, J B Miller, C Moore, R J Anthony, U R Kortshagen & E K Hobbie. 2013. Temperature Dependent Photoluminescence of Size-purified Silicon Nanocrystals. ACS Applied Materials and Interfaces 5(10). 4233–4238. doi:10.1021/am400411a. [174] Van Sickle, Austin R, Joseph B Miller, Christopher Moore, Rebecca J Anthony, Uwe R Kortshagen & Erik K Hobbie. 2013. Temperature dependent photoluminescence of size- purified silicon nanocrystals. ACS applied materials & interfaces 5(10). 4233–4238. [175] Vanrenterghem, B, M Bele, F R Zepeda, M äala, N Hodnik & Tom Breugelmans. 2018. Cutting the Gordian Knot of electrodeposition via controlled cathodic corrosion enabling the production of supported metal nanoparticles below 5 nm. Applied Catalysis B: Environ- mental 226. 396–402. [176] Vasiliev, Igor, James R Chelikowsky & Richard M Martin. 2002. Surface oxidation eects on the optical properties of silicon nanocrystals. Physical Review B 65(12). 121302. [177] Vaterrodt, Anne, Barbara Thallinger, Kevin Daumann, Dereck Koch, Georg M Gueb- itz & Mathias Ulbricht. 2016. Antifouling and Antibacterial Multifunctional Polyzwitte- rion/Enzyme Coating on Silicone Catheter Material Prepared by Electrostatic Layer-by-Layer Assembly. Langmuir 32(5). 1347–1359. doi:10.1021/acs.langmuir.5b04303. 103 [178] Vendamme, Richard, Takuya Ohzono, Aiko Nakao, Masatsugu Shimomura & Toyoki Kuni- take. 2007. Synthesis and Micromechanical Properties of Flexible, Self-Supporting Polymer- SiO$_2$ Nanofilms. Langmuir 23(5). 2792–2799. doi:10.1021/la062084g. [179] Wang, Chuan, David Hwang, Zhibin Yu, Kuniharu Takei, Junwoo Park, Teresa Chen, Biwu Ma & Ali Javey. 2013. User-interactive electronic skin for instantaneous pressure visualization. Nature materials 12(10). 899. [180] Wang, Jiangwei, Zhi Zeng, Christopher R Weinberger, Ze Zhang, Ting Zhu & Scott X Mao. In situ atomic-scale observation of twinning-dominated deformation in nanoscale 2015. body-centred cubic tungsten. Nature Materials 14(6). 594–600. doi:10.1038/nmat4228. [181] Wang, Lihua, Pengfei Guan, Jiao Teng, Pan Liu, Dengke Chen, Weiyu Xie, Deli Kong, Shengbai Zhang, Ting Zhu, Ze Zhang, Evan Ma, Mingwei Chen & Xiaodong Han. 2017. New twinning route in face-centered cubic nanocrystalline metals. Nature Communications 8(1). 1–7. doi:10.1038/s41467-017-02393-4. [182] Wang, Lihua, Xiaodong Han, Pan Liu, Yonghai Yue, Ze Zhang & En Ma. 2010. In Situ observation of dislocation behavior in nanometer grains. Physical Review Letters 105(13). 1–4. doi:10.1103/PhysRevLett.105.135501. [183] Wang, Lihua, Jiao Teng, Pan Liu, Akihiko Hirata, En Ma, Ze Zhang, Mingwei Chen & Xi- aodong Han. 2014. Grain rotation mediated by grain boundary dislocations in nanocrystalline platinum. Nature Communications 5. 1–7. doi:10.1038/ncomms5402. [184] Wang, Lihua, Jiao Teng, Xuechao Sha, Jin Zou, Ze Zhang & Xiaodong Han. 2017. Plastic Deformation through Dislocation Saturation in Ultrasmall Pt Nanocrystals and Its in Situ Atomistic Mechanisms doi:10.1021/acs.nanolett.7b01416. [185] Wang, Zhixin. 2011. Polydimethylsiloxane mechanical properties measured by macroscopic compression and nanoindentation techniques . [186] Weber, W H & John Lambe. 1976. Luminescent greenhouse collector for solar radiation. Applied optics 15(10). 2299–2300. [187] Wilcoxon, JPs & GA Samara. 1999. Tailorable, visible light emission from silicon nanocrys- tals. Applied physics letters 74(21). 3164–3166. [188] Woodard, Austin, Lihua Xu, Alejandro A Barragan, Giorgio Nava, Bryan M Wong & Lorenzo ni. 2018. On the non-thermal plasma synthesis of nickel nanoparticles. Plasma Processes and Polymers 15(1). 1700104. doi:10.1002/ppap.201700104. [189] Yang, Zhenyu, Mita Dasog, Alexander R Dobbie, Ross Lockwood, Yanyan Zhi, Al Meldrum & Jonathan G C Veinot. 2014. Hybrid Materials: Highly Luminescent Covalently Linked Silicon Nanocrystal/Polystyrene Hybrid Functional Materials: Synthesis, Properties, and Processability (Adv. Funct. Mater. 10/2014). Advanced Functional Materials 24(10). 1344. doi:10.1002/adfm.201470064. 104 [190] Yu, H & W E Buhro. 2003. Solution–Liquid–Solid Growth of Soluble GaAs Nanowires. Advanced Materials 15(5). 416–419. doi:10.1002/adma.200390096. [191] Yue, Yonghai, Pan Liu, Ze Zhang & Xiaodong Han. 2011. Approaching the Theoretical Elastic Strain Limit in Copper Nanowires 3151–3155. [192] Zamiri, A & S De. 2011. Mechanical properties of hydroxyapatite single crystals from nanoindentation data. Journal of the Mechanical Behavior of Biomedical Materials 4(2). 146–152. doi:10.1016/j.pmrj.2014.02.014.Lumbar. [193] Zhang, Ning, Qian Deng, Yu Hong, Liming Xiong, Shi Li, Matthew Strasberg, Weiqi Yin, Yongjie Zou, Curtis R Taylor, Gregory Sawyer & Youping Chen. 2011. Deformation mechanisms in silicon nanoparticles. Journal of Applied Physics 109(6). 63534. doi: 10.1063/1.3552985. [194] Zhang, Qing, Jing-Juan Xu, Yan Liu & Hong-Yuan Chen. 2008. In-situ synthesis of poly (dimethylsiloxane)–gold nanoparticles composite films and its application in microfluidic systems. Lab on a Chip 8(2). 352–357. [195] Zhang, Xin-Xiang, Bi-Bo Xia, Hai-Ping Ye, Yu-Lu Zhang, Bo Xiao, Liang-Hong Yan, Hai-Bing Lv & Bo Jiang. 2012. One-step sol–gel preparation of PDMS–silica ORMOSILs as environment-resistant and crack-free thick antireflective coatings. Journal of Materials Chemistry 22(26). 13132–13140. [196] Zhang, Z, X D Han, Y F Zhang, K Zheng, X N Zhang, Y J Hao, X Y Guo, J Yuan & Z L Wang. 2007. Low-temperature in situ large strain plasticity of ceramic SiC nanowires and its atomic-scale mechanism. Nano Letters 7(2). 452–457. doi:10.1021/nl0627689. [197] Zhao, P, W Zheng, Y D Meng & M Nagatsu. 2013. Characteristics of high-purity Cu thin films deposited on polyimide by radio-frequency Ar/H2 atmospheric-pressure plasma jet. Journal of Applied Physics 113(12). 123301. doi:10.1063/1.4795808. [198] Zhou, Shu, Xiaodong Pi, Zhenyi Ni, Yi Ding, Yingying Jiang, Chuanhong Jin, Christophe Delerue, Deren Yang & Tomohiro Nozaki. 2015. Comparative study on the localized surface plasmon resonance of boron-and phosphorus-doped silicon nanocrystals. ACS nano 9(1). 378–386. [199] Zhou, Zhiyong, Louis Brus & Richard Friesner. 2003. Electronic structure and luminescence of 1.1-and 1.4-nm silicon nanocrystals: oxide shell versus hydrogen passivation. Nano Letters 3(2). 163–167. 105