Unique Mechanisms of Colloidal Stability Probed by Surface-Specific Vibrational Spectroscopy by Ashley N. Mapile A dissertation accepted and approved in partial fulfillment of the requirements for the degree of Doctor of Philosophy in Chemistry Dissertation Committee: Jeffrey Cina, Chair Lawrence Scatena, Advisor Carl Brozek, Advisor Cathy Wong, Core Member Daniel Steck, Institutional Representative University of Oregon Summer 2024 2 © 2024 Ashley N. Mapile This work is openly licensed via CC BY-NC 4.0 DISSERTATION ABSTRACT Ashley N. Mapile Doctor of Philosophy in Chemistry Title: Unique Mechanisms of Colloidal Stability Probed by Surface-Specific Vibrational Spectroscopy The stability of nanoparticles suspended in a solution, known as colloids, is crucial for their application in drug delivery systems, the solution processibility required for drop-casting films, and the long-term storage or transport of sensitive chemical materials. While current mechanisms for colloidal stability include implicit models of solvation – namely Derjaguin-Landau-Verwey- Overbeek (DLVO) and Flory-Huggins theories – these classical approaches neglect the role of specific solvent-surface interactions. Analyzing these surface-specific interactions in colloidal stability becomes increasingly relevant for nanosized particles, which have a highly accessible surface area compared to their bulk counterparts. This dissertation seeks to understand unconventional mechanisms of colloidal stability that are not explained by traditional theories alone, with oil-in-water emulsions and nanoparticles of metal-organic frameworks (nanoMOFs) as model materials. Leveraging the surface-specific spectroscopic technique, vibrational sum frequency scattering spectroscopy (VSFSS), this work provides a molecular-level understanding of the specific surface interactions that contribute to colloidal stability. In particular, emulsions can be stabilized by a steric layer of polymer alone, with colloidal behavior tunable by pH, electrolyte concentration, molecular weight, and temperature. These sterically-stabilized emulsions find applications in drug delivery systems that must withstand extreme physiological conditions. For bare nanoMOFs, an ordered solvation shell and solvent-metal surface binding contribute to unforeseen long-term stability in common solvents. Additionally, nanoMOFs coated with a polymeric binding agent – similar to those used in the paint industry – yield ultra-strong mixed-matrix membranes for gas separation technologies. Ultimately, this work bridges molecular interfacial chemistry with material properties, emphasizing the importance of understanding mechanisms of colloidal stability. This dissertation includes previously published co-authored material. 2 CURRICULUM VITAE NAME OF AUTHOR: Ashley N. Mapile GRADUATE AND UNDERGRADUATE SCHOOLS ATTENDED: University of Oregon, Eugene, OR University of Puget Sound, Tacoma, WA DEGREES AWARDED: Doctor of Philosophy, Physical Chemistry, 2024, University of Oregon Master of Science, Physical Chemistry, 2021, University of Oregon Bachelor of Science, Chemistry, 2019, University of Puget Sound AREAS OF SPECIAL INTEREST: Vibrational Spectroscopy Colloids and Soft Matter Interfacial Chemistry PROFESSIONAL EXPERIENCE: Co-Chair, Gordon Research Seminar in Vibrational Spectroscopy, August 2024 Associate Editor, Journal of Emerging Investigators, 2023-Present Newsletter Complier, Society for Applied Spectroscopy, 2023-Present Graduate Research Assistant, University of Oregon Department of Chemistry and Biochemistry, 2019-2024 Graduate Teaching Assistant, University of Oregon Department of Chemistry and Biochemistry, 2019-2024 Undergraduate Researcher, National Science Foundation Research Experience for Undergraduates (NSF-REU), University of Oregon, Summer 2018 Undergraduate Researcher, NSF-REU, University of Washington, Summer 2017 GRANTS, AWARDS, AND HONORS: Honorable Mention, National Science Foundation Graduate Research Fellowship Program, 2021 PUBLICATIONS: 1. Mapile, A. N.; Scatena, L.F. Bulking Up: The Impact of Polymer Sterics on Emulsion Stability. In Review. 2024 2. Mapile, A. N.; Svensson Grape, E.; Brozek, C. K. Solvation of Nanoscale Materials. Chem. Mater. 2024. https://doi.org/10.1021/acs.chemmater.4c01518. https://doi.org/10.1021/acs.chemmater.4c01518 3 3. Mapile, A. N.; LeRoy, M. A.; Fabrizio, K.; Scatena, L. F.; Brozek, C. K. The Surface of Colloidal Metal–Organic Framework Nanoparticles Revealed by Vibrational Sum Frequency Scattering Spectroscopy. ACS Nano 2024, 18 (20), 13406–13414. https://doi.org/10.1021/acsnano.4c03758. 4. Mapile, A. N.; Scatena, L. F. Stabilizing Strands: Exploring the Kinetic Stability of Polymer-Coated Emulsions with Surface Specific Spectroscopy. Colloids Surfaces A 2024, 697, 134414. https://doi.org/10.1016/j.colsurfa.2024.134414. 5. LeRoy, M. A.; Perera, A. S.; Lamichhane, S.; Mapile, A. N.; Khaliq, F.; Kadota, K.; Zhang, X.; Ha, S.; Fisher, R.; Wu, D.; Risko, C.; Brozek, C. K. Colloidal Stability and Solubility of Metal–Organic Framework Particles. Chem. Mater. 2024. https://doi.org/10.1021/acs.chemmater.3c03191. 6. Huang, J.; Marshall, C. R.; Ojha, K.; Shen, M.; Golledge, S.; Kadota, K.; McKenzie, J.; Fabrizio, K.; Mitchell, J. B.; Khaliq, F.; Davenport, A. M.; LeRoy, M. A.; Mapile, A. N.; Debela, T. T.; Twight, L. P.; Hendon, C. H.; Brozek, C. K. Giant Redox Entropy in the Intercalation vs Surface Chemistry of Nanocrystal Frameworks with Confined Pores. J. Am. Chem. Soc. 2023, 145 (11), 6257–6269. https://doi.org/10.1021/jacs.2c12846. 7. Carpenter, A. P.; Christoffersen, E. L.; Mapile, A. N.; Richmond, G. L. Assessing the Impact of Solvent Selection on Vibrational Sum-Frequency Scattering Spectroscopy Experiments. J. Phys. Chem. B 2021, 125 (12), 3216–3229. https://doi.org/10.1021/acs.jpcb.1c00188. 8. Tran, E.; Mapile, A. N.; Richmond, G. L. Peeling Back the Layers: Investigating the Effects of Polyelectrolyte Layering on Surface Structure and Stability of Oil-in-Water Nanoemulsions. J. Colloid Interface Sci 2021, 599, 706–716. https://doi.org/10.1016/j.jcis.2021.04.115. 9. Wilson, K. S.; Mapile, A. N.; Wong, C. Y. Broadband Single-Shot Transient Absorption Spectroscopy. Opt. Express, OE 2020, 28 (8), 11339–11355. https://doi.org/10.1364/OE.390938. https://doi.org/10.1021/acsnano.4c03758 https://doi.org/10.1016/j.colsurfa.2024.134414 https://doi.org/10.1021/acs.chemmater.3c03191 https://doi.org/10.1021/jacs.2c12846 https://doi.org/10.1021/acs.jpcb.1c00188 https://doi.org/10.1016/j.jcis.2021.04.115 https://doi.org/10.1364/OE.390938 4 ACKNOWLEDGMENTS I am extremely grateful to receive the support of so many passionate and dedicated individuals over the course of my chemistry career. From starting as a music major to finishing a Ph.D. it’s been quite a journey, and I couldn’t have done it without help. First and foremost, thank you to Geri Richmond for inspiring me to study at the University of Oregon and providing me with a home in the Richmond Lab. I still remember when I introduced myself to Geri at my first ACS conference in New Orleans as a scared little undergraduate unaware of my future. After an REU in the Richmond lab (and meeting Geri’s chickens), I was sold on Eugene and pursuing a graduate degree. I truly wouldn’t be here if it wasn’t for her. Geri’s efforts to promote equality in science and her ability to craft engaging presentations are two aspects that I hope to carry on in my own career. Second, thank you to my advisors Larry Scatena and Carl Brozek who took me under their wing and directed me through the difficult, winding path of graduate school. Larry, thank you for your tough conversations that challenged me to think deeper about my results and push myself as a researcher. I had a great time writing these manuscripts and flushing out the ideas with you (even despite the tough peer review process!). I’m honored to have learned from one of the best laser brains around. Carl, to think that we’ve only worked together for three years! We’ve cranked out a lot of impressive work (and more on the way!). Thank you for making the scientific process fun in a time when I was feeling discouraged and drained. Brainstorming in your office, drawing out figures, making two-week plans, analyzing data...to the regular person, these might seem like boring tasks, but our discussions helped me realize what I loved most about science: the investigative process and collaborating with others. Diamonds are made under pressure, right? Just like the perspective we wrote at the last minute! We make a great scientific team and I hope we get to work together again. Thank you to my committee members Jeff Cina, Cathy Wong, and Daniel Steck who asked difficult questions and pushed me to be the best researcher I could be. I also appreciated the advice and teachings of the UO faculty, particularly Marina Guenza, Jim Prell, Andy Marcus, Scott Hansen, Julia Widom, Mark Lonergan, Chris Hendon, Karl Reasoner, Mike Haley (and Rocky), and Amanda Cook (and Pyewacket). Thank you to the chemistry office, especially Leah O’Brien and Christi Mabinuori whose helpful smiles always brightened my day. Thank you to Priscilla 5 Lewis (and Shiloh) who made the logistics of graduate school so much easier and helped me be involved in the PURS and REU program. I am also grateful to my professors at the University of Puget Sound who inspired me to pursue graduate school and enriched my education in more ways than they will ever know. Thank you especially to Dan Burgard, Amanda Mifflin, Steven Neshyba, Eric Scharrer, Jo Crane, Luc Boisvert, Gerard Morris, and Jennifer Nelson. Hack hack chop chop! To my lab mates past, present, and future: Andrew Carpenter, Rebecca Altman, Brandon Schabes, Emma Tran, Marc Foster, Konnor Jones, Checkers Marshall, Kevin Fabrizio, Jiawei Huang, Erik Svensson Grape, Michael LeRoy, Jacob McKenzie, Audrey Davenport, Faiqa Khaliq, Quinn Valentine, Golnaz Navidi, Racheal Fisher, Rachel Galfo, Dario Nunez, and Adam Mather, thank you for being there every day for a good laugh, a tough conversation, or just a coffee walk. For the Richmond Lab members, thank you for providing me with the foundations for graduate school and inspiring me to be just like you! To the Brozek group, thank you for the fun, wild times, and for welcoming a physical chemistry person like me into your materials world. Finally, the biggest thank you to the four most important beings in my life. A huge thank you and all the love to my partner, Cassidi Howard. Thank you for being there at the exciting and stressful times, the times when I was grumpy and grumbling with Adobe Illustrator, and for all the times that I’ve practiced my talks for you. Thank you to my parents, Kim Mapile and Todd Mapile, without a doubt, I wouldn’t be finishing this crazy thing if it wasn’t for your unwavering support. And to Pepper, thank you for your kisses and cuddles! I love you all so much. 6 For Nibbler. 7 TABLE OF CONTENTS Chapter Page CHAPTER I: INTRODUCTION .................................................................................................. 15 I.A. SOLVATION AT NON-POROUS INTERFACES .......................................................... 16 I.A.1. Colloidal Stability and Solvation ................................................................................ 16 I.A.2. Capped Colloidal Nanoparticles .................................................................................. 18 I.A.3. Bare Nanoparticle Solvation ....................................................................................... 19 I.B. SOLVATION AT PSEUDO-POROUS INTERFACES .................................................... 20 I.B.1. Polymer Solubility ....................................................................................................... 20 I.B.2. Hydration of Proteins................................................................................................... 21 I.C. SOLVATION OF OIL-IN-WATER EMULSIONS .......................................................... 23 I.C.1. Emulsion Kinetic Stability........................................................................................... 23 I.C.2. Steric Stabilization of Emulsions ................................................................................ 25 I.D. SOLVATION AT POROUS INTERFACES .................................................................... 26 I.D.1. Simple Porous Structures ............................................................................................ 27 I.D.2. Porous Frameworks ..................................................................................................... 30 CHAPTER II: VIBRATIONAL SUM FREQUENCY SCATTERING SPECTROSCOPY (VSFSS) ........................................................................................................................................ 33 II.A.VSFSS THEORY .............................................................................................................. 33 II.B. LASER SETUP ................................................................................................................. 35 CHAPTER III: EMULSIONS STABILIZED BY POLY(ACRYLIC ACID) ............................. 38 III.A. AN INTRODUCTION TO POLY(ACRYLIC ACID) ................................................... 38 III.B. MATERIALS AND METHODS .................................................................................... 40 III.B.1. Materials ................................................................................................................... 40 8 III.B.2. Emulsion Formation.................................................................................................. 40 III.B.3. Dynamic Light Scattering and Zeta Potential ........................................................... 41 III.C. RESULTS AND DISCUSSION ...................................................................................... 41 III.C.1. Impact of PAA Concentration .................................................................................. 41 III.C.2. Emulsions Stabilized by PAA Alone ........................................................................ 42 III.C.3. Addition of Salt to PAA-stabilized Emulsions ......................................................... 48 III.D. CONCLUSIONS ............................................................................................................. 53 CHAPTER IV: IMPACT OF POLYMER STERICS ON EMULSION STABILITY ................ 54 IV.A. AN INTRODUCTION TO POLYMER STERIC HINDRANCE .................................. 54 IV.B. MATERIALS AND METHODS .................................................................................... 57 IV.B.1. Materials ................................................................................................................... 57 IV.B.2. Emulsion Formation ................................................................................................. 57 IV.B.3. Dynamic Light Scattering and Zeta Potential ........................................................... 57 IV.C. RESULTS AND DISCUSSION - EMULSIONS COATED WITH PAA OF VARYING MOLECULAR WEIGHT ......................................................................................................... 58 IV.C.1. Emulsion Characteristics .......................................................................................... 58 IV.C.2. Entropy and Enthalpy Favored Adsorption .............................................................. 59 IV.C.3. VSFSS of Varying Molecular Weight PAA Emulsions at pH 4 .............................. 60 IV.C.4. VSFSS of Varying Molecular Weight PAA Emulsions at pH 2 .............................. 64 IV.C.5. C=O Region and Summary ....................................................................................... 66 IV.D. RESULTS AND DISCUSSION - EMULSIONS COATED WITH PMAA .................. 67 IV.D.1. Emulsion Characteristics .......................................................................................... 67 IV.D.2. VSFSS of PMAA Coated Emulsions ....................................................................... 67 IV.E. RESULTS AND DISCUSSION - EMULSION THERMAL STABILITY .................... 70 IV.F. CONCLUSIONS ............................................................................................................. 71 9 CHAPTER V: COLLOIDAL STABILITY OF ZIF-8 ................................................................. 73 V.A. AN INTRODUCTION TO BARE MOF NANOPARTICLES ....................................... 73 V.B. MATERIALS AND METHODS ..................................................................................... 76 V.B.1. Materials..................................................................................................................... 76 V.B.2. Synthesis of nanoZIF-8 Particles ............................................................................... 76 V.B.3. nanoZIF-8 Sample Preparation for VSFSS ................................................................ 76 V.B.4. Powder X-Ray Diffraction ......................................................................................... 77 V.B.5. FT-IR Spectroscopy ................................................................................................... 77 V.B.6. Dynamic Light Scattering and Zeta Potential Measurements .................................... 77 V.C. RESULTS AND DISCUSSION ....................................................................................... 78 V.C.1. Ordered Solvation Shells Revealed by VSFSS .......................................................... 78 V.C.2. Spontaneous Ordering within the ZIF-8 Nanoparticles ............................................. 82 V.D. CONCLUSIONS .............................................................................................................. 85 CHAPTER VI: CHARACTERIZING THE MOF-POLYMER INTERFACE FOR IMPROVED MEMBRANE FORMULATION ................................................................................................. 86 VI.A. AN INTRODUCTION TO MIXED MATRIX MEMBRANES .................................... 86 VI.B. MATERIALS AND METHODS .................................................................................... 89 VI.B.1. Materials ................................................................................................................... 89 VI.B.2. Synthesis of nanoZIF-8 ............................................................................................. 89 VI.B.3. Formation of MMMs ................................................................................................ 89 VI.B.4. Dynamic Light Scattering and Zeta Potential ........................................................... 90 VI.B.5. FT-IR Spectroscopy .................................................................................................. 90 VI.B.6. Scanning Electron Microscopy ................................................................................. 90 VI.C. RESULTS AND DISCUSSION ..................................................................................... 90 VI.C.1. Water Stability of ZIF-8 with Binding Agent ........................................................... 90 10 VI.C.2. Characterization of Colloids ..................................................................................... 92 VI.C.3. Ordered Polymer Layers Measured by VSFSS ........................................................ 94 Poly(acrylic acid) .................................................................................................................. 94 Poly(acrylamide) ................................................................................................................... 95 Poly(N-vinyl acetamide) ........................................................................................................ 97 VI.C.4. Characterization of MMMs ...................................................................................... 98 VI.C.5. Tensile Strength Testing ........................................................................................... 99 VI.C.6. Force-Probe AFM ..................................................................................................... 99 VI.D. CONCLUSIONS ........................................................................................................... 100 CHAPTER VII: ORDERED SOLVATION SHELLS AND THE STABILITY OF POROUS AND NON-POROUS COLLOIDS ............................................................................................ 101 CHAPTER VIII: THE FUTURE OF VSFSS FOR PROBING UNIQUE MECHANISMS OF COLLOIDAL STABILITY ........................................................................................................ 104 APPENDIX A: SUPPLEMENTARY INFORMATION FOR CHAPTER III .......................... 108 APPENDIX B: SUPPLEMENTARY INFORMATION FOR CHAPTER IV........................... 123 APPENDIX C: SUPPLEMENTARY INFORMATION FOR CHAPTER V ............................ 131 APPENDIX D: SUPPLEMENTARY INFORMATION FOR CHAPTER VI .......................... 139 APPENDIX E: SUPPLEMENTARY INFORMATION FOR CHAPTER VII ......................... 143 APPENDIX F: CALCULATION OF NANOPARTICLE SURFACE DENSITY .................... 144 APPENDIX G: VSFSS NORMALIZATION PROCEDURE .................................................... 147 APPENDIX H: DETERMINATION OF PARTICLE CONCENTRATION FOR VSFSS DETECTION LIMIT .................................................................................................................. 151 REFERENCES ........................................................................................................................... 153 11 LIST OF FIGURES Figure Page Figure I. 1 A) Typical DLVO graph ............................................................................................ 17 Figure I. 2. Illustration of emulsion formation ............................................................................ 24 Figure I. 3 Illustration of A) 1-dimensional (polymers, nanotubes). ........................................... 28 Figure I. 4 Representation of A) framework materials. ............................................................... 31 Figure II. 1 A) Schematic of VSFSS............................................................................................33 Figure II. 2 Schematic of the VSFSS experiment ....................................................................... 36 Figure III. 1 Chemical structure of poly(acrylic acid)..................................................................38 Figure III. 2 Characteristics A) hydrodynamic diameter ............................................................ 41 Figure III. 3 VSFSS measurements. ............................................................................................ 44 Figure III. 4 Characteristics (top panel: hydrodynamic diameter ............................................... 48 Figure III. 5 VSFSS measurements (ssp polarization) ................................................................ 50 Figure III. 6 Illustration depicting the behavior of PAA ............................................................. 52 Figure IV. 1 A) Hydrodynamic diameter......................................................................................58 Figure IV. 2 Critical adsorption parameter (δc)........................................................................... 60 Figure IV. 3 VSFSS measurements (ssp polarization) ................................................................ 61 Figure IV. 4 Molecular picture of the hexadecane/water droplet interface ................................. 63 Figure IV. 5 VSFSS measurements (ssp polarization). ............................................................... 64 Figure IV. 6 Molecular picture of the hexadecane/water droplet interface ................................. 65 Figure IV. 7 VSFSS measurements (ssp polarization) ................................................................ 68 Figure IV. 8 Temperature-dependent diameter. ........................................................................... 70 Figure V. 1 A) VSFSS measurements of nanoZIF-8 particle.......................................................79 Figure V. 2 A) VSFSS measurements (ssp polarization)............................................................. 81 Figure V. 3 A) VSFSS measurements (ssp polarization)............................................................. 83 Figure VI. 1 Representation of the ZIF-8 nanocrystal surface.....................................................87 Figure VI. 2 PXRD patterns. ....................................................................................................... 91 Figure VI. 3 A) Hydrodynamic diameter ..................................................................................... 92 Figure VI. 4 VSFSS measurements (ssp polarization) ................................................................ 96 Figure VI. 5 A) Image of MMMs ................................................................................................ 98 12 Figure VII. 1 VSFSS measurements (ssp polarization)..............................................................102 Figure VII. 2 Dynamic light scattering distribution curves ....................................................... 103 Figure VIII. 1 Temperature dependent hydrodynamic diameter................................................105 Figure VIII. 2 a) Zeta potential of Cu(TA)2 nanoparticles ....................................................... 106 Figure A. 1 Zeta potential of 500 ppm PAA-stabilized..............................................................108 Figure A. 2 Visual representation of the time-dependent colloidal stability ............................. 109 Figure A. 3 Equilibrium surface pressure. ................................................................................. 110 Figure A. 4 Visual representation of the time-dependent colloidal stability. ............................ 111 Figure A. 5 Hydrodynamic diameter distribution curves of colloids ........................................ 111 Figure A. 6 Dynamic light scattering and Zeta Potential measurements ................................... 112 Figure A. 7 VSFSS measurements (ssp polarization) ................................................................ 116 Figure A. 8 VSFSS measurements (ssp polarization) ................................................................ 116 Figure B. 1 VSFSS measurements (ssp polarization).................................................................123 Figure B. 2 Fit parameter trends for VSFSS measurements ...................................................... 124 Figure B. 3 Equilibrium surface pressure. ................................................................................. 125 Figure B. 4 Polydispersity index (PDI) of nanoemulsions ........................................................ 126 Figure C. 1 SEM images of ZIF-8 nanoparticles........................................................................131 Figure C. 2 PXRD patterns of powder nanoZIF-8..................................................................... 131 Figure C. 3 FT-IR spectra. ......................................................................................................... 131 Figure C. 4 Impact of DMF absorption on VSFSS spectra. ...................................................... 132 Figure C. 5 Size distribution curves of nanoZIF-8 particles...................................................... 133 Figure C. 6 Acid digestion 1H NMR ......................................................................................... 133 Figure C. 7 TGA trace of nanoZIF-8. ........................................................................................ 134 Figure C. 8 Surface zeta potential of nanoZIF-8 ....................................................................... 134 Figure C. 9 VSFSS measurements (ssp polarization) ................................................................ 135 Figure D. 1 pH measurements of ZIF-8......................................................................................139 Figure D. 2 Pictures of ZIF-8 colloids ....................................................................................... 139 Figure D. 3 A) Amplitude, B) Gaussian linewidth, and C) peak position trends ...................... 140 Figure D. 4 FT-IR spectra of polymeric membrane films ......................................................... 140 Figure E. 1 Dynamic light scattering curves...............................................................................143 Figure G. 1 Non-resonant response from KNbO3 crystal...........................................................147 13 Figure G. 2 Raw signal obtained from one scan. ....................................................................... 147 Figure G. 3 Average signal trace ............................................................................................... 148 Figure G. 4 Averaged signal trace. ............................................................................................ 148 Figure G. 5 Averaged signal trace. ............................................................................................ 149 Figure G. 6 Signal trace from colloidal ZIF-8 in DMF normalized .......................................... 149 Figure G. 7 Completely normalized signal trace. ...................................................................... 150 Figure H. 1 Signal-to-noise ratio................................................................................................152 14 LIST OF TABLES Table Page Table IV. 1 Integrated area of VSFSS spectra ............................................................................. 69 Table A. 1 Emulsion zeta potential.............................................................................................109 Table A. 2 Polydispersity index (PDI) values ............................................................................ 113 Table A. 3 Fitting parameters for CH spectra ............................................................................ 117 Table A. 4 Fitting parameters for C=O spectra .......................................................................... 118 Table A. 5 Fitting parameters for spectra ................................................................................... 119 Table A. 6 Fitting parameters for spectra ................................................................................... 120 Table A. 7 Fitting parameters for spectra ................................................................................... 121 Table A. 8 Fitting parameters for spectra. .................................................................................. 122 Table B. 1 Characteristics (diameter, polydispersity index, and zeta potential).........................125 Table B. 2 Fitting parameters for C-H spectra ........................................................................... 127 Table B. 3 Fitting parameters for C=O spectra. ......................................................................... 128 Table B. 4. Fitting parameters for C-H spectra .......................................................................... 130 Table B. 5 Fitting parameters for C=O spectra .......................................................................... 130 Table C. 1 Fitting parameters for spectra of ZIF-8.....................................................................136 Table C. 2 Fitting parameters for spectra of ZIF-8 .................................................................... 137 Table C. 3 Fitting parameters for spectra of ZIF-8. ................................................................... 138 Table D. 1 Fitting parameters for spectra of ZIF-8.....................................................................141 Table D. 2 Fitting parameters for spectra of ZIF-8 .................................................................... 141 Table D. 3 Fitting parameters for spectra of ZIF-8 nanoparticles. ............................................. 142 15 CHAPTER I: INTRODUCTION This introduction provides a background on the traditional theories used to describe solvation and stability of common colloids (nanoparticles, polymers, and proteins) whose mechanistic details are important for application to our nano-sized emulsions and MOFs. Next, this introduction specifies the unique characteristics of the systems studied here – namely the kinetic stability of emulsions and the porosity of nanoMOFs – and presents avenues for measuring colloidal behavior. The introduction contains material published in Chemistry of Materials as a Perspective written with co-authors Erik Svensson Grape and Carl K. Brozek in 2024. Chapter II provides a technical background on vibrational sum frequency scattering spectroscopy (VSFSS) and details the experimental setup used in this work. Chapters III and IV describe two research projects that study the role of steric bulk from a polymer layers as a mechanism for stability of hexadecane emulsions. The role of pH- and salt-sensitivity is introduced in Chapter III and molecular weight dependence is probed in Chapter IV. Chapter III was published in Colloids and Surfaces A: Physiochemical and Engineering Aspects in 2024 and Chapter IV is currently submitted to Soft Matter, with both manuscripts supervised and written with Lawrence F. Scatena. Chapter V was published in ACS Nano in 2024 in collaboration with Michael A. Leroy, Kevin Fabrizio, Lawrence F. Scatena, and Carl K. Brozek and introduces VSFSS as an emerging technique to measure the surface of bare, colloidally stable MOF nanoparticles. Chapter VI contains unpublished work in preparation for submission co-authored with Michael A. Leroy, Audrey M. Davenport, Lawrence F. Scatena, and Carl K. Brozek that combines the two main aspects of this dissertation, nanoMOF particles and polymer coatings, where the surface structure of carbonyl polymers at very low concentrations is measured and proposed as a solution to homogenously disperse nanoMOF particles in a polymeric membrane. Finally, Chapter VII details unpublished work in collaboration with Adam P. Mather and Carl K. Brozek that connect the fundamentals of colloidal stability discovered in this work, namely solvation shells, to the relevant parameter of critical aggregation concentration of both porous and non-porous inorganic colloids. Chapter VIII will conclude this dissertation with remarks on the exciting and broad future of VSFSS for studying buried, colloidal surfaces. 16 I.A. SOLVATION AT NON-POROUS INTERFACES I.A.1. Colloidal Stability and Solvation In the 1940s, two teams of researchers independently studied the forces contributing to particle colloidal stability. Boris Derjaguin and Lev Landau in the Soviet Union presented their theory of colloidal stability that invoked short-range van der Waals attractions between particles overcome by electrostatic surface repulsions.1 While the electrical double layer had previously been introduced by Hermann von Helmholtz (1853),2 Louis Georges Gouy (1910),3 David Leonard Chapman (1913),4 and Otto Stern (1924),5 Derjaguin and Landau pioneered the notion of electrostatic potentials at curved interfaces where the surface electric field decays as a function of the particle radius. Meanwhile, in the Netherlands, Evert Verwey and Theodoor (Jan) Overbeek developed the now-ubiquitous potential energy curves of two interacting spherical particles as a function of interparticle distance and electrolyte concentration.6,7 Ultimately, a model of the forces governing nanoparticle stability in solution was named DLVO theory for the four authors involved.8 DLVO theory has been thoroughly derived to describe the colloidal stability of hard- shell nanoparticles in a variety of electrolyte concentrations,9,10 surface charge composition,11,12 solvent environments,13 and many other scenarios.13–18 DLVO theory can be summarized by Eq. 1: 𝑊(𝐷) = 𝑊𝑣𝑑𝑤 + 𝑊𝑒𝑙𝑒𝑐 (1) where W(D) represents the interaction energy between neighboring particles, Wvdw is the attractive energy due to van der Waals interactions and Welec is the repulsive electrostatic energy (Figure I.1A). The entropy and steric pressures of the system have been incorporated into extended DLVO theory (XDLVO)19,20 as described by Eq. 2: 𝑊(𝐷) = 𝑊𝑣𝑑𝑤 + 𝑊𝑒𝑙𝑒𝑐 + 𝑊𝑜𝑠𝑚𝑜𝑡𝑖𝑐 + 𝑊𝑒𝑛𝑡𝑟𝑜𝑝𝑖𝑐 (2) where the addition of the Wosmotic term accounts for the repulsive energy caused by neighboring nanoparticles with an increased overlap region (typically due to increased surface sterics), while the Wentropic term accounts for the attractive energy arising from an increase in solvent entropy upon particle coalescence. This entropy-driven aggregation is often termed the hydrophobic effect. Although developed as a general model, DLVO theory applies best to “hard” colloidal systems, such as colloidal quantum dots, core-shell metallic nanoparticles, or metal oxides with 17 well-defined boundaries between the nanoscale object and the suspending solution. Whereas hard- shelled colloids maintain their rigidity in solution, “soft-shelled” particles stretch and adapt to nearby fluid interfaces (Figure I.1B).21–24 Hard-shell particles are convenient systems to study as short-range van der Waals interactions decay exponentially as a function of distance from the particle without the need to consider interfacial charge-screening. On the other hand, soft-shell particles, such as oil droplets, metal nanoparticles coated with a polymer layer, proteins clusters, and even living cells,25 present complications to DLVO theory as solvent intercalation, ion adsorption,26 spontaneously ordered solvation shells , and surface roughness,15 together comprising so-called “non-DLVO forces”. Experimental attempts to bridge these two types of systems, such as core-shell colloids with soft exteriors and hard interiors,21,27 impart improved thermodynamic stability and, in some cases, unique optical behavior.28,29 Theoretical models of Figure I. 1 A) Typical DLVO graph showing the interaction energy as a sum of electrostatic repulsion and attractive van der Waals interactions. B) Illustration of the soft-to-hard range of nanoscale materials ranging from polymers/proteins to hard spheres, such as metal oxide nanoparticles. C) The lattice model of polymer solvation as described by Flory-Huggins theory. D) Illustration and ion examples in relation to the Hofmeister series of solvation. Figure I.1 A) Typical DLVO graph showing the interaction energy as a sum of electrostatic repulsion and attractive van der Waals interactions. B) Illustration of the soft-to-hard range of nanoscale materials ranging from polymers/proteins to hard spheres, such as metal oxide nanoparticles. C) The lattice model of polymer solvation as described by Flory-Huggins theory. D) Illustration and ion examples in relation to the Hofmeister series of solvation. 18 such systems account only for the hard-soft interface, while neglecting the possibility of low density and heterogeneous surfaces and their associated energetics. According to the IUPAC definition, a colloid involves a “molecule or polymolecular particle dispersed in a medium” with “at least in one direction a dimension roughly between 1 nm and 1 µm”.30 Examples of colloids include a solid in a gas (smoke), a liquid in a gas (aerosol), or a liquid in a liquid (emulsion), in addition to many other dispersion types that do not require solvent as a medium. Understanding the stability of heterogeneous mixtures in solution is critical to their implementation in displays, coatings, or membranes, but the clear phase boundary creates significant challenges because of the difficulty in studying the chemistry that emerges at the colloidal interface. The process of dissolution, by contrast, involves the formation of a single, homogeneous phase.31 Unlike solutions, the heterogeneity of colloidal dispersions creates thermodynamic instability resulting in their eventual phase separation. Suspended particles can be filtered out or mechanically separated, whereas separating dissolved particles requires additional chemical transformations.32,33 In the following sections, we provide a brief overview of the chemistry underlying the colloidal stability of non-porous inorganic nanoparticles with and without functionalized surfaces, and how it critically involves solvent. I.A.2. Capped Colloidal Nanoparticles Nearly all non-porous inorganic nanoparticles require surface functionalization for long-term stability. The strong van der Waals attractions between bare metallic surfaces results in nearly instantaneous aggregation and sedimentation. Semiconductor nanocrystals, specifically quantum dots, the topic of the Nobel Prize in Chemistry in 2023,34,35 represent another key example of nanoparticles that find utility only by surface functionalization. Attaching surfactant ligands to prevent Ostwald ripening36,37 allows semiconductor nanoparticles to remain stable at precise sizes that dictate their tunable optical properties.38–40 Functionalizing quantum dots with either an inorganic shell also diversifies the available solvents for dispersion.41,42 Most notably, the addition of water-soluble ligands, such as citrates,43 polymers,44,45 or other small molecules46,47 allows for the use of metallic nanoparticles within cellular media. A wide variety of ligands have been studied for their interaction with various solvent media.48,49 By contrast, nearly all porous colloidal materials, as described below, require no surface functionalization for colloidal stability. 19 Therefore, for context, we must understand the mechanism of colloidal stability for nanoparticles prepared without capping agents. I.A.3. Bare Nanoparticle Solvation Few reports have documented bare nanoparticles exhibiting long-term colloidal stability. The examples described below are restricted in their industrial applications due to highly defective surfaces, the necessity of a stabilizing supporting material, and the need for post-synthetic cleaning or filtration. For example, while it has been shown that metallic nanoparticles can be stabilized by a partially oxidized surface,50 this chemical decomposition generates a heterogeneous and rough surface with mixed valence sites.51,52 An alternate approach that avoids surface oxidation was achieved by synthesizing bare copper nanoparticles with an electron-donating gadolinium support.53 Implementing nanoparticles in biological sensing applications cannot be achieved, however, with a bulky and potentially toxic supporting material. Another approach involves “sterilization” of bare gold nanoparticles by autoclave.54 Studies suggest that < 5-nm particles assemble into larger 10-30 nm particles during this process. This size-focusing serves as a form of filtration to achieve a narrow size dispersity, but size selection does not always ensure colloidal stability.55 Because these few examples of bare nanoparticles are short-lived colloids, they are often quickly drop-casted back to their solid state.56–59 In the few studies of bare metallic nanoparticles in solution without a stabilizing support, the ionic structure of the supporting electrolyte and the resulting electrical double layer at the surface has proven to be a key, but poorly understood feature contributing to colloidal stability. Simulations suggest that facet-specific ion adsorption and an ordered primary solvation shell promote gold nanoparticle stability in water.52,60 Experimental studies also suggest water stability of bare nanoparticles arises from ion adsorption dependent on electrolyte concentration and following the so-called Hofmeister series for anions.54,61–63 The colloidal stability of nanoparticles in solvents without electrolyte, however, have only been studied as computational simulations of bare particles64 or through weakly ligated experimental systems where a portion of the surface remains deliberately exposed.65 Thus, achieving a metal nanoparticle colloid stabilized by solvation forces alone remains an open challenge and would open an avenue for tuning the opto- electronic and electrochemical electrical properties of metallic nanoparticles in a wide variety of 20 solutions. To understand the colloidal stability of bare nanomaterials, therefore, we instead turn to the supramolecular solvation chemistry of polymers and proteins. I.B. SOLVATION AT PSEUDO-POROUS INTERFACES I.B.1. Polymer Solubility At nearly the same time that Derjaguin, Landau, Verwey, and Overbeek developed their model of colloidal stability, Paul Flory and Maurice Huggins independently considered how polymers could dissolve despite distinct differences in molecular size from the surrounding solvent.66,67 While the entropy of mixing a simple molecule in solution can be described by Gibbs energy of mixing: ∆𝐺𝑚𝑖𝑥 = ∆𝐻𝑚𝑖𝑥 − 𝑇∆𝑆𝑚𝑖𝑥 (3) where H is enthalpy, T is temperature, and S is entropy, this relationship only considers the gross interaction between a molecule and its surroundings. It neglects synergistic behavior between individual units within a chain, such as in a polymer (Figure I.1C). Flory and Huggins adapted the Gibbs energy of mixing for polymers as: ∆𝐺𝑚𝑖𝑥 = 𝑅𝑇(𝑛1𝑙𝑛𝜙1 + 𝑛2𝑙𝑛𝜙2 + 𝑛1𝜙2𝜒12) (4) which now considers the number of moles (𝑛) and the volume fraction (𝜙) of the solvent (component 1) and polymer (component 2). 𝜒12 is material-specific and describes the synergistic interaction between the polymer and the solvent, allowing for specific descriptions of polymer solubility. We introduce the term “pseudo-porosity” to describe the solvation of polymers and related systems. While polymers lack permanent porosity (the ability to maintain a rigid, porous structure when suspended in solution), their solvation mechanism depends on the large solvent accessible surface areas carved out by the wrapping of individual monomer units. In Flory-Huggins theory, a polymer solution is modeled as a lattice of cells containing either polymer monomer units or a solvent molecule. Based on this image, the individual mixing components of a monomer unit with the polymer is well described by equation 4. However, realistic intermolecular forces from hydrophobicity,68 ionic screening,69 and other polymer or solvent-specific interactions70,71 integral to solvation and polymer conformation are absent from eq. 4. 21 In addition to the pseudo-porosity of conventional polymers, intrinsically porous polymers and porous polymer membranes also exist. Typical examples of porous polymer membranes include polycarbonate, polyester, or cellulose with pore diameters ranging from nanometers to 10s of µms. The narrow, one-dimensional confinement of these pores within a membrane have led to a versatile platform for studying solvent-controlled ion-transport under spatial confinement akin to the ion-transport across cellular membranes and carbon nanotubes.72,73 Chemical functionalization of these materials leads to specific ion effects, such as appended carbonate groups showing preferential binding of metal cations.74 Polymers of intrinsic microporosity (PIMs), on the other hand, contain voids of 2 nm or less.75 Controlling the solubility of PIMs is critical for creating robust films with consistent pore sizes.76 The commonly used PIM-1, however, dissolves only in tetrahydrofuran (THF) and chloroform (CHCl3), leading to reduced solution processability and restricting the incorporation of fillers or other composites. Recent work has introduced post- synthetic modifications of PIMs to improve their solubility,77 with the key result that specific solvent interactions with appended moieties improves solubility. In understanding the solvation of porous materials discussed in Section 3, these reports of porous polymers offer the insight that solution stability of nanoscale materials benefits from monomer units accessible to solvent. I.B.2. Hydration of Proteins Proteins comprise another class of macromolecules whose solvation depends on solvent interacting with accessible building blocks, yet with the added compositional diversity of amino acid mixtures and H-bonding networks. For proteins, solvation can be described as monomer-by-monomer solubility or by secondary structure solvation via protein folding to bury hydrophobic moieties. As a result, proteins arrange into secondary structures, like helices or sheets, and fold into native states to satisfy solvent interactions with their respective amino acid chains. Numerous studies have quantified surface interactions between proteins and water in solution,78,79 with one notable computational investigation into the hydrogen bonding of a protein in water as a function of solvent accessible surface area and protein conformation.80 These authors report that the solvation free energy of a protein decreases linearly as a function of the solvent accessible surface area of the protein. This finding suggests that an increase in the surfaces available for solvation promote stability in solution. As the basis of Flory-Huggins theory, leveraging enthalpically favorable 22 solvent-material interactions is critical for solubility and stability. In a permanently porous material, as explained below, the unparalleled solvent accessible surface areas provide increased sites for solvation interactions via internal and external surfaces. The unique folding ability of a protein allows it to adapt to an otherwise thermodynamically unstable environment. In contrast, porous framework materials, such as metal-organic frameworks (MOFs), cannot protect select components by geometrical reconfiguration. Protein folding or unfolding can be induced by addition of ions that “salt-in” or “salt-out” the macromolecule of interest (protein, polymer, nanoparticle, etc.) vis-à-vis the Hofmeister series (Figure I.1D). Ion specific effects in protein precipitation was introduced by Franz Hofmeister in the late 1880s to describe how different salts impact the solubility of proteins despite possessing the same net charges.81,82 Kosmotropes, or “structure-makers” are ions that interact more strongly with water than with the protein themselves, causing the protein to remain in its native folded state and salt- out of solution. Common examples of kosmotropes include citrate, sulfate, and phosphate anions or magnesium, calcium, or lithium cations. On the other hand, chaotropes, or “structure-makers” interact closely with individual protein units, promoting an unfolded, solubilized state of the protein. Chaotropes include iodide, nitrate, and tetrafluoroborate anions or calcium, magnesium, and aluminum cations. Polyoxometalates (POMs), for example, are super-chaotropes due to their large, delocalized charge and low surface densities.83 While Hofmeister’s initial experiments provide a basis for harnessing specific-ion effects, recent work has produced a more detailed understanding of salts and their impact on colloidal stability and solubility, especially cooperative contributions from ion pairs.84–87 Vibrational spectroscopy and molecular dynamic simulations reveal that cations follow the Hofmeister series through strong backbone-salt interactions and weaker interactions with negatively charged side chains.88 Anions, however, despite following the Hofmeister series at the backbone, exhibit a reversed trend on positively charged amino acid residues. For this reason, the Hofmeister series for anions holds only when the backbone-salt interactions outweigh the salt interactions with side chains. Tuning the colloidal stability of metal- based nanoparticles requires understanding the specific-ion effect in solvents beyond water.89,90 A comprehensive study of the ion effects in non-aqueous solvents found that water is not unique in its role in the Hofmeister series.91 In fact, aprotic solvents also show an ion specificity due to the inherent molar volume and electrostriction of the ion, or the ability of the ion to slightly deform. 23 These results confirm that a Hofmeister trend persists regardless of solvent identity, thereby providing a basis for tailoring the colloidal stability of non-aqueous materials. Specific ion effects are critical to the colloidal stability of bare nanoparticles. In one study, gold nanoparticles without any organic ligands at the surface were stabilized by as little as 10 µM of a chaotropic anion.61 The electrolyte solutions were added during nanoparticle synthesis, thereby acting as a non-organic capping ligand. Additional electrolyte may destabilize colloids, however. Reports indicate that increasing salt concentration to the mM regime often leads to nanoparticle aggregation, as exhibited by metallic nanosheets titrated with potassium salts of varying valency anions92 and other metal nanoparticles.93,94 A likely causes is that with increased salt concentrations electrolyte ions screen the effective charge density that normally prevents aggregation. Nevertheless, tuning salt concentrations could serve as a promising strategy for controlling the solvation of porous nanoparticles with bare surfaces. I.C. SOLVATION OF OIL-IN-WATER EMULSIONS I.C.1. Emulsion Kinetic Stability Formulating stable emulsions and understanding the mechanism of oil droplet stability has far reaching implications, from the development and enhancement of new cosmetic products, to the efficacy of drug delivery systems and oil spill remediation technologies.95–97 Traditionally, oil droplets with diameters of 20-200 nm are classified as nanoemulsions while larger droplets with diameters of 200-1000 nm in diameter can be referred to as miniemulsions.98 Regardless of the size, surfactant-coated emulsions are useful laboratory analogues for modeling commercially-used dispersants that reduce the surface tension of a large oil patch and enable the formation of stable, small oil droplets that are continuously distributed by waves and eventually biodegraded by marine microbes.99 However, these oil droplets require high energy mixing for formation and are thermodynamically unstable leading to eventual phase separation given enough time.100–103 In practice, ultrasonication is necessary to input the energy required for the system to enter the kinetically trapped state (Figure I.2). The addition of surfactants, electrolyte, polymers, and/or proteins can provide a protective layer that prevents particle coalescence over time. DLVO theory explains that particle-particle repulsion in emulsions is commonly accounted for by balancing 24 attractive van der Waals interactions with repulsive electrostatic forces, such as those typically achieved by an ionic surfactant. A zeta potential, or surface charge, of about ± 30 mV is typically necessary to a stabilize a colloid. Extended DLVO theory further accounts for steric effects with the addition of a destabilizing osmotic pressure term and a stabilizing entropic term. In most applications, emulsion stabilization is ensured by surface functionalization that provides electrostatic or steric repulsion, or a combination of the two. Beyond oil remediation, the curved interface of emulsions have proven to be of interest to measure the molecular packing behavior as compared to planar oil/water interfaces. The Richmond/Scatena lab and others have postulated that the geometry of the interface and the thermodynamic instability of emulsions are responsible for the observed differences in molecular orientation and conformation at the two interfaces.104–106 In one regard, ultrasonication forces both the formation of oil into small droplets and the adsorption of surfactants or polymers to the interface. The addition of ultrasonicating energy can cause unexpected interfacial behavior as compared to what has been exhibited at a planar oil/water interface where small molecules Figure I. 2. Illustration of emulsion formation from a phase separated state to a dispersion via ultrasonication. The dispersed emulsion is kinetically trapped and will eventually return to the phase separate state. Figure I.2 25 thermodynamically adsorb to the interface at an equilibrium time scale. Not only do emulsions differ from their planar counterparts in terms of kinetic stability, but they exhibit a curved geometry, instead of flat. While the droplet geometry increases overall interfacial surface area for chemical adsorption, it also enhances steric and electrostatic crowding effects which can reduce interfacial coverage of surfactants, polymers, or ions.107 Keeping these aspects of thermodynamic and kinetic stability in mind for the planar and droplet oil/water interfaces, respectively, are important when interpreting VSFS spectra from each chemical environment. I.C.2. Steric Stabilization of Emulsions While non-toxic surfactants are useful for environmental applications, the introduction of proteins as colloid stabilizers have the potential to encapsulate and selectively deliver hydrophobic drugs.108–112 It is hypothesized that colloids are stabilized by protein coronas through steric hinderance or, in other words, a thick protein shell prevents droplets from coalescing. While recent work using 2D-IR spectroscopy and sum frequency scattering spectroscopy have investigated the newfound importance of protein-stabilized colloids, the molecular conformation and detailed mechanism of stability remain difficult to probe due to the complicated nature of long-chain proteins and their resulting coronas.113,114 Polyelectrolytes are useful analogues for studying protein coronas as both macromolecules have varying functional groups along a repeating backbone, a nuanced folding or coiling structure with the ability to hydrogen bond or form secondary structures, and charge localization along the chain dictated by environmental pH. Yet, polyelectrolytes are often easier to study than proteins due to their known monomer units and simple tunability of molecular weight, percent protonation, and secondary structure. In addition to their pH-dependent behavior in bulk solution, acids and polyelectrolytes adsorbed at interfaces are known to have a surface pKa ~1 unit more alkaline than the bulk pKa due to interfacial solvent dielectrics reducing the solvation energy of the acid functional groups.115–120 This pH-dependence is critical for materials such as polyelectrolyte coarcervates, where stabilization across a variety of conditions can be a challenge.121,122 The adsorption behavior of polymers has generated both theoretical123 and experimental124,125 studies that seek to understand interfacial conformation. In particular, a “train, loop, tail” representation has been developed that describes how polymers lie flat at the surface (train), coil beyond the surface into the bulk (loop) and extend at the ends of the polymer chains (tail) to emphasize 26 favorable surface-polymer interactions or bulk-polymer interactions, dependent on the polymer identity and interface.126–128 Here, VSFSS is a powerful tool to probe polymer conformation at interfaces. Previous work from Marc Foster in the Richmond Laboratory investigated the role of carboxylic acid surfactants at the curved oil/water interface.106 These experiments were fundamental in developing a rigorous experimental procedure to measure the carbonyl and carboxylate modes via VSFSS. A non-resonant contribution from the CaF2 windows used in VSFSS contribute to a large background that requires de-timing of the visible pulse to deconvolute the resonant vibrational modes of interest.129 Additionally, Marc found that the carbonyl and carboxylate modes at the oil/water droplet interface differed in both spectral line shape and peak position then the same modes measured at a planar oil/water interface or the planar air/water interface.104,130 These differences were attributed to the difference in solubility of the hydrophobic phases with the alkyl chains of the surfactant. The improved solubility of the surfactant in the CH- rich oil phase as compared to air, for example, would increase the ordering of the head group at the droplet interface. At the same time, it was proposed that the curved geometry of the droplet resulted in charge repulsion across the droplet that caused a lower population of surfactants at the droplet surface. With this increased surfactant head group area, the carbonyl and carboxylate modes exhibited improved ordering at the oil/water droplet interface. In this dissertation, the ability of a low-charge polymer to stabilize emulsions without the assistance of a surfactant is explored. The colloidal stability of this polymer system was further probed by adjusting the pH, salt identity and concentration, molecular weight of the polymer, backbone substitution, and temperature. The adjustment of these variables provide insight into the physiological robustness of polymer-stabilized emulsions with tunable parameters that can improve drug delivery platforms. I.D. SOLVATION AT POROUS INTERFACES To understand the solubility and stability of porous colloids, we suggest using the aforementioned models of solvation, namely electrostatic or steric repulsion, the hydrophobic effect, monomer-by- monomer solubility, and specific ions interactions. By possessing both internal and external surfaces, porous colloids exhibit surface area-to-volume ratios and accessible void spaces far 27 exceeding any class of non-porous or pseudo-porous material.131 For example, the surface density of the commonly studied zeolitic imidazolate framework ZIF-8 (Zn(2-methylimidazolate)2), has a surface density of 7 atoms/nm2 while non-porous ZnO has a surface density of 78 atoms/nm2 (Appendix F). Additionally, as illustrated by the example above, the internal and external surfaces of porous materials resemble the heterogeneous surfaces of proteins. We expect the surface of porous materials to be susceptible to specific solvent-surface interactions as opposed to generalized hard-shell interactions between a homogeneous nanoparticle surface or surfactant ligands and the supporting solvent. In the following sections, we describe common permanently porous materials (i.e., materials that retain porosity in solution) and propose mechanisms for their colloidal stability and solubility. This research field is nascent, without consensus around theory that explains the colloidal stability of porous materials. We propose, given the examples below, that specific solvent-surface interactions, such as through ordered solvation shells and chemical interactions between the material building blocks and the solvent provide the colloidal stability of porous materials. I.D.1. Simple Porous Structures As a starting point, nanotubes serve as a basis for understanding the solvation of a one-dimensional porous material (Figure I.3A). The pores, or tunnels, of nanotubes can range from 0.5-2 nm in diameter and are typically grown from a metal catalyst via chemical vapor deposition. One might assume that water should be excluded from the nanotube interior due to its hydrophobicity and because water absorption is entropically disfavored.132 Surprisingly, water adsorbs inside the pores of carbon nanotubes, leading to the phenomenon of “water wires”.133 One explanation proposes that water wires and carbon nanotube solubility results from the entropy of water flowing through the pores and from free rotation of water molecules. Studies also suggest dissolution in water does not involve favorable enthalpic interactions between water and the nanotube, suggesting water solubility of other hydrophobic materials through entropy-favored interactions. Changing the nanotube polarity, via simulations or experimental crystal engineering, allows for tuning of the water wire mobility, specifically in terms of water migration from one opening of the tube to the other, functioning as a nanosized garden hose. Carbon nanotubes also dissolve in ionic liquids,134,135 where both solvation shells and internal hydrogen bonding are postulated to stabilize the particles. Similar solvation behavior has been observed for carbon nanotubes in benzene,136 28 alcohols,137 and polymer solutions,138 however nearly all studies are computational. Although challenging, experimental studies of nanomaterials solvation will lay the foundation for designing their application in solution state applications, such as drug delivery or membranes. Whereas nearly all nanotubes are homogenous and comprised of carbon, the solvation of 2D materials introduces the probability of surface heterogeneity (Figure I.3B). In addition to the atomically homogeneous example of graphene, 2D materials include the heterogeneous boron nitride, metal chalcogenides, or metal oxides. In these materials, solvent accessible surface areas exist between sheets (interlayer), introducing another unique solvation environment in addition to interior and exterior pores. Despite their enhanced surface areas, most nanosheets require a surfactant or post-synthetic modification for long-term colloidal stability, similar to 3D, non- porous inorganic analogs.139–141 As illustrated by the examples below, electrical double layers are thought to spontaneously assemble at the surface of nanosheets, providing a large electrostatic surface repulsion (as measured by zeta potential) and site-specific water interactions at the surface. These results suggest that in high-surface area materials without deliberately added surface Figure I. 3 Illustration of A) 1-dimensional (polymers, nanotubes), B) 2-dimensional (layered or sheet-like materials), and C) 3-dimensional structures (porous materials) and their possible solvent interactions. Figure I.3 29 capping agents, specific solvent interactions are important for colloidal stability. Surprisingly, hydrophobic 2D materials, including graphene and MoS2, disperse in aqueous solutions via exfoliation-induced oxidation.142,143 Studies suggest that ultrasonication causes etching of edge- groups that improves the solubility in water. In a notable study, graphene ultrasonicated at high- temperatures exhibited long-term colloidal stability in water while the sample sonicated at low- temperatures maintained a pristine morphology and was unstable in water.144 The introduction of edge-site functionalization to graphene oxide in the form of hydroxyl and carboxyl groups also contributes to the increased water stability. However, for heterogeneous nanosheets of hexagonal boron nitride, MoS2, WS2, and MoSe2, colloidal stability in water was achieved after sonication at both high- and low-temperatures and no surface functionalization was observed. 3D porous materials with relatively simple compositions include microporous nanoparticles such as silica and zeolites (Figure I.3C). Mesoporous silica (SiO2) has pore diameters of 2-50 nm, while microporous silica has much smaller pores below 2 nm. Zeolites are aluminosilicates with even smaller pore diameters ranging from 0.3-0.8 nm. As with colloidal nanoparticles, polymers, and proteins detailed above, the colloidal stability of 3D porous nanomaterials depends on proper electrolyte concentration,145 surfactant surface coverage,146,147 and polymer coatings.148 Studies remain largely empirical and an underlying mechanism of colloidal stability remains unclear. Representative studies include the finding that mesoporous silica is typically synthesized with a surfactant like hexadecyltrimethylammonium bromide (CTAB) acting as both a structure-directing molecule (to synthesize a specific shape) and a stabilizing capping ligand.149–151 These surfactant-capped materials can be stable for upwards of a year in water.152 Mesoporous silica nanoparticles can also be surface functionalized as hydrophobic, which renders them useful drug delivery agents.153,154 Little is known about the solvation and colloidal stability of mesoporous silica, although recent work suggests that performing dialysis solvent exchange or coating the nanoparticles with proteins improves dispersability.155,156 A potential method for probing mechanisms of colloidal stability in porous materials would rely on surface functionalization. Although porous, silica and zeolites lack chemical tunability beyond their typical inorganic compositions (Figure I.4B). Instead, organic-inorganic framework materials assemble from a wide variety of metallic and organic-compounds, resulting in a diverse 30 range of pore diameters, aperture shapes, solvent-accessible surface area, morphology, and well- developed methods for post-synthetic modulation of surface compositions. I.D.2. Porous Frameworks Understanding the colloidal stability of porous frameworks attracts intense recent attention, in part because 3D porosity challenges conventional mechanisms of solvation, such as the notion of electrostatic forces between smooth, uniform, hard spheres. Porous framework materials metal- organic frameworks (MOFs) are pursued for a wide range of applications due to their unparalleled performance at selective gas separation, water filtration, carbon sequestration, and other areas leveraging tunable guest-host interactions (Figure I.4A). Practical implementation as solution processible and reusable materials, such as thin film membranes, demands their ability to suspend as uniform, stable colloids in a range of solvents. Stability in water would facilitate their utility in biological applications, for example, while compatibility with low-boiling solvents would render them amenable to spraying coating and other forms of industrial production at-scale. In addition to a lack of knowledge of solvation structure, few studies exist for any form of surface functionalization. Reported surface ligands deviate from those well studied with conventional quantum dots.157 Solvent interactions likely dictate the interaction of porous materials with polymers in the so-called mixed matrix membranes envisioned for chemical separation technologies. Little is known about microscopic aspects of the polymer-porous colloid interface, except that great care must be taken to prevent polymers from intercalating and clogging pores,158,159 which can be detrimental to gas sorption and chemical separation applications. In this section, we will describe the current theories of colloidal stability for porous frameworks and the challenges in studying them. Unlike typical, nonporous nanoparticles, MOFs exhibit colloidal stability without the need for conventional capping ligands.160 In fact, while traditional surfactants such as dodecanoic acid or cetyltrimethylammonium bromide can be included in a MOF synthesis to impart size or shape control,161,162 these ligands do not remain with the MOF nanoparticles after washing, as we have observed by nuclear magnetic resonance spectroscopy, thermogravimentric analysis, and other analytical techniques.160,163 Instead, while post-synthetic addition of ligand dyes157 or DNA/protein coronas164 leads to functionalized surfaces, few if any studies detail how they impact colloidal stability. Instead, we recently demonstrated that the solvent identity plays a greater role in 31 stabilizing the surface as only solvents that can dissolve the organic linker can suspend the MOF nanoparticle.160 This result strongly suggests that the interaction of MOF nanoparticles resembles the dissolution of cage molecules, polymers, and other macromolecules systems where solubility depends solvent-monomer energetics and accessible void spaces (Figure I.4C,D). In addition to specific solvent-surface interactions, classical electrostatic arguments of DLVO theory could explain the stability of uncapped MOF nanoparticles, where surface charges would arise from deprotonated linker molecules or open metal sites. Electrostatics alone is unlikely to account for the full mechanism of colloidal stability, however.165 Recent reports of zeta potentials—indicators of net surface charge— typically at the border (±30 mV) of values required for colloidal stability by DLVO theory.160,166 These reduced zeta potentials can be attributed to the low surface density of porous nanoparticles and the weak, short-range electrostatic fields produced by surface defects that quickly decayed through porous channels.167 Binding of solvent or polymers to open metal sites at the surface of MOF nanoparticles improves colloidal stability, Figure I. 4 Representation of A) framework materials, comprised of inorganic/organic building units that are assembled into porous frameworks and B) inorganic porous materials, which are made up of secondary building units. In C) both material classes show distinct internal vs. external surfaces due to their high solvent-accessible surface areas. Direct functional groups or open metal sites within the materials D) may induce different solvent arrangements. 32 while also reducing the apparent surface charge.159,163 In other words, zeta potentials may not relate directly to MOF colloidal stability. Instead of traditional electrostatics, we suggest the intrinsic porosity of nanoMOFs improves colloidal stability through entropic effects. For nonporous colloids, entropy favors aggregation because it removes solvation shells and disorders excluded solvent. This process is termed the “hydrophobic effect” and was introduced as a mechanism to describe porous nanoparticles,168 where the total amount of ordered solvent must be lower than nonporous colloids due to the low density surfaces. As a result, the potential energy favoring aggregation decreases as well. To examine the interaction of solvent with porous interfaces, we use VSFSS to measure the orientation of N,N-dimethylformamide (DMF) or water solvation shells at the exterior of Zn(2- methylimidazolate)2 (ZIF-8) colloids. Although predicted to exist, solvation shells of colloids were previously documented only by atomic force microscopy and x-ray studies.169–173 In addition to providing a protective shell of steric repulsion, we propose that solvent interactions with open metal sites and linker monomer units improves solvation energetics akin to Flory-Huggins theory. We also observed that the bridging 2-methylimidazolate linkers spontaneously align at the solvent interface in a manner that expands the internal pore volumes. This lattice flexibility, like a protein, may further enhance favorable solvent interactions and colloidal stability. Taken together, these recent reports provide a roadmap for preparing and stabilizing porous materials in solution. This thesis describes research which probes the surface structure of colloidal materials that are stabilized by non-DLVO forces, including factors like steric stabilization, low electrostatic potential, sparse surface density, particle porosity, and surface heterogeneity. A background on the surface-specific technique of vibrational sum frequency scattering spectroscopy is provided in the following chapter, laying the groundwork for using this powerful tool to study the interfaces of uniquely stable colloidal materials. 33 CHAPTER II: VIBRATIONAL SUM FREQUENCY SCATTERING SPECTROSCOPY (VSFSS) II.A.VSFSS THEORY Developed by Roke et. al.,174 vibrational sum frequency scattering spectroscopy (VSFSS) is a scattering adaptation of the reflection-based vibrational sum frequency spectroscopy introduced by Shen et. al..175 While reflection VSFS has been commonly used to probe planar air/water176, oil/water177, oxide/water 178, and silica/water179 interfaces, the scattering variation of this technique enables investigation into the interface of nano-sized colloids.180,181 In the scattering experiment, a visible and tunable infrared (IR) pulse are overlapped spatially and temporally at the Figure II. 1 A) Schematic of VSFSS, in which a scattered sum frequency response (blue) is generated by overlapping a visible (green) and IR (red) pulse spatially and temporally at a curved interface. The energy level diagram of VSFSS is shown in the bottom right. B) Cartoon representation of the possible scenarios in which a sum frequency signal could be expected. IR transition dipoles at anisotropic surfaces (top) are sum frequency active while mismatched or directly opposing transition dipoles (bottom) are sum frequency inactive. C) Schematic of the ssp polarization scheme in which the sum frequency, visible beams, and IR beams are s, s, and p polarized, respectively. This polarization combination probes vibrational dipoles perpendicular to an interface. 34 curved interface to generate a scattered sum frequency response that is the sum of the two incoming beams (Figure II.1A). VSFSS directly probes vibrational modes at interfaces due to selection rules under the electric dipole approximation prohibiting contributions from a centrosymmetric environment such as a bulk liquid phase. The signal indicates two main attributes of the chemical environment: the population of molecules at the interface and the average molecular orientation of vibrational modes, also referred to as net molecular ordering. Only highly ordered resonant vibrations generate a sum frequency response (Figure II.1B). In VSFSS, the intensity of the scattered SF response is proportional to the intensities of the incoming visible and IR beams (𝐼𝐼𝑅𝐼𝑣𝑖𝑠) and the square modulus of the second-order nonlinear susceptibility (𝛸(2)) as shown in Equation 5. 𝐼𝑆𝐹 ∝ 𝐼𝐼𝑅𝐼𝑉𝑖𝑠|𝜒(2)| 2 (5) In a centrosymmetric environment where all directions are equivalent, 𝜒𝑖𝑗𝑘 (2) is identical in two opposite directions such as: 𝜒𝑖𝑗𝑘 (2) = 𝜒−𝑖−𝑗−𝑘 (2) . Additionally, 𝜒𝑖𝑗𝑘 (2) is a third rank tensor and obeys the property that changing the sign of the three subscripts is the same as reversing the axis system like: 𝜒𝑖𝑗𝑘 (2) = −𝜒−𝑖−𝑗−𝑘 (2) . To satisfy both of the above symmetry properties, 𝜒𝑖𝑗𝑘 (2) must equal 0, and thus a SF response in a centrosymmetric environment is forbidden. Therefore, VSFS is a surface specific technique as the interface of two materials in non-centrosymmetric and results in non-zero tensor elements. VSFS involve 3 electric fields (visible, IR, and SF), thus 𝜒(2) is a 3x3x3 tensor with 27 elements. These elements are reduced through symmetry considerations to result in seven non- zero elements that contribute to a SF response. Only four of these elements are unique. These contributions physically represent different orientations of the vibrational mode dipoles relative to the interface and can be probed through a variety of polarization combinations. The possible polarization combinations and the respective 𝜒𝑖𝑗𝑘 (2) element(s) that contribute to the spectrum are: pss (𝜒𝑧𝑦𝑦 (2) ), sps (𝜒𝑦𝑧𝑦 (2) ), ssp (𝜒𝑦𝑦𝑧 (2) ), and ppp (𝜒𝑧𝑧𝑧 (2) , 𝜒𝑧𝑥𝑥 (2) , 𝜒𝑥𝑧𝑥 (2) , 𝜒𝑥𝑥𝑧 (2) ), where the polarizations are listed in the order of SF, visible, and infrared. The resonant portion of the second-order susceptibility is dependent on both the net orientation and population of molecules at the interface (Equation 6). 35 𝜒𝑅 (2) = 𝑁 𝜀0 〈𝛽𝑉〉 (6) Where 𝑁 is the number of molecules at the interface, 𝜀0 is the vacuum permittivity, and 〈𝛽𝑉〉 is the average macroscopic hyperpolarizability which accounts for the orientation of molecules at the interface. The 𝑋(2) has non-resonant and resonant terms which are accounted for in the fitting equation (7) as developed by Bain et al. |𝛸(𝜔)(2)| 2 = |𝛸𝑁𝑅 (2) 𝑒𝑖𝜙 + ∑ ∫ 𝐴𝑣𝑒𝑖𝜙𝑣𝑒 −( 𝜔𝐿−𝜔𝑣 Г𝑣 ) 2 𝜔𝐿 − 𝜔𝐼𝑅 + 𝑖Г𝐿 𝑑𝜔𝐿 +∞ −∞𝑣 | 2 (7) where the amplitude of the non-resonant susceptibility is described by 𝛸𝑁𝑅 (2) with a phase 𝜙.182 The summation of all vibrational transitions that are SFG active describes the resonant susceptibility, where 𝐴𝑣 is the peak amplitude, 𝜙 is the phase, Г𝐿 is the Lorentzian linewidth describing homogenous broadening, and Г𝑣 is the Gaussian linewidth describing inhomogeneous broadening. Equation 5 is used to fit all experimental spectra. A practical aspect of VSFSS is the ability to change the polarization of the incoming and collected beams to isolate vibrational responses from unique molecular orientations relative to the interface. A three-letter nomenclature denotes the polarization of the beams in the sum- frequency, visible, and IR beams, respectively (Figure II.1C). In this dissertation, ssp, sps, and ppp polarization schemes are used. The ssp polarization scheme probes IR transition moments perpendicular to the interface while the sps polarization probes IR transition moments aligned parallel to the interface. The ppp polarization combination probes a combination of transition moments in both the perpendicular and parallel contributions. II.B. LASER SETUP The VSFSS system used to collect the data presented in this dissertation is engineered as follows and in Figure II.2. A Ti:Sapphire regenerative amplifier laser (Coherent Libra) generates an 800 nm, ~80 fs fundamental pulses with a 1 kHz repetition rate. A portion of that beam is used as the visible pulse while the remaining is sent through an optical parametric amplifier (Coherent OPerA Solo) to generate a broadband infrared (IR) beam through difference frequency 36 generation. The visible and IR pulses are then directed along a series of mirrors, lenses, and polarizers to become spatially and temporally overlapped at the sample stage. The sample cell is composed of a CaF2 window in the front and a quartz cuvette in the back (Helma QS) with an optical path length of 200 µm. The IR beam is focused at the sample to a spot size of ~80 µm with a parabolic gold mirror, while the visible beam is focused right after the sample cell to a ~500 µm spot size. The scattered SFG response is collected at an angle of ~60°, collimated with a plano- convex lens, and focused into a spectrograph and accompanying charge-coupled device intensifier (Princeton Instruments IsoPlane SCT320 and PI-MAX4). For all experiments, the visible pulse energy was 25 µJ while the IR pulse energies were 2-3 µJ and 5 µJ for the C=O and C-H stretching regions, respectively. To account for daily fluctuations in laser power, a single trace in a figure is the result of at least 3 averaged trials across different days. One trial consists of 2 signal and 2 background measurements. Each trial is averaged, background subtracted, and normalized by a non-resonant response from KNbO3 to account for changes in IR spectral shape. In the C-H region, spectra are further normalized by the integrated SFG intensity (from 2800-3000 cm-1) generated from a d- hexadecane emulsion in D2O stabilized with 1 mM AOT. Each trial is further normalized by the size of the droplet (determined by a monomodal distribution from DLS) through a scattering pattern developed by Roke et. al. that accounts for the percentage of scattered signal collected at 60° which is dependent on droplet size and beam polarizations.183,184 More details on the Figure II. 2 Schematic of the VSFSS experiment utilized in this dissertation with visible (green), IR (red), and SF (blue) beam paths. Mirrors are omitted for clarity. 37 normalization procedure for VSFSS utilized in the Richmond/Scatena laboratory can be seen in a in Appendix G.185 Additionally, a calculation of the particle concentration necessary to overcome the signal to noise threshold for VSFSS is provided in Appendix H. The above procedure was used to collect all of the VSFS spectra presented in this dissertation and are not repeated in the experimental sections of each chapter for clarity. Chapter III begins the experimental section of this thesis by using VSFSS to study the surface structure of emulsions stabilized with a steric layer of polymer. 38 CHAPTER III: EMULSIONS STABILIZED BY POLY(ACRYLIC ACID) This work was published in volume 697 of the journal Colloids and Surfaces A: Physiochemical and Engineering Aspects in June 2024. Ashley N. Mapile is credited with conceptualization, formal analysis, investigation, writing – original draft, writing – review and editing, and visualization. Lawrence F. Scatena assisted in conceptualization, writing – review and editing, supervision, and funding acquisition. Supplementary information for this chapter is provided in Appendix A III.A. AN INTRODUCTION TO POLY(ACRYLIC ACID) As explained in the introduction to this thesis, polyelectrolytes are useful analogues to studying protein coronas that are known to stabilize colloids due to the tunable nature of polymers. Many characteristics of polymers, such as the chain length (via molecular weight), chemical functionality, % protonation, and morphology can all be adjusted synthetically or by environmental conditions. Poly(acrylic acid) (PAA, Figure III.1) is a simple polyelectrolyte that can mimic protein behavior at interfaces. PAA, which contains a single carboxylic acid repeating group, can expand in solution when deprotonated or hydrogen bond with itself to form coils when protonated and is a useful material in forming zwitterionic polymer coarcervates.186–190 Through pH- dependent zeta potential measurements (Appendix A, Figure A1) the interfacial pKa of PAA at the hexadecane/water emulsion interface was determined to be ~5.3, about 0.8 units more alkaline Figure III. 1 Chemical structure of poly(acrylic acid) and percent protonation curve as a function of pH for surface (solid, orange line) and bulk (dotted, purple line). Percent protonation was calculated with the Henderson-Hasselbalch equation, the known bulk pKa (4.5), and the experimentally derived interfacial pKa (5.3). 39 than the bulk pKa (Figure III.1). Therefore, studying the pH-dependent conformational behavior of PAA adsorbed at the emulsion interface can provide insight into droplets stabilized by pH- sensitive protein coronas as well as a detailed understanding of the mechanisms of colloidal stability. The adsorption behavior of PAA without the addition of a surfactant, has been studied at both the planar CCl4/water and air/water interfaces using vibrational sum frequency spectroscopy (VSFS). These studies found that in conditions above the pKa (pH >4.5), the polymer is interfacially inactive, but the addition of charge-screening cations promotes adsorption of the polymer to the interface.105,191–194 Below pH 4.5, the polymer is interfacially active and exhibits a well-ordered structure as evident by strong CH and C=O VSFS spectral features. At a planar interface, molecular dynamics are thermodynamically determined by equilibrium surface activity and surface-active molecules develop a stable configuration given time. On the other hand, emulsions require the input of external energy to form droplets resulting in a kinetically trapped interface.98,101 During the formation of emulsions, surface-active molecules initially adsorb to the interface via ultrasonication but subsequently adsorb/desorb through equilibrium thermodynamics after formation of the curved interface. Thus, the conformational ordering of interfacial molecules adsorbed to emulsions is dictated by the thermodynamic equilibrium of molecular adsorption and the kinetic energy encouraging the platform of these colloids to destabilize and coalesce. Stabilization of emulsions is best described by the theory developed by Derjaguin, Landau, Verwey, and Overbeek (DLVO) which states that both electrostatic and steric repulsion prevents droplet coalescence and keeps the emulsions in a kinetically trapped state.8,98,101,195 Reports of emulsions coated with solid polymer particles (such as Pickering emulsions),196 nanoparticles with a polymeric coating,197 and emulsions formed with surfactant/polymer mixtures198 suggest unique behavior at curved hydrophobic interfaces as opposed to their planar counterparts. While there have been reports of emulsions stabilized with polymer alone, observed through Janus particles199 or magnetic nanoemulsions,200 and specifically emulsions coated with diblock copolymers,201,202 they lack insight into the detailed mechanism of stability promoted by polymers or the molecular level information to provide a structural understanding of the droplet interface. Here, we demonstrate the stability of emulsions coated with PAA and provide a picture of polymer molecular conformation with the surface-specific vibrational sum frequency scattering spectroscopy (VSFSS). Despite the lack of surface charge thought to be necessary for droplet 40 stabilization, PAA-coated emulsions prepared at pH 2 are stable and exhibit interfacial polymer ordering. Surprisingly, the highly charged pH 6 polymer does not stabilize emulsions and exhibits no molecular organization. With the addition of salt, we observe charge screening at the interface and a disruption to polymer ordering because of localized electrostatics. We propose that emulsions are stabilized through the steric hindrance of an adsorbed PAA layer, whose molecular conformation and charge-dependent behavior can provide insight to protein corona-stabilized droplets and their application as drug delivery vehicles. III.B. MATERIALS AND METHODS III.B.1. Materials All materials were used as delivered without further purification. Hexadecane (≥99%), dioctyl sodium sulfosuccinate (AOT, ≥97%), poly(acrylic acid) (PAA, average Mv ~ 450,000), calcium chloride (CaCl2, anhydrous, ≥97%), magnesium chloride (MgCl2, anhydrous, ≥98%), sodium chloride (NaCl, BioXtra, ≥99.5%), and ammonium chloride (NH4Cl, ACS reagent, ≥99.5%) were purchased from Sigma-Adrich. Deuterated hexadecane (n-hexadecane-d34, 98.6% D), sodium deuteroxide (NaOD, 99.5% D), and deuterium chloride (DCl, 99.8% D) were purchased from CDN Isotopes. Deuterium oxide (D2O, 99.9% D) was purchased from Cambridge Isotope Labs. All glassware was copiously cleaned in a bath containing sulfuric acid (98% Sigma-Aldrich) and AlNOCHROMIX oxidizer from Godax Laboratories Inc. After sitting in the acid bath for at least 24 hours, glassware was rinsed for at least 2 min. with 18.2 MΩ-cm water and dried in an oven. III.B.2. Emulsion Formation Emulsions consisting of hexadecane (or d-hexadecane) suspended in water (or D2O) were prepared by ultrasonication (Branson Sonifier 250) of the sample at 5% output power (~12.5%) at 20 kHz for 5 minutes at a constant duty cycle. The ultrasonication probe tip was placed at the interface of the aqueous solution and oil layer to ensure homogenous mixing. The AOT-coated emulsion standard was formed by ultrasonicating AOT stock solution with oil to give a 1mM AOT emulsion with 2.5% v/v oil in water. PAA-stabilized emulsions were formed in the same fashion with DCl or NaOD added to reach the desired pD (pH). Emulsion pH was measured using MilliporeSigma MColorpHast pH strips and an Oakton Instruments Portable Meter Kit. The detection of pH values can differ from pD values by 0.43,203 however for ease of communication, pH is used in this paper. 41 III.B.3. Dynamic Light Scattering and Zeta Potential Emulsion hydrodynamic diameter (Z-average), polydispersity index (PDI), and zeta potential were measured by a Malvern Zetasizer Nano ZS. Details regarding the theory and practice of dynamic light scattering (DLS) and zeta potential can be found elsewhere.204,205 From DLS measurements, hydrodynamic diameter and PDI was reported from an average of at least three measurements. DLS and zeta potential measurements were collected by pipetting 1 mL of emulsion solution into a Malvern folded capillary zeta cell, therefore maintaining consistent v/v oil concentrations. Zeta potential values were reported from an average of at least five measurements. Determination of the surface pKa using zeta potential measurements is detailed in the supporting information (Figure S1) and follows a procedure developed by Haes et. al.206 III.C. RESULTS AND DISCUSSION III.C.1. Impact of PAA Concentration Emulsions were formed with concentrations of PAA ranging from 1 ppm to 4000 ppm in the natural pH conditions of PAA (pH 4) and measured by dynamic light scattering and zeta potential (Figure III.2). For context, previous experiments at the CCl4/water interface used 5 ppm PAA while air/water interface experiments used 4500 ppm.105,191 Figure A2 and Table A1 demonstrate that emulsions prepared at or below 250 ppm PAA were unstable after 1 week (Appendix A). Bare emulsions (those without deliberately added surface active agents) are known to have a negative zeta potential due to miniscule amounts of surface active contaminants.180,207 The zeta potential of our hexadecane/water emulsions (-32 mV) suggest minor surface active impurities. The presence Figure III. 2 Characteristics A) hydrodynamic diameter, B) polydispersity index, and C) zeta potential) of emulsions coated with varying amounts of PAA at pH 4. Points shaded orange indicate 500 ppm PAA. 42 of these impurities were identified in previous emulsion studies from our laboratory that employed meticulous and extensive cleaning procedures to minimize their presence.180 After PAA addition, the polymer is much more surface active than the impurities and displaces any contaminants that contributed to the negative surface charge. As additional PAA is added, charge screening and hydrogen bonding causes a reduction in the magnitude of the zeta potential. At a concentration of 500 ppm, PAA-stabilized emulsions exhibit good colloidal stability, moderately polydisperse size distributions, and a highly charged surface layer as indicated by the orange shaded points in Figure III.2. Additionally, as confirmed by pendant drop surface tensiometery, discussed in Appendix A and Figure A3, PAA adsorbs to the hexadecane/water interface at a concentration of 500 ppm. Emulsions formed using 250 ppm PAA and below resulted in unstable droplets presumably due to a lack of surface population. Despite a remarkably high magnitude zeta potential (≥30 mV) emulsions stabilized with 100 ppm PAA or below are unstable and crash out quickly after formation, likely due to a thin or completely absent polymer layer. While the addition of PAA at concentrations above 500 ppm leads to more stable emulsions, as indicated by the long-term stability, significant charge screening occurs making them difficult targets for measuring the interplay between electrostatics and steric repulsion when salt ions are added. The slight increase in emulsion size and reduction in zeta potential with increasing concentration of PAA is due to polymer layering at the surface which extends the hydrodynamic diameter and charge screens the surface as is consistent with other polymer-layering studies.16,201,208,209 Emulsions stabilized with 1500 ppm PAA or higher have a low zeta potential (~-7 mV) and are colloidally stable up to a week, solidifying the observation that sterics alone are enough to stabilize emulsions. Due to the desire to study the impact of surface charge screening, the emulsions used in this paper are formed with 500 ppm PAA as they have long-term colloidal stability, are moderately polydisperse and have appreciable interfacial charge. III.C.2. Emulsions Stabilized by PAA Alone Emulsions coated with 500 ppm PAA were formed at pH 2, pH 4, and pH 6 via ultrasonication of 2.5% v/v hexadecane in water. Pendant drop surface tensiometery measurements indicate that high amounts of PAA adsorb to the hexadecane/water interface at pH 2 and pH 4 while little adsorption is observed at pH 6 (Figure A4). Emulsions at pH 2 and pH 4 maintained colloidal stability after one week, as evident by their visual turbidity, while pH 6 oil droplets phase separated (Figure 43 A4). Dynamic light scattering of pH 2, pH 4, and pH 6 colloids measured 20 minutes after formation revealed hydrodynamic diameters of 477.7, 363.2, and 761.0 nm, respectively, demonstrating that the average size of emulsions prepared at pH 2 and 4 are in the nanosized regime while pH 6 emulsions are inherently larger. Narrow, monodisperse colloids have a PDI < 0.1, while polydisperse systems have a PDI > 0.1.210 Moderate polydispersity is defined by a PDI from 0.1-0.4 and describes the pH 2 and pH 4 emulsions studied `here (PDIs of 0.247 and 0.272, respectively), while the pH 6 emulsions have broad polydispersity (PDI of 0.628). Similarly, size distribution plots from DLS measurements are monomodal for emulsions at pH 2 and pH 4, but multimodal for pH 6 (Figure A5). Table S2 lists the PDI of each emulsions system studied by VSFSS which are in line with recent studies on emulsion formulations.201