<<

The Pennsylvania State University

The Graduate School

Department of Chemical Engineering

ZETA POTENTIALS AND MINERAL REPLACEMENT AT SURFACES IN

SATURATED SALT SOLUTIONS

A Dissertation in

Chemical Engineering

by

Astha Garg

 2017 Astha Garg

Submitted in Partial Fulfillment of the Requirements for the Degree of

Doctor of Philosophy

August 2017 ii

The dissertation of Astha Garg was reviewed and approved* by the following:

Darrell Velegol Distinguished Professor of Chemical Engineering Dissertation Advisor Chair of Committee

Ali Borhan Professor of Chemical Engineering

Manish Kumar Assistant Professor of Chemical Engineering

Ayusman Sen Distinguished Professor of Chemistry

James H. Adair Professor of Materials Science and Engineering, Biomedical Engineering and Pharmacology

Janna Maranas Professor of Chemical Engineering and Materials Science and Engineering Graduate Program Coordinator of Chemical Engineering

*Signatures are on file in the Graduate School

iii

ABSTRACT

Surfaces, fixed or mobile, exposed to salt solutions interact with the solution and with other surfaces through physical interactions, chemical reactions and transport. Understanding these phenomena is key to discerning the system behavior and subsequently manipulating it. A good handle on saturated salt systems can be of immense benefit to modern science and industry. However, their behavior is not adequately understood, especially from a colloidal and surface chemistry perspective. The complexity of interactions under saturated salt conditions and experimental difficulties in carrying out certain measurements have contributed to our limited understanding in this regard. The subject of this thesis is to characterize and transform the nature of surfaces exposed to saturated salt solutions through an in-depth understanding of the , fluid transport and dissolution- precipitation phenomena in these systems.

The work presented here is conceptually focused on two subjects – and mineral replacement. First, I describe careful measurements and interpretation of zeta potential which is a measure of electrical potential at the surface-solution interface. Second, I characterize mineral dissolution-precipitation reactions in terms of fluid flow, the rate and extent of mineral replacement and porosity, under a variety of controlled conditions. These experiments offer insights into the behavior of systems at saturated salt and provide quantitative measurements of key parameters that can be used to evaluate and manipulate the system behavior.

iv TABLE OF CONTENTS

List of Figures ...... x

List of Tables ...... xix

Acknowledgements ...... xx

Chapter 1 Motivation and Research Goals ...... 1

1.1 Motivation ...... 1 1.2 Research Goals ...... 6 1.2.1 Zeta potential at high salt and saturated salt conditions ...... 7 1.2.2 Mineral replacement in the KBr-KCl system ...... 10 1.2.3 Chemical micro-fracking of calcite through mineral replacement ...... 14 1.3 Organization of the thesis...... 16 1.4 References ...... 17

Chapter 2 Theoretical background on Electrokinetics ...... 20

2.1 Definition of zeta potential and the electrical double layer ...... 20 2.1.1 Examples and estimates ...... 21 2.2 Applications of zeta potential...... 22 2.3 Classical theory of the EDL ...... 23 2.3.1 Spatial distribution of charge – the Poisson-Boltzmann equation ...... 23 2.3.2 Relationship between charge density and surface potential – the Gouy- Chapman model ...... 24 2.4 Methods of zeta potential measurement ...... 25 2.4.1 based techniques ...... 25 2.4.2 Streaming potential ...... 26 2.4.3 Electroacoustics ...... 26 2.5 Colloidal stability ...... 27 2.5.1 DLVO theory...... 28 2.5.2 Rapid aggregation ...... 29 2.6 Electrokinetics ...... 30 2.6.1 Electrophoresis and electro-osmosis ...... 30 2.6.2 Diffusiophoresis and diffusioosmosis ...... 32 2.7 References ...... 33

v Chapter 3 Particle Zeta Potentials Remain Finite in Saturated Salt Solutions ...... 36

3.1 Abstract ...... 36 3.2 Introduction ...... 37 3.3 Methods and Materials: ...... 41 3.3.1 Measuring Zeta Potentials ...... 41 3.3.2 Experimental Details ...... 42 3.3.3 Particle tracking and analysis ...... 46 3.3.4 Method Validation ...... 49 3.3.5 Monte Carlo Simulations of the EDL...... 52 3.4 Results & Discussion ...... 53

3.4.1 Particle Motions at Frequency 20 ...... 53 3.4.2 Zeta potential measurements of sPSL and aPSL ...... 57 3.4.3 Stout and Khair zeta potential ...... 61 3.4.4 Monte Carlo simulation of the EDL ...... 62 3.5 Summary and Conclusions ...... 68 3.6 Author Contributions ...... 69 3.7 Copyright Notice ...... 69 3.8 Acknowledgements ...... 69 3.9 References ...... 70

Chapter 4 Relative Roles of Kinetics, Transport and Thermodynamics in Pseudomorphic Mineral Replacement ...... 77

4.1 Abstract ...... 77 4.2 Introduction ...... 78 4.3 Background ...... 80 4.3.1 Steps in the replacement process ...... 80 4.3.2 Thermodynamics - Lippmann Diagrams ...... 82 4.3.3 Porosity generation ...... 85 4.4 Questions ...... 86 4.4.1 Passivation ...... 86 4.4.2 Solution to solid ratios ...... 87 4.4.3 Porosity ...... 87 4.4.4 Fluid transport ...... 87 4.4.5 Interfacial ...... 88

vi 4.5 Methods and Materials ...... 88 4.5.1 Replacement experiments ...... 88 4.5.2 Fluorescence optical microscopy ...... 89 4.5.3 Electron Microscopy ...... 90 4.5.4 Elemental analysis ...... 90 4.6 Results and Discussion ...... 91 4.6.1 Replacement rate of KCl vs KBr ...... 91 4.6.2 Varying solution to solid ratios ...... 93 4.6.3 Effect of external transport ...... 98 4.6.4 Replaced layer composition for replaced KCl...... 98 4.6.5 Replaced layer composition for KBr ...... 102 4.6.6 Accuracy of the EDS composition measurements ...... 104 4.6.7 Solution composition as a function of distance ...... 107 4.7 Conclusions ...... 109 4.7.1 Passivation ...... 109 4.7.2 Solution to solid ratios ...... 110 4.7.3 Porosity ...... 111 4.7.4 Fluid transport ...... 112 4.7.5 Interfacial concentration ...... 113 4.8 Contributions ...... 114 4.9 Acknowledgements ...... 114 4.10 References ...... 115

Chapter 5 Chemical microfracking - Enhancing calcite porosity through pseudomorphic mineral replacement ...... 118

5.1 Abstract ...... 118 5.2 Objectives...... 119 5.3 Background ...... 121 5.3.1 pMRR of calcium carbonate...... 121 5.3.2 Porosity measurement ...... 122 5.4 Methods and Materials ...... 123 5.4.1 Crystals ...... 123 5.4.2 Buffer solutions ...... 124 5.4.3 Replacement experiments ...... 125

vii 5.4.4 Electron Microscopy and Elemental Mapping ...... 128 5.4.5 Image processing for rim width measurement ...... 128 5.4.6 Mercury porosimetry ...... 129 5.4.7 Dye penetration using confocal microscopy ...... 130 5.4.8 Calculations using VMinteq ...... 130 5.5 Results and discussion ...... 131 5.5.1 Effect of temperature on replacement at pH 8.11, 200 C vs. 50 C ...... 131 5.5.2 Effect of pH on observed replacement ...... 132 5.5.3 Replacement at pH 5 at 50 C ...... 138 5.5.4 Replacement with time ...... 139 5.5.5 Effect of external transport ...... 140 5.5.6 Effect of Phosphate concentration ...... 143 5.5.7 Effect of Ca+2 addition...... 144 5.5.8 Mercury porosimetry at 6 days...... 145 5.5.9 Confocal at 6 days and 12 days. For nicely cut vs. hammered...... 146 5.6 Conclusions ...... 148 5.7 Future work and improvements to experimental setup ...... 149 5.8 Acknowledgements ...... 150 5.9 References ...... 151

Chapter 6 Future Work and Broader Impact ...... 154

6.1 Zeta potential at high and saturated salt ...... 154 6.1.1 Scientific advance ...... 154 6.1.2 Implications for saturated salt systems ...... 155 6.1.3 Broader applicability of technique to biological systems...... 155 6.1.4 Future directions ...... 156 6.2 Pseudomorphic mineral replacement reactions ...... 157 6.2.1 Scientific advance ...... 157 6.2.2 Application to enhanced oil recovery ...... 157 6.2.3 Applicability to synthesis of templated or doped materials ...... 158 6.2.4 Future research directions ...... 159 6.3 References ...... 159

Appendix A Supplemental Information for Zeta at High Salt ...... 160

viii A.1 Lattice Model of the Electric Double Layer ...... 160 A.2 Continuum Model of the Electric Double Layer ...... 164 A.3 Summary of zeta potential measurements ...... 167 A.4 Contributions ...... 169 A.5 Copyright Notice ...... 169 A.6 References ...... 169

Appendix B Boundaries can steer active Janus spheres...... 170

B.1 Abstract ...... 170 B.2 Introduction ...... 171 B.3 Results and Discussion ...... 172 B.3.1 Experimental Characterization of the Orientational Quenching ...... 172 B.3.2 Steering the Active ...... 175 B.3.3 Theoretical description of the catalytic colloid near a surface ...... 178 B.3.4 Electrostatic colloidal forces...... 179 B.3.5 Phenomenology...... 183 B.4 Conclusion ...... 193 B.5 Discussion: exploring the limits for boundary steering phenomena ...... 194 B.6 Methods ...... 197 B.6.1 Janus Particle Preparation ...... 197 B.6.2 Sample Preparation and Gravitaxis of Janus ...... 197 B.6.3 Microscopy and Image Analysis ...... 199 B.6.4 Colloid settling and diffusion ...... 200 B.6.5 Electrophoresis/Rotational Electrophoresis ...... 201 B.6.6 Lithographic Manufacture of Rectangular Channels ...... 205 B.7 Acknowledgements ...... 207 B.8 Contributions ...... 208 B.9 References ...... 208

Appendix C Origins of concentration gradients for diffusiophoresis ...... 214

C.1 Abstract ...... 214 C.2 Introduction ...... 215 C.3 Essential Theory of Diffusiophoresis ...... 218 C.4 Origins of Concentration Gradients ...... 225

ix C.4.1 Molecular Exclusion ...... 225 C.4.2 Chemical Reaction ...... 231 C.4.3 Diffusive mixing ...... 233 C.4.4 Externally-imposed ...... 237 C.4.5 Salt or crystal dissolution ...... 241 C.4.6 Molecular Crystallization ...... 244 C.4.7 Sedimentation and screening ...... 247 C.4.8 Evaporation and condensation ...... 248 C.5 Conclusions ...... 250 C.6 Acknowledgments ...... 251 C.7 Contributions ...... 251 C.8 Copyright Notice ...... 252 C.9 References ...... 252

x LIST OF FIGURES

Figure 1-1: Zeta potential, dissolution and precipitation in a geological reservoir. A simplified view of a geological reservoir that has charged colloids suspended in a saturated salt solution. The calcite surface and NaCl crystals participate in dissolution – precipitation equilibria...... 2

Figure 1-2: Charged surfaces in salt concentration gradients can cause flows in dead-end pores through diffusiophoresis...... 3

Figure 1-3: Our hypothesis of chemical-mechanical fracking. pMRR at the fractured surfaces would create additional porosity, opening up more volume for fluid flow...... 5

Figure 1-4: Zeta potential and electrical double layer for a negatively charged particle in water ...... 8

Figure 1-5: Electrophoresis of a particle with positive zeta potential. Particle velocity, u is directly proportional to the electric fields and zeta potential on the particle’s surface. ... 9

Figure 1-6: Schematic of mineral replacement of parent AB by a porous layer of the guest AC, mediated by a solution of AC. For replacement to proceed, the interfacial fluid contacting the phase AB should be saturated...... 11

Figure 1-7: Camera and electron microscopy images of replacement in the KBr-KCl system. KBr crystal in saturated KCl solution shows replacement of ~ 2 mm over 24 hours, while the KCl crystal shows replacement of ~ 0.2 mm over the same time...... 12

Figure 2-1: Zeta potential and electrical double layer for a negatively charged particle in water ...... 21

Figure 2-2: A sample DLVO plot. The parameters are :  potential = 100 mV, = 1 nm, particle diameter 2 µm, Hamaker constant = 1.4 x 10-19 J...... 29

Figure 2-3: Fluid and particle velocity when an is applied on fluid in a closed capillary...... 31

Figure 3-1: (a) Schematic illustration of the experimental setup showing polystyrene latex (PSL) particle in a glass capillary subject to an oscillating electric field. The electrolyte solution is characterized by its viscosity (), conductivity (ke), (), and salt concentration (n0). (b) The Debye length (-−1) decreases as a function of NaCl concentration becoming smaller than the Bjerrum length () at 0.3 M and the ions themselves (ca. 0.3 nm) above 2 M. (c) An illustration of the EDL at two different salt highlighting the relevant length scales such as ion size, Debye length (shown on left and right edges for low and high salt, respectively), ion-ion spacing, and surface charge spacing (red semi-circles)...... 40

Figure 3-2: Electronics and instrumentation. A function generator generated a sine wave that was amplified by the amplifier. The from the amplifier was connected across the sample cell, and a Keithley source-measure-unit recorded the current

xi through a labview interface. The high speed camera was synchronized with the function generator through an external trigger port...... 46

Figure 3-3: Data analysis and filtering. (a)Fourier transform of the current data i(!) shows a peak only at the applied frequency. This observation suggests that particle motions at frequency 2!0 cannot be attributed to effects that depend linearly on the current. (b) Raw position vs. time for sPSL, showing that 0f0 velocity is so high that all other information is hidden due to it. (c) The filtered position, obtained by subtracting cubic splines fitted to the data split into 24 adjacent sections. This removes information only at low frequencies, f ≤ 24 Hz. The filtered position was used to calculate the fourier transforms shown in Figure 3-5 c,d. All data are for sPSL in 1 M NaCl, f0 = 300Hz and E = 829 V/m...... 47

Figure 3-4: (a) Mobility (u/E) vs. salt concentration for sulfated PSL in various salts. Error bars are obtained from 90% confidence intervals. (b), (c) data compiled for (b) viscosity25–27] and (c) relative permittivitys28–32; at T = 25 ◦C. (d) Measured conductivity for each of the salt solutions at T = 25 ◦C...... 48

Figure 3-5: Filtered position x(t) vs. time for a 2.9 m sPSL particle in (a) 10 mM NaCl for E = 1120 V/m and (b) 5.4 M NaCl for E = 733 V/m. Magnitude of the Fourier components for particle velocity |X()| as a function of frequency for (c) 10 mM and (d) 5.4 M NaCl, respectively. The red dot highlights the peak at the applied frequency f0 = 300 Hz. At high salt concentrations, an additional peak is seen at 600 Hz and attributed to field-induced thermal convection (section 3.4.1)...... 50

Figure 3-6: Phase of Fourier transform of position at f0 for (a) sulfated and (b) amidine- functionalized polystyrene latex particles. A phase of 0 implies ζ < 0 and a phase of ±π implies ζ > 0. For situations where the phase was neither of ±π or 0, the measurement was considered as noise...... 50

Figure 3-7: If the frequency of electric field is low (∼ 1 Hz), the ζ measured would be a function of the frequency22 due to electro-osmotic flow (EOF) from the capillary walls being transmitted to the center. (a) We show here that the measured ζ potential of sulfated PSL particles does not depend on frequency for f0 ≥ 300 Hz. Lower frequencies were not applied because electrode degradation led to strong flows. (b) The particle velocity at the applied frequency is linear with E (i.e., mobility does not depend on electric field). This validates that we are operating in the linear response regime...... 51

Figure 3-8: (a) and (b) X(20) plotted against electrical power for sPSL and aPSL for the salt concentrations where the non-linear effect was statistically significant (1-5 M). 2 The X(20) peaks are roughly linear with power (E ). (c) and (d) X(20) plotted against the velocity at zero frequency for sPSL and aPSL in NaCl for the same concentrations, showing that the two are correlated as can be expected from an E2 effect. (e) Phase for each peak at 20 showing that X(20) is in phase with the applied electric field, which is also a sine wave...... 56

Figure 3-9: (a) Zeta potentials of 2.9 m sPSL particles vs. concentration in five different monovalent salts. (b) Zeta potentials of 3.3 m aPSL particles vs. concentration in

xii

four different monovalent salts. Only data for high salt concentrations  100 mM are shown; see Figure 3-10 for a larger salt concentration range. The error bars represent 90% confidence intervals. The green curves show the prediction of the GC model assuming different surface charge densities. The solid curves are fits obtained for data in NaCl...... 58

Figure 3-10 ζ potential plotted against NaCl salt concentration on a logarithmic scale, for each of the 3 particles tested (circles) along with chi-squared fits for each particle (solid lines). The number below each particle name is the best fit value of σ...... 60 figure 3-11: (a) schematic illustration of the lattice model of a charged surface in a symmetric electrolyte. (b) within the simulation cell averaged over the 1 and 2-directions. the solid markers represent the average of 12 independent monte carlo simulations; the solid curves are predictions of the gc model; the x markers denote the zeta potentials. the simulation parameters are charge density, 휎 = 0.0625푒/ℓ2 and bjerrum length, 휆 = 2ℓ within a cell of dimensions 퐿1 = 퐿2 = 8ℓ and 퐿3 = 32ℓ; the salt concentrations are 푛푐 = 0.005ℓ − 3 (low salt) and 푛푐 = 0.32ℓ − 3 (high salt). (c) zeta potential vs. concentration for different three different  corresponding to conditions similar to experiments on spsl. the markers are results of the mc simulations; solid curves are the gc predictions of equation (3). simulation parameters are the same as in (b) unless stated otherwise...... 63

Figure 3-12: (a) Electric potential near a highly charged surface averaged over the 1 and 2- directions. The solid markers represent the average of 12 independent Monte Carlo simulations; the solid curves are predictions of the GC model; the x markers denote the zeta potential. The simulation parameters are 휎 = 0.0188푒/ℓ2, 휆 = 4ℓ, and 푛푐 = 0.32ℓ − 3 within a cell of dimensions 퐿1 = 퐿2 = 8ℓ and 퐿3 = 32ℓ . (b) Electric potential in the plane 푥3 = ℓ for a particular realization of the surface charge distribution; other parameters correspond to those in (a). (c) Effective zeta potentials 2 vs. concentration for three different  corresponding to conditions similar to experiments on aPSL. The markers are results of the MC simulations; solid curves are the GC predictions of equation (3). Simulation parameters are the same as in (a) unless stated otherwise...... 65

Figure 4-1: Lippmann diagram for the KCl-KBr system.The y axis is the sum of activity products of KBr and KCl and the x-axis shows both xKBr (mole fraction of KBr in solid) - and Br,aq (aqueous activity fraction of Br ). The plot consists of data generated by using a mathematical formulation of the thermodynamics as explained by Pollok et al15(p 217-220). Mathematica was used to calculate the curves...... 83

Figure 4-2: Solubility phase diagram for the KBr-KCl system. The top curve represents solid concentrations and the bottom cure, corresponding solution concentrations, connected by horizontal tie-lines. The curves were calculated using the mathematical formulation of the thermodynamics as explained by Pollok et al15 (p 217-220)...... 85

Figure 4-3: Electron microscopy images of partially replaced cross sections of (left) KCl crystal in saturated KBr solution for 푛퐵푟( 푎푞)푛퐾퐶푙(푠) = 50 and (right) KBr crystal in saturated KCl solution for 푛퐵푟( 푎푞)푛퐾퐶푙(푠) = 150, both at t = 30 mins. The scale bars are both 100 m...... 92

xiii Figure 4-4: Rim width vs. solution to solid ratio. Fluorescence optical microscopy images of dried cross-sections of partially replaced KBr (top row) and KCl (bottom row) crystals, at times 30 mins and 5 hours respectively, for varying solution to solid ratio shown at the bottom left corner for each image...... 93

Figure 4-5: Solubility phase diagram for KBr-KCl annotated with the solution to solid ratios teted in experiments. The green lines correspond to KBr crystal in KCl experiments, and red lines to KCl crystal in KBr...... 95

Figure 4-6: Replaced rim thickness for KCl crystals in saturated KBr solution for various times and solution to solid ratios. Note that the 1.75 mm rim thickness point corresponds to the maximum rim thickness possible since it corresponds to complete replacement. Note that the first time-point corresponds to 30 mins...... 96

Figure 4-7: Effect of exposed surface area on KCl replacement, while keeping the volume of crystal constant...... 97

Figure 4-8: Top and middle: Secondary electrons (SE) images of the cross section of a replaced KCl crystal before (top) and after (middle) it was smoothed using ultra- microtomy. The squares in the middle image represent areas from which the spectrum was averaged to obtain each of the plotted points. Bottom: Solid composition of the rim of a partially replaced KCl crystal obtained through quantitative EDS measurements. The dark area to the right is the unreplaced KCl crystal. Solid lines are guide to the eye...... 99

Figure 4-9: Electron images of blade-cut cross-sections of partly replaced KCl crystals at various solution to solid ratios and times...... 102

Figure 4-10: (Top) SE2 image of the cross section of a replaced KBr crystal smoothed using ultra-microtomy. Bottom: Solid composition of the rim of a partially replaced KBr crystal obtained through quantitative EDS measurements. The non-porous area to the very right is the unreplaced crystal. The white boxes show regions from which spectral data was averaged to get each point on the plot...... 103

Figure 4-11: A zoomed in section of the crystal shown in Figure 4-9 The crystal is oriented perpendicularly to that in Figure 4-9 so that the squares numbered 36-38 lie at the same distance from the replacement front. EDS data was acquired from the regions marked Map Data 36, 37 and 38...... 106

Figure 4-12: Circles represent bromide mole fractions in solid phases (considering only Br and Cl, excluding K) as a function of distance from the edge of crystal for replaced KCl and KBr crystals shown in Error! Reference source not found. and Figure 4-9 respectively, and marked as KCl, solid and KBr, solid respectively. Squares represent corresponding solution concentrations calculated using solubility phase diagram by assuming solid-solution local equilibria. All lines are guides to the eye...... 107

Figure 5-1: Schematic showing how an enhanced porosity from mineral replacement could increase gas yields. Higher the surface area exposed to the replacing fluid, greater the additional porosity generated through replacement...... 119

xiv Figure 5-2: Iceland spar calcite crystal (left), calcite crystals after hammering, sieving and washing (middle), sample S01 (see Table 5-3) calcite crystals after replacement for 6 days (right)...... 124

Figure 5-3: (Left) Multi position heater-stirrer used for the replacement experiments, (Middle) calcite crystals in the reaction vessel, placed on a polypropylene net have turned white after replacement, (Right) calcite crystals in a heat-sealed polypropylene net bag for replacement experiment with mixing. The crystals have turned white and joined together due to extra precipitation...... 126

Figure 5-4: Parr High P and T cell we used for calcite replacement. The vessel can fill upto 50 mL and can go upto 5000 psi and 350 0C. The reactor was pressurized with air...... 127

Figure 5-5: FESEM images of calcite crystals soaked in 2 M phosphate (buffer P1, pH 8.11) for 6 days at (a) 50 C and (b) 200 C ...... 131

Figure 5-6: Rim width vs. pH at 50 and 80 C. FESEM images of cross-sections of calcite crystals after replacement for 3 days in various buffers described in Table 1...... 133

Figure 5-7: Mean and maximum rim thickness observed for calcite crystals as a function of pH, in 3 days...... 134

Figure 5-8: Calcite solubility with pH using Visual Minteq. The simulation was done using a pH sweep at fixed temperature and total phosphate concentration, along with calcite specified as an infinite solid allowed to reach a saturation in solution...... 135

Figure 5-9: Speciation diagram of phosphate plotted by doing a pH sweep in Visual Minteq with 2 M phosphate at 50 C. The plot at 80 C (not shown here) shifts by less than half a pH point...... 137

Figure 5-10: EDS map of sample S03 (buffer P2, pH 5). (a) Electron image and (b), (c) elemental maps of phosphorous and carbon respectively overlaid on the electron image, for a cleaved calcite crystal from S03 removed at 6 days...... 139

Figure 5-11: Plot of rim thickness with time for S3 – calcite in pH 5 at 50 C, with and without mixing. The error bars are standard deviations calculated based on the variations in thickness from a single image...... 140

Figure 5-12: FESEM images of cleaved calcite crystals showing the effect of phosphate concentration and mixing on texture of the replaced phase. Comparing (a) and (c), we can see that at 2 M the replaced layer consists of small crystals as opposed to much larger crystals in (c) at 0.5 M. There is no significant change due to mixing at 2.2 M (part (b)) but at 0.5 M, mixing changes the texture and porosity of the replaced phase drastically...... 141

Figure 5-13: S90, Runaway reaction with mixing at pH 5. Left : Crystals in one of the 3 bags suspended in the reaction were transformed from separate crystals to one big white mass, taking shape of the bag. (b) The solution had a gelatinous precipitate of white which did not settle in 2 weeks...... 143

xv Figure 5-14: FESEM images of outside surfaces of replaced crystals. At 2.2 M (image a), the outer surface has roughness on a much smaller length-scale than at 0.5 M (image b) without mixing. In the presence of mixing, the outer surface at 2.2 M does not change much (image not shown) but at 0.5 M, the outer surface has much finer roughness due to mixing (image c). These images correspond to the rim cross-sections shown in Figure 5-12...... 144

Figure 5-15: Replacement along a fracture. FESEM image and EDS maps (carbon and phosphorous respectively) of the cross-section of a crystal from the experiment with buffer P2 with 1 mM Ca+2 added...... 145

Figure 5-16: (Left)Cumulative mercury intrusion pore volume as a function of pressure for various partly replaced crystals (S03, S07, S10) and untreated calcite crystals.(Right) Differential intrusion volume against equivalent pore diameter on a logarithmic scale. The area under the curve corresponds to total intrusion...... 146

Figure 5-17: Z-projections of maximum fluorescence from confocal microscopy. Crystals (a)-(c) were soaked in dye without pre-cutting, while (d) was cut before soaking. Crystals in (e) were saw-cut crystals rather than fractured crystals, replaced in buffer B15 (acetate) instead of P2 (HCl) used for (a)-(d) for 12 days...... 147

Figure A-1: Schematic illustration of the lattice model...... 161

Figure A-2: Average potential φ (scaled by e=4π퓵) as a function of position x3 (scaled by 퓵) as computed by the lattice model (black circles). Here, the simulation cell is L1 = L2 = 8퓵 by L3 = 32퓵 and contains M = 4 surface charges and N = 20 ion pairs; the inverse temperature is β = 1. The error bars represent 95% confidence intervals based on ten independent simulations. The open circles represent the average ζ- potential, which is assumed equal to the electric potential at x3 = 0.5퓵. The solid curve shows the solution to the nonlinear Poisson-Boltzmann equation for the same conditions...... 164

Figure B-1: Brownian rotational quenching and alignment near a planar surface (a) Left: Selection of frames from fluorescent microscopy videos (15 µm x 15 µm field of view) for fluorescent platinum-polystyrene (PS) Janus spheres of varying radii, the PS side of the colloid appears bright (a), near to a planar interface in de-ionized (DI) water, and in 10 % aqueous H2O2 solutions. Right: Polar angle, ϴ(t) for typical Janus particles determined from fluorescent microscopy videos (Note the a=2.4 µm particle in water shows strong gravitational alignment constraining ϴ close to 0° ). (b, c) Schematic 3D orientation and experimental trajectories (45 seconds duration, red line) for Pt-PS Janus particle with radius a = 1 µm in (b) DI water settled under gravity against a planar glass substrate and (c) 10 % H2O2 solution at either the top (+g) or bottom (-g) planar surface of a rectangular glass cuvette. (d) Polar Mean Square Angular Displacement (MSAD) as a function of time for three differently-sized Janus spheres. In each graph, the black “Water” line represents the MSAD for Janus particle settled at a planar interface under gravity in water (n>20). The additional curves represent the MSAD for Janus particles with speeds in the defined ranges, at both the top (n>20) and bottom (n>20) planar surfaces of a rectangular cuvette...... 174

xvi Figure B-2: Particles moving along geometric boundaries, at speeds of up to 10 µm/s. (a- b) schematics of Janus particles encountering multi-planar geometries. Red axis indicates forbidden rotations due to proximity to a plane, green axis indicates unquenched axis of rotation: (a) Janus particle encountering a planar edge while moving along a 2D surface, expected to result in Brownian rotational quenching about two orthogonal axes. (b) Janus particle confined within a square groove; parallel vertical walls confine the rotational diffusion about one axis; however, if the particle descends to the base of the groove, it is confined about two orthogonal axes. (c-f) Overlaid still frames from fluorescence microscopy videos with equal time gaps: yellow line shows complete trajectory, green line shows location of vertical cuvette walls, red arrows indication direction of motion: (c) a = 1.55 µm Janus particle (10 % H2O2) moving at the bottom of a rectangular glass cuvette a long way away from the edges. (d) a = 1.55 µm Janus colloid (10 % H2O2) moving along the curved edge of a glass cuvette – inset shows a colloid reaching the cuvette boundary. (e) a = 2.4 µm Janus colloid (10 % H2O2) moving along the straight edge of a glass cuvette- left hand inset shows a magnified region, right hand inset shows a “stuck” aligned agglomerate formed at the cuvette boundary. (f) a = 2.4 µm Janus colloid (10 % H2O2) moving within a rectangular section groove (width = 8.75 µm). (g) SEM image of a section of the rectangular grooves (widths 7-9 µm) used in (f)...... 177

Figure B-3: Electrophoretic behaviour for Janus colloids. (a) Schematic representation of the rotational electrophoresis experiment. Left hand side shows the relevant physical quantities, a Janus sphere with hemispheres with two different zeta potentials (Pt and PS), giving a dipole vector 풆. The dipole vector rotates in an electric field, and the right hand side 3D schematics depict the effect of switching the direction of E. 1. Represents the initial misaligned dipole and applied field orientation immediately after the E field direction is switched, stages 2-5 show two possible rotations to re- align the dipole with the applied field: on the left hand side about an out of plane axis, with constant polar angle, ϴ, and on the right about an axis parallel to the plane where polar angle changes; at position 5 풆 reaches the steady state. The black arrows show the direction of translational motion, which is always aligned with the applied field (see b). (b) Typical position vs. time curves obtained by tracking a Janus particle (a = 2.4 µm) above a glass interface with E = 2.5 V/cm in 1 mM NaCl. E pointed in the negative x direction first and then switched every second. The red circles are times when E changed directions. (c) Typical  vs. time curve for rotation of a Janus colloid (2.5 V/cm, 5 % H2O2, 1 mM NaCl). We changed the direction of E after the particle aligned with the applied field. (d) f() vs. t for (b), from equation 21 (see Methods). (e, f) show still frames from a fluorescence microscopy video for a Pt-PS Janus particle rotating about an out of plane axis (e) and about an axis parallel to the plane (f) from the point at which the applied E-field polarity was reversed to the depicted direction (red arrow). (g) Measured zeta potentials for both Pt (blue markers) and PS (red markers) at two time points following sample preparation, each with and without hydrogen peroxide...... 180

Figure B-4: Schematics of the colloidal swimmer near a substrate, showing the geometry (a) and the surface slip velocity flow field (b) indicates the direction of propulsion...... 183

Figure B-5: MSAD data re-scaled to allow comparison with theory, together with fits to equation 16 with estimated values for B (see equation 10)...... 187

xvii Figure B-6: Schematics of the colloidal swimmer near a substrate, showing the field lines created by the image distribution and 퐄퐢퐦퐚퐠퐞 at the location of the Janus sphere (a), and the real electric field lines (b)...... 190

Figure B-7: Calibration of cell for electrophoretic measurements. (a) Markers represent experimentally measured velocities of sPSL particles as a function of height from the centre. The solid line represents the best fit parabola using spsl and wall as parameters in Bowen’s equation. (b) Shows the spsl and wall calculated from the fit in fig a for 2 independent trials, C1 and C2. The last row shows the spsl measured using Malvern Zetasizer...... 204

Figure B-8: Relationship between particle size, fuel concentration and velocity. Average propulsion velocity versus fuel concentration. Supplementary Figure 1 shows the mean velocity as a function of hydrogen peroxide (H2O2) concentration for all of the Pt-PS (Platinum-Polystyrene) Janus particles investigated in section 1. “Top” refers to particles swimming at the top surface of the cuvette, and “Bottom” to particles sediment at the bottom of the cuvette...... 207

Figure C-1: Causes of fluid transport and interconversions of driving forces. These include well-known causes like pressure gradients, electrical or magnetic gradients, and gravitational gradients. However, chemically-driven flows, caused by chemical concentration gradients, are not widely-appreciated. The prevalence of chemical gradients, and thus chemical potential gradients, makes diffusiophoretic transport more widespread and pervasive than has been realized...... 216

Figure C-2: Essential mechanism of electrolyte diffusiophoresis. The mechanism consists of two parallel additive phenomena: electrophoresis caused by an in-situ electric field (E) generated by a concentration gradient (or more precisely, a chemical potential μ gradient) of NaCl, and chemiphoresis caused by a gradient of NaCl concentration, and therefore a gradient of pressure in the EDL, tangential to the particle surface. This pressure gradient drives fluid flow along the particle surface inside the EDL. For a β negative salt like NaCl, both mechanisms transport the negatively-charged particle towards higher ionic concentration...... 220

Figure C-3: Diffusiophoretic velocity increases and dominates over convective velocity with reduction in porosity. 20 mM NaCl is rejected by an ideal semipermeable membrane with water flux of 23 LMH. Porosity change is time dependent, and also depends on the type of salt (훃) and particle zeta potential 퐩. With higher values of 훃 and 퐩 the diffusiophoretic velocity increases further...... 228

Figure C-4: Generating gradients by diffusive mixing in the laboratory. (a) 2 cm long and 0.15 m thick glass capillaries were filled with different concentrations of a solute and placed in a petri dish that contained no solute91. At the mouth of the capillary, two solutions of different concentrations meet, creating a concentration gradient. (b) A capillary was filled with salt solution and one end was sealed off. The other end was dipped into a bigger capillary that served as a reservoir containing only DI water14. Again, two salt concentrations were placed adjacent, giving a gradient...... 234

xviii Figure C-5: Boundary Layer Diffusiophoresis (BLDP). (a) A transverse salt gradient is created due to the concentration boundary layer as the DI water from outer reservoir flows diffusioosmotically on the inner capillary that contains 10 mM NaCl or KCl. This causes tracer particles to deflect towards the walls due to BLDP as they enter the capillary102. (b) & (c) Tracks of tracer particles as they enter from the outer capillary containing DI water into the inner capillary containing 10 mM NaCl in (b) and 10 mM KCl in (c). The deflection towards the wall is attributed to BLDP. The deflection is small in this case since KCl has a much lower  value14...... 237

Figure C-6: Schematic of hypothesized viral transport from MTOC compartment to nuclear pore in eukaryotic cell. One of the transport mechanisms could be diffusiophoresis driven by pH gradients (훁퐜/퐜) near the vicinity of nucleus from cytosol...... 240

Figure C-7: Left17: Diffusioosmotic pumping of 1.4 m sulfated polystyrene latex tracer particles by two interacting 7 m CaCO3 particles settled close to a glass slide. Right17: A clear exclusion region of tracers develops around a barium sulfate micropump on a glass surface. Despite the very low solubility of BaSO4, the diffusiophoretic pumping still occurs. These tracers, after being ejected, exhibit mostly Brownian motion. The scale bars are 10 m...... 242

Figure C-8: A schematic of diffusiophoresis in crystallization: A small seed crystal of size r0 is spontaneously formed, with density,, radius r0 and >0 (say). Its formation depletes salt of -value, over a length scale R so that the salt concentration goes from saturation, Cs to that of the supersaturated bulk, Cb. The d\Debye length in bulk is - 1. The salt gradient gives rise to a spontaneous electric field and the particle moves towards higher salt concentration as indicated by the direction of udp...... 246

xix LIST OF TABLES

Table 3-1: Zeta potentials of sPSL particles averaged over 1 M to 5.5 M for each salt. The uncertainties are calculated as root mean square of the 90% confidence intervals for the measurement for each salt concentration, for each salt...... 60

Table 5-1: Summary of previous and current work done on dissolution- precipitation of calcium carbonate in phosphate solutions. A summary of other work done at even higher temperatures can be found in Yoshimura et al11...... 122

Table 5-2: Buffer compositions used for replacement experiments. * mono refers to ammonium phosphate monobasic as the source of phosphate...... 124

Table 5-3: Conditions for experiments reported. All experiments were carried out with fractured crystals...... 125

Table 5-4: Conditions for additional mixing experiments...... 142

Table A-1: Summary of zeta potential measurements ...... 167

Table B-1: Measured values of zeta potential of the Janus particle, platinum and polystyrene halves of the Janus particle in water and 5% H2O2 at time 10 mins and 60 mins. This table shows the actual values of the zeta potential measured for each particle and the averages for each trial, referenced in Figure 3g of the manuscript. The averages indicate very good consistency within a batch and the intra-batch variability is similar for trials with and without H2O2. Thus it is clear that the method of subtracting velocities of tracer particles in the presence of bubbles is effective in removing the effect of complex convective flows around a bubble...... 182

Table B-2: Diffusion coefficients parallel to a wall (D||) and the corresponding gap heights (h) of a of Janus colloids in both water and KNO3 solution at 21oC. Values in brackets are the standard deviation of the mean values...... 200

Table C-1: Categories of concentration gradient origins. Some of the systems “known” to involve diffusiophoresis (DP) are recent discoveries (e.g., diffusive mixing). Also, for some of the categories, the authors are not aware of any known examples in the literature...... 216

Table C-2: Ionic diffusivities at 25˚C and temperature independent β values of common electrolyte species38. Gradients of some salts (e.g., KCl) will result primarily in chemiphoresis. For the reaction CaCO3 + H2O = Ca+2 + OH- + HCO3-, apparent β of CaCO3 is based on equation (3)...... 221

Table C-3: Chemical reactions given in the literature that produce diffusiophoretic transport...... 232

Table C-4: Summary of the effects of externally induced diffusiophoresis...... 240

xx ACKNOWLEDGEMENTS

This Dissertation resulted from an almost 5 years long journey through successes and failures, thrills and disappointments, fear and excitement of the unknown. And for this journey, I want to thank my fellow travelers, mentors, guides, friends and rest-houses on the way.

I want to thank my advisor and the greatest teacher ever - Darrell Velegol who guided me through this journey, just as Krishna guided Arjuna in the Geeta. Time spent learning from him not just about colloids, but also about life and about generally staying afloat in a world that’s always running have contributed a great deal in making me a ‘doctor of philosophy’ in the real sense of the word.

Next I want to thank my mom and dad for standing by me in this hopeless labor of love. I want to thank them for believing that if I pursued this love for research, it will certainly take me somewhere, even if they – or I – didn’t know where this ‘somewhere’ was when the journey began. I want to thank them for raising me up to dream big things and to ignore the crowd, and I only follow their examples in what I do.

Coming to my partner in crime and in life, my husband Siddhartha, I am not sure if I should thank him for the dissertation or the dissertation for him. We met on our first journey to Penn State and he has been a witness and the greatest fellow traveler ever on this journey right from day 0. I want to thank him for his beautiful company along the way, his very useful advise that served as a balancing act against my own nature, and for our favorite leisure together – food.

I would also like to thank my in-laws for standing by me in the ups and downs of graduate school, offering me wisdom, strength and patience in times when I absolutely needed it. I want to thank my sister and brother-in-law, who were my go to place for every vacation. Apart from offering a

xxi vacation home, they have brought to life my nephew Saharsh – a 1 year old little bundle of joy. His each little triumph on the journey of ‘growing up’ is an inspiration to me to keep fighting.

I want to also thank my lab-mate Rajarshi Guha for his help, ideas and advice, Sruti who was my first collaborator at Penn State and who taught me the first things around the lab, Abhishek for his inspiring ideas some of which led me to the questions in my thesis and Farzad for long-winded discussions about science, behavior, culture, sexism and food. I want to thank Charles Cartier and

Sambeeta Das who I much enjoyed working with and learning from. I am thankful to undergraduate student Ibrahim who lent a helping hand in my pursuits in the lab, and Alex for his tenacity and good questions.

I am grateful to the many professors who taught me, collaborated with me or mentored me at Penn

State through research, courses and offering valuable feedback being on my thesis committee. I very much enjoyed and learnt immensely from collaborating with Prof. Kyle Bishop, Prof.

Ayusman Sen, Prof. Manish Kumar and Prof. Christopher Gorski at Penn State. Courses taught by

Prof. Kyle Bishop, Prof. Ali Borhan, Prof. Scott Milner, Prof. Kristen Fichthorn, Prof. Christine

Keating, Prof. Kwadwo Osseo Asare and Prof. Jim Adair all shaped my knowledge and thought process that have helped me write this dissertation. Money isn’t everything but certainly, my research would not have been possible without it. I am grateful to Dow Chemicals, NSF MRSEC,

NSF CBET and Halliburton that provided funding for my research.

This work would also not have been possible without my friends Fish, Peng, Animesh, Soumalya,

Sreemoyee, Sravani Di, Sambeeta and Wenjie who always seemed to appear each time I needed company and always willing to share food that I very much enjoyed cooking for them. Last but not the least, I must acknowledge the contribution of tea, coffee and the various cafes in State College where I wrote my thesis - Café Lemont, Webster’s and Barranquero café, Lila Yoga Studio, my

xxii go-to place for keeping my sanity and audible for the greatest source of knowledge and endless entertainment to recharge me for more writing.

1

Chapter 1

Motivation and Research Goals

1.1 Motivation

The objective of this thesis is to answer questions about the behavior of saturated salt systems from the perspective of surface charge, fluid flows and dissolution-precipitation reactions. Saturated salt solutions exposed to colloidal particles or surfaces are encountered in a broad range of scientific, industrial and medical problems – in the natural processes that transform rocks1 and degrade building materials2, in methods of enhanced oil recovery3,4 and carbon dioxide capture5, in the synthesis of simple6 or designer inorganic particles7 and even in finding a cure for kidney stones8.

In each of these systems, colloidal particles or surfaces interact with each other and with the saturated salt solution. However, our understanding of these systems is often empirical, rather than fundamental. Creating simpler models of these complex systems based on an understanding of the fundamentals is essential to be able to scientifically control, improve, or tailor them to specific applications. This thesis addresses the gap in our knowledge from primarily two perspectives that each involve fixed or mobile surfaces in saturated salt solutions – (i) zeta potential and electrokinetics, and (ii) mineral replacement caused by dissolution – precipitation reactions.

Aqueous saturated salt solutions – solutions saturated with respect to a particular salt crystal – abound in nature, and by definition, contain a surface in chemical equilibrium with the solution.

Consider for example, a simple picture of a geological reservoir that contains limestone walls and sodium chloride deposits in equilibrium with the surrounding water – a salt solution saturated with

2 respect to CaCO3 and NaCl (Figure 1-1). The reservoir also contains droplets of oil and other colloidal particles such as clay particles. The interactions of each of these interfaces - the walls, the oil droplets, and the clay particles - with each other and with the solution lead to changes in the system over time. A key problem in our systems of interest is predicting and controlling such changes in the system namely, material transport, adhesion, chemical composition and material properties. These changes ultimately depend upon the physical and chemical interactions such as those arising due to surface charge or chemical reactions such as dissolution and precipitation. The phenomena of surface charge, and dissolution-precipitation form the subject of this dissertation, as discussed below.

Figure 1-1: Zeta potential, dissolution and precipitation in a geological reservoir. A simplified view of a geological reservoir that has charged colloids suspended in a saturated salt solution. The calcite surface and NaCl crystals participate in dissolution – precipitation equilibria.

A measure of charge at the interface is provided by the zeta potential, which is a key property involved in adhesion and electric field driven transport of fluid and particles. A starting point for my work on zeta potential was an important discovery that increased oil yields - the discovery that pumping low salinity fresh-water instead of high salinity sea-water gives rise to enhanced oil

3 recovery from geological reservoirs9. The mechanism of this behavior is still a subject of debate, but two possible explanations hinge on the existence of a zeta potential on exposed surfaces in the system. First, Tang and Morrow showed that the salinity dependent yield was contingent upon the presence of colloidal particles10. They further proposed that repulsive or attractive electrostatic interactions could have a role to play since these are salt concentration dependent colloidal interactions10. Secondly, Kar et al11 showed that salt concentration gradients could drive fluid flow and the transport of oil droplets suspended in water provided that surfaces and particles had a finite zeta potential. This was based on the phenomenon of electrolyte diffusiophoresis12, the transport of fluid at a charged surface or of charged particles in the presence of a salt concentration gradient

(Figure 1-2). Thus, two possible explanations of enhanced oil recovery with low salinity brines each relied on the presence of a finite charge on colloidal particles or oil droplets under reservoir conditions – typically high salt and saturated.

Figure 1-2: Charged surfaces in salt concentration gradients can cause flows in dead-end pores through diffusiophoresis. However, the presence of a finite zeta potential under high and saturated salt conditions (for eg. 5.3

M NaCl) is neither known nor easy to test. The primary difficulty is that at high salt concentrations the existing theory is not applicable and existing experimental measurement methods do not work.

The first key contribution of this dissertation is an electrophoretic technique to measure zeta

4 potential of model polystyrene latex particles at high salt conditions (5.3 M NaCl). Our measurements reveal a finite zeta potential up to saturation which continue to follow the theory developed for low salt concentration for some particles even after major assumptions of the classical theory break down.

The second set of contributions in this dissertation focus on pseudomorphic mineral replacement reactions (pMRRs). pMRRs, also called coupled-dissolution-reprecipitation (CDR) reactions are reactions where the space occupied by a parent mineral comes to be occupied by a guest mineral, such that the external morphology of the parent mineral is preserved and porosity is created in the guest phase13–15. These fluid-mediated reactions cause changes in the chemical composition as well as material properties such as porosity and hardness. We hypothesized that these reactions when carried out controllably in a geological reservoir could enhance oil recoveries by creating additional porosity when used in conjunction with conventional mechanical fracking used by oil companies

(Figure 1-3). In modern fracking operations commercialized in 1949 in a patent16 licensed to

Halliburton, fluid is pumped at high pressure to mechanically rupture the rocks, thereby creating fluid pathways for the oil to flow out. We aimed to find the solution composition that would cause pMRR of a reactive reservoir rock such as calcite to create additional porosity in the fracking process, creating chemical-mechanical-fracking.

5

Figure 1-3: Our hypothesis of chemical-mechanical fracking. pMRR at the fractured surfaces would create additional porosity, opening up more volume for fluid flow. Carrying out and controlling this reaction under reservoir conditions requires sound knowledge of the mechanism. However, the mechanism of these reactions is poorly understood, especially the inter-relationships and relative contributions of thermodynamics, kinetics and fluid transport. In the second part of my thesis, I have tried to build a better theoretical understanding of the replacement process by studying the model replacement system KBr-KCl. An overarching question in this investigation was coming up with general principles to better predict and control mineral replacement reactions, such that these principles are applicable to both high and low solubility systems. I have questioned some of the assumptions that have prevented a better understanding of this system so far. I have contributed to understanding the mechanism of replacement and reconciling many existing results by performing well-controlled experiments to measure the rates of replacement and composition of fluid under a variety of conditions.

Building on a better understanding of the replacement process, I studied the far more complicated and much less understood replacement system – calcite in phosphate solutions, for application to increasing porosity in calcitic geo-reservoirs. In this work, I mapped the replacement process under

6 conditions that were relevant to our intended application but have never been traced before.

Through my systematic investigations, I have contributed not only the conditions necessary for chemically generating micro-porosity in calcite, but also a deeper understanding of the replacement process.

1.2 Research Goals

In this dissertation, I sought to answer questions about the behavior of saturated salt systems from the perspective of surface charge, fluid flows and dissolution-precipitation reactions. Below is a brief look at the three problems I worked on. The description below provides the specific questions investigated, the approach and the key results obtained.

1. Zeta potential at high and saturated salt conditions

What is the zeta potential on model polystyrene latex particles under high and

saturated salt conditions (eg. 5 M NaCl)? How does the zeta potential depend upon

the surface functionalization or salt type?

2. Mineral replacement in the KBr-KCl system

(a) How does the rate of replacement of KCl (or KBr) crystals in saturated KBr (or

KCl) solutions vary with solution to solid ratios and fluid mixing?

(b) What is the composition of the replaced layer for each of the above systems as a

function of position in the crystal? How do these composition differences relate to

the thermodynamics and fluid transport rate?

3. Chemical micro-fracking of calcite through mineral replacement

7 (a) What is the dependence of rate of calcite replacement in 2 M phosphate solutions

on the pH in the range 4-8, at 50 and 80 C? Does the rate of replacement correlate

with the pH dependent rate of calcite dissolution?

(b) What is the increase in porosity generated by pMRR in comparison to bare crystals,

and how deep does it penetrate into the crystal?

1.2.1 Zeta potential at high salt and saturated salt conditions

What is the zeta potential on model polystyrene latex particles under high and saturated salt conditions (eg. 5 M NaCl)? How does the zeta potential depend upon the surface functionalization or salt type?

Most interfaces attain a surface charge in water such as solid-liquid (particles), gas-liquid (bubbles) or liquid-liquid ( droplet) interfaces17. The charge may be a result of ionization of surface groups or due to adsorption of ions from solution. This surface charge is characterized by an electrical potential which is equal to the zeta potential () at the slipping plane. The surface charge attracts counter-ions which form the electrical double layer (EDL) (Figure 1-4). The classical theory of electrokinetics – the Gouy-chapman theory – predicts that the zeta potential is directly proportional to the surface charge density () and is inversely proportional to the square root of salt concentration17. Under the usual assumption of a constant surface charge, the theory predicts that the zeta potential becomes very small at high salt concentrations. However this theory is not expected to work at salt concentrations greater than about 140 mM where its assumptions of no ion-ion correlations and point-sized ions breaks down18.

8

Figure 1-4: Zeta potential and electrical double layer for a negatively charged particle in water

For particles in salty media, one might therefore anticipate that the structure of the EDL and the corresponding zeta potentials will differ significantly and qualitatively from expectations built largely on the study of dilute electrolytes. It is critical to identify and understand these differences in order to predict the behaviors of colloids in salty environments such as human blood

(≈150 mM), seawater (≈600 mM), wastewater (≈2 M during reverse osmosis), and geological reservoirs (≈5 M). The standard methods used to determine zeta potentials of particles at low ionic strengths are often inapplicable at high salt concentrations, and therefore, few studies reliably report zeta potentials for colloidal particles at high . We are especially interested in zeta potential at saturated salt conditions, where the assumptions of classical theory break down most dramatically.

This led me to my questions – What is the zeta potential on model polystyrene latex particles under high and saturated salt conditions (eg. 5 M NaCl)? How does the zeta potential depend upon the

9 surface functionalization or salt type? In this work I have investigated these questions for model polystyrene latex particles which are commercially available, have been well characterized and studied extensively by other researchers at low salt concentrations. However, the measurement technique is applicable to any colloids large enough to be seen by optical microscopy and with density close to that of the medium such as oil droplets and bacterial cells suspended in water.

In this dissertation, I describe an electrophoresis based technique using sinusoidal electric fields and high speed microscopy that we developed to carry out these measurements. Electrophoresis is the motion of charged particles when subjected to an electric field with a velocity proportional to the applied electric field and to the particle’s zeta potential12 (Figure 1-5). The displacement of the particles at high salt was small and noisy, but the use of Fourier transforms enabled the resolution of the velocity at the applied frequency. This was used to calculate the zeta potential.

Figure 1-5: Electrophoresis of a particle with positive zeta potential. Particle velocity, u is directly proportional to the electric fields and zeta potential on the particle’s surface. My measurements show that not only is the zeta potential finite and measurable up to saturation, it is also in accordance with the classical Gouy- Chapman theory for sulfated polystyrene latex particles even after critical assumptions of the theory are violated. While the zeta potential is too small to stabilize these particles electrostatically against aggregation, it is high enough to cause a

10 measurable electrokinetic transport under the influence of applied electric fields (3 – 10 V/cm) or spontaneous electric fields created by salt concentration gradients (diffusiophoresis). Additionally, for amidine functionalized polystyrene latex particles, I find a reversal in zeta potential between

0.5 – 1 M salt, most prominently for KBr among the salts we tested. This is the first observation of charge reversal for particles in simple monovalent salts. Lastly, I also present Monte-Carlo simulations of a simple lattice model of the electrical double layer that support the experimental results.

The development of a simple, reliable technique to carry out zeta potential measurements at high salt is a major advance in this field. This is especially so since commercial instruments often fail to give reliable measurements at high salt, and are effectively black-boxes that contribute further to a lack of reliability. While we focused on saturated salt concentrations here, I anticipate that this technique would be of much use in biological sciences where bodily fluids – urine and blood have high salt concentrations, at the borderline of the range of applicability commercial instruments.

Scientifically, the finding of a finite charge at high salt that follows the classical theory is an important first step in developing the electrokinetic theory for saline aqueous solutions and for the emerging field of ionic liquids.

1.2.2 Mineral replacement in the KBr-KCl system pMRRs, also called coupled-dissolution-reprecipitation (CDR) reactions are fluid mediated mineral replacement reactions where the space occupied by a parent mineral comes to be occupied by a guest mineral such that the external morphology of the parent mineral is preserved and porosity is created in the guest phase13–15. These reactions proceed by dissolution of the parent phase into the fluid that subsequently gets saturated with respect to the guest phase and results in precipitation

11 (Figure 1-6). For pMRR to proceed, the replaced phase must contain porosity so that internal surface can get exposed to the fluid that mediates the reaction.

Figure 1-6: Schematic of mineral replacement of parent AB by a porous layer of the guest AC, mediated by a solution of AC. For replacement to proceed, the interfacial fluid contacting the phase AB should be saturated. pMRRs can be observed in high solubility systems such as KBr crystal in saturated KCl solution, and low solubility minerals such as the replacement of calcium carbonate to hydroxyapatite when exposed to phosphate solutions. Recently, these reactions have found application in new material synthesis and in this thesis I have tried to use it to enhance the porosity in geological reservoirs.

The reactions are also widespread in nature, yet poorly understood owing to a complex inter- dependence between the three primary factors – thermodynamics, kinetics and transport. In order to use it for industrial applications, it is desirable to understand the mechanism of these reactions clearly. With the objective of getting closer to creating a mathematical model of the replacement process, I studied pMRR in the KBr-KCl system. This high salt system has been used previously by researchers as a model system where replacement proceeds within the time-frame of a few hours,

12 and the thermodynamics of the system are well-known19–21. An interesting property of this system is that replacement of KBr in saturated KCl proceeds much faster than that of KCl in saturated KBr.

My questions below try to find the causes and consequences for this difference.

The complete replacement of a 3 mm KBr crystal in saturated KCl solution happens within 2 hours, resulting in a porous KCl crystal. But the reverse reaction – KCl crystal in saturated KBr solution was reported to stop within a few minutes (Figure 1-7). The difference between the two cases has been attributed to the fact that KBr has a larger lattice size and higher solubility than KCl. This, according to previous researchers are the two factors that control porosity in the replaced phase, and a higher porosity leads to a higher replacement rate as more surface area is exposed. However, we hypothesized that solid to solution ratio, as well as fluid transport must also play a role in the replacement process, and could potentially change the outcome, speed and extent of replacement.

Figure 1-7: Camera and electron microscopy images of replacement in the KBr-KCl system. KBr crystal in saturated KCl solution shows replacement of ~ 2 mm over 24 hours, while the KCl crystal shows replacement of ~ 0.2 mm over the same time. This was the starting point for my first question - How does the rate of replacement of KCl

(or KBr) crystals in saturated KBr (or KCl) solutions vary with solution to solid ratios and fluid mixing? I hypothesized that I might be able to control the reaction through solution to solid ratios and transport rate. The solution to solid ratios are known to affect the end point of the reaction for

13 KBr in KCl system, but its effect on the rate of pMRR has not been studied. Additionally, the effect of transport has been completely left out of previous studies even though the dissolution of highly soluble salts is generally transport controlled. In my experiments, I found that for solution to solid molar ratios in the range 1-50, KCl replacement in KBr depends on the ratio such that if the ratio is 1, replacement stops in 2 hours but when the ratio is 50, an entire 3 mm crystal can be replaced in 18 hours. On the contrary, KBr crystals are completely replaced for all the ratios tested within 1 hour.

Next, I asked the question - What is the composition of the replaced layer for each of the above systems as a function of position in the crystal? How do these composition differences relate to the thermodynamics and fluid transport rate? To answer this question, I carried out replacement experiments of KCl and KBr, and then measured the composition over an internal cross-section of the replaced crystal using quantitative Electron Dispersive Spectroscopy (EDS or EDX). These measurements showed the presence of gradients in Br:Cl ratio in the rim, as a function of position within the crystal. The solid composition could be related directly to solution composition within the crystal through a mixed solid-solution equilibrium relationship. Through this, I concluded that the solution found in the pore space of the replacement rim in a KCl crystal has a far steeper concentration gradient than the solution in the rim of KBr crystals. This behavior correlates directly with the reduced porosity, slower internal transport rate and subsequently a lower rate of replacement of KCl than KBr.

My study of these reactions helps show that the KBr-KCl reaction can proceed both ways, and that internal transport is the rate limiting factor in KCl replacement. Additionally, it provides many insights into the mechanism that are essential in the formulation of a mathematical model for this system that can eventually help us design replacement reactions for various industrial applications.

14 1.2.3 Chemical micro-fracking of calcite through mineral replacement pMRR of calcite could be used as a way to generate porosity in calcite-rich oil and gas reservoirs, which could result in significant economic benefit. The literature reports replacement of calcite in ammonium hydroxy phosphate (NH4)2HPO4 solutions and in sodium fluoride NaF. Of these, I have considered the use of phosphate rather than fluoride since pumping fluoride in large quantities under the earth might be hazardous for ground water resources. For application to geological reservoirs, we are interested in T between 50-80C. However, previous studies of calcite replacement in 2 M phosphate report replacement at high T (120C ≤ T ≤ 200C) and no replacement for T ≤ 80C. Previous studies have also hypothesized that dissolution is the rate limiting step for this replacement. Since the kinetics slow down at low temperatures, we hypothesized that lowering the pH might allow us to increase the rate of replacement at lower temperatures.

This led me to my first question – What is the dependence of rate of calcite replacement in 2 M phosphate solutions on the pH in the range 4-8, at 50 and 80 C? Does the rate of replacement correlate with the pH dependent rate of dissolution? To answer this question, I performed replacement experiments of calcite in buffers at various pHs in the range 4-8, containing 2 M phosphate and measured the replacement rim width at 3 days through electron microscopy images.

At 50 C, I found that the replacement width was roughly monotonic with pH and followed the expected dissolution rate – the highest replacement was at the lowest pH. But at 80 C, the replacement width was non-monotonic, unlike the dissolution rate. The highest replacement at 80

C was obtained at pH 7, clearly suggesting that the dissolution rate alone isn’t controlling the rate of replacement, and that the more complex precipitation rate may need to be accounted for.

15 Having obtained some replacement, we are interested in getting a high porosity on a controllable and observable time scale of days, preferably with low concentrations of added salts. This brought me to my second question - What is the increase in porosity generated by pMRR in comparison to bare crystals, and how deep does it penetrate into the crystal? For porosity measurements, I used mercury intrusion porosimetry of replaced crystals and found a 10-fold increase in the porosity of calcite replaced for 6 days at 50 or 80 C compared to untreated calcite. To measure the depth of penetration, I soaked replaced crystals in a fluorescent dye for a day and then used confocal microscopy to see how far into the crystal the dye had penetrated. Through these experiments, I found that the dye can penetrate into the crystal at a mm scale even though the rim width is <100

m. This deep penetration is only found in crystals with internal fractures, showing that it is due either to ‘opening up’ access to pre-existing fractures, or due to replacement proceeding much faster at internal fractures.

These reactions have previously been studied from a geological and synthesis perspective for several decades but I have for the first time investigated pMRR from the perspective of enhanced oil recovery, mapping the replacement behavior in regimes that were never tested before – a range of pH, medium temperatures and fairly long times (up to 12 days). While this work focused on calcite, further work might permit the use of pMRR to generate porosity in sandstone or clays.

Previous studies of calcite replacement focused on high temperatures ( T >= 120 C) and failed to find any replacement at lower temperatures. A key contribution of this work is the finding that replacement can be achieved at these low temperatures by setting the right pH of the phosphate solution. Another key contribution is quantifying the total porosity as well as it’s quality – the depth of pores using dye penetration confocal microscopy.

16 1.3 Organization of the thesis

Chapter 2 sets up a theoretical background on the definition, significance and measurement techniques for zeta potential. Chapters 3,4 and 5 form the main body of new findings during the course of my dissertation, and each of these is structured as a stand-alone project. Chapter 3 discusses the zeta potential work, chapter 4 discusses the work on pMRR in the KBr-KCl system, and chapter 5 describes the work on pMRR of calcite. Chapter 6 concludes this work by discussing the scientific significant of this work, applications and possible paths for future work.

Appendices 1-3 continue the discussion on electrokinetics and colloids after Chapters 2 and 3.

Appendix 1 presents details of the theoretical model for Monte Carlo simulations of the electrical double layer at high salt concentrations, done in collaboration with prof. Kyle Bishop. Appendix 2 is derived from a manuscript I co-authored on the steering of active Janus particles close to boundaries. Appendix 3 is derived directly from a review paper on diffusiophoresis written in collaboration with Prof. Darrell Velegol, Rajarshi Guha, Dr. Abhishek Kar and Prof. Manish

Kumar. As a whole, the present work presents techniques to probe the physics and visualize the fluid flows in saturated salt solutions.

All the work presented here has major individual contribution from me, but was done collaboratively. Each of the chapters 3-5 and the appendices A-C has a section each on contributions and acknowledgements to credit these people who were collaborators, mentors or staff without whose assistance this dissertation would not have been possible. The work in Chapter

3, and appendices A-C consists of work already published in the academic literature. The credits for this work have been given appropriately in a footnote one the first page of each of those chapters.

17 1.4 References

(1) Putnis, A. Mineral Replacement Reactions. Rev. Mineral. Geochemistry 2009, 70, 87–124.

(2) Flatt, R. J. Salt Damage in Porous Materials: How High Supersaturations Are Generated. J.

Cryst. Growth 2002, 242, 435–454.

(3) Austad, T.; Shariatpanahi, S. F.; Strand, S.; Black, C. J. J.; Webb, K. J. Conditions for a

Low-Salinity Enhanced Oil Recovery (EOR) Effect in Carbonate Oil Reservoirs. Energy

and Fuels 2012, 26, 569–575.

(4) Alvarado, V.; Manrique, E. Enhanced Oil Recovery: An Update Review. Energies, 2010, 3,

1529–1575.

(5) Pruess, K.; Müller, N. Formation Dry-out from CO2 Injection into Saline Aquifers: 1.

Effects of Solids Precipitation and Their Mitigation. Water Resour. Res. 2009, 45.

(6) Gupta, A. K.; Gupta, M. Synthesis and Surface Engineering of Iron Oxide Nanoparticles

for Biomedical Applications. Biomaterials 2005, 26, 3995–4021.

(7) Guo, Y.; Zhou, Y.; Jia, D.; Tang, H. Fabrication and Characterization of Hydroxycarbonate

Apatite with Mesoporous Structure. Microporous Mesoporous Mater. 2009, 118, 480–488.

(8) Adair, J. .; Aylmore, L. a. .; Brockis, J. .; Bowyer, R. . An Electrophoretic Mobility Study

of Uric Acid with Special Reference to Kidney Stone Formation. J. Colloid Interface Sci.

1988, 124, 1–13.

(9) Morrow, N.; Buckley, J. Improved Oil Recovery by Low-Salinity Waterflooding. J. Pet.

Technol. 2011, 63, 106–112.

(10) Tang, G. Q.; Morrow, N. R. Influence of Brine Composition and Fines Migration on Crude

Oil/brine/rock Interactions and Oil Recovery. J. Pet. Sci. Eng. 1999, 24, 99–111.

18 (11) Kar, A.; Chiang, T.-Y.; Ortiz Rivera, I.; Sen, A.; Velegol, D. Enhanced Transport into and

out of Dead-End Pores. ACS Nano 2015, 9, 746–753.

(12) Anderson, J. Colloid Transport By Interfacial Forces. Annu. Rev. Fluid Mech. 1989, 21, 61–

99.

(13) Putnis, A. Mineral Replacement Reactions: From Macroscopic Observations to

Microscopic Mechanisms. Mineral. Mag. 2002, 66, 689–708.

(14) Altree-Williams, A.; Pring, A.; Ngothai, Y.; Brugger, J. Textural and Compositional

Complexities Resulting from Coupled Dissolution–reprecipitation Reactions in

Geomaterials. Earth-Science Rev. 2015, 150, 628–651.

(15) Pedrosa, E. T.; Putnis, C. V.; Renard, F.; Burgos-Cara, A.; Laurich, B.; Putnis, A. Porosity

Generated during the Fluid-Mediated Replacement of Calcite by Fluorite. CrystEngComm

2016, 18, 6867–6874.

(16) Reistle Jr Carl E. Method of Treating Earth Formations. US 2547778 A, 1949.

(17) Hunter, R. J. Zeta Potential in Colloid Science: Principles and Applications; Academic

Press, 1981.

(18) Storey, B. D.; Bazant, M. Z. Effects of Electrostatic Correlations on Electrokinetic

Phenomena. 2012, 56303, 1–11.

(19) Putnis, C. V.; Mezger, K. A Mechanism of Mineral Replacement: Isotope Tracing in the

Model System KCl-KBr-H2O. Geochim. Cosmochim. Acta 2004, 68, 2839–2848.

(20) Pollok, K.; Putnis, C. V.; Putnis, a. Mineral Replacement Reactions in Solid Solution-

Aqueous Solution Systems: Volume Changes, Reactions Paths and End-Points Using the

Example of Model Salt Systems. Am. J. Sci. 2011, 311, 211–236.

19 (21) Kar, A.; Mceldrew, M.; Stout, R. F.; Mays, B. E.; Khair, A.; Velegol, D.; Gorski, C. A. Self-

Generated Electrokinetic Fluid Flows during Pseudomorphic Mineral Replacement

Reactions. Langmuir 2016.

20

Chapter 2

Theoretical background on Electrokinetics

In this chapter, I introduce the major concepts that form the basis of the work on zeta potential and electrokinetic transport in this dissertation. The zeta potential () is a key parameter that is a measure of the surface charge on a particle or surface. In what follows below, I discuss the definition, applications, classical theories and measurement techniques for  potential. Next I provide a background on colloidal forces which will enable us to appreciate the challenges in colloidal formulation, especially at high salt concentration.

Lastly, I discuss two – electrophoresis and diffusiophoresis which depend on the

 potential. This is intended to set a theoretical background that will help understand the motivation, questions, techniques and applications of my work on zeta potential.

2.1 Definition of zeta potential and the electrical double layer

Most particles attain a surface charge in water which may be due to several reasons such as ionization of surface groups as in the case of sulfate groups or silica groups, or due to adsorption of ions on the surface.

This surface charge is characterized by a potential called the surface potential, 휓0 on the particle surface.

The surface charge attracts counter-ions from the solution which form the electrical double layer (EDL). In a fairly simple picture of the double layer, it is thought to consist of a fixed layer of counter-ions held firmly close to the surface, called the Stern layer, and a diffuse layer where the counter-ions are more loosely attached, called the Gouy Chapman (GC) layer. While the Stern layer consists mostly of counter-ions, the diffuse layer consists of both counter-ions and co-ions, with a deficit of co-ions. Over the diffuse layer, the potential drops to zero as we go from the Stern layer to the neutral bulk of the solution.

21

Since the Stern layer is fixed on the surface of the particle, it forms a slip plane1. The potential at the slip plane is defined as the zeta potential,  which can usually be estimated using measurements of electrokinetic parameters. Since under most situations 휓0 is not directly measurable, it is approximated to be the zeta () potential.

Figure 2-1: Zeta potential and electrical double layer for a negatively charged particle in water

2.1.1 Examples and estimates

 potential of most particles and surfaces in water is negative, typically in the range -15 mV to -150 mV at

1 mM NaCl. Examples of surfaces with negative zeta potential at close to neutral pH include oil droplets2, dust particles3, sulfated4 or caboxylated5 polystyrene latex particles, glass or silica surfaces6 and platinum particles7. Some of these surfaces may become positively charged upon lowering the pH, such as silica or carboxyl surfaces. Positively charged surfaces are less common in nature, but adsorption of positively charged species such as multivalent cations8, eg. Ca+2, proteins9 and surfactants10 can make surfaces positively charged. Synthetic positively charged particles such as amidine-functionalized polystyrene

22 latex11 particles are made positive by functionalization with amidine groups or other positively charged groups. Within the approximation of a thin Debye layer (a >> 1), the  potential reduces in magnitude with salt concentration.

2.2 Applications of zeta potential

The sign of the  potential – positive or negative – can be easily obtained based on experimental measurements, and usually its magnitude can also be estimated based on electrokinetic models. This ability to probe a fundamental surface property by experimental means makes it powerful in understanding many things about the system. Here are some applications of the zeta potential:

1. Surface charge density: By applying theories of zeta potential, an estimate of the surface

charge density can be made. Another independent estimate of surface charge density can be

obtained by other means such as . Comparing these two estimates provides us insights

into the solid-solution interface, into the charging of surface groups, arrangement of ions

around the charged surface and adsorption of ions from solution.

2. Electrostatic forces: Zeta potential is one of the key parameters in estimating the electrostatic

forces between particles and/or surfaces. Electrostatic forces play an important role in

stabilizing colloidal suspensions against aggregation, or in adhesion of colloids (such as

particles or bacteria) to surfaces.

3. Electrophoresis: The zeta potential is directly proportional to the speed with which a charged

particle moves in an electric field. Electrophoresis is used to separate proteins on the basis of

their charge. Electrophoresis is also responsible for the motion of charged particles in a salt

concentration gradient where an electric field is generated internally, through a phenomenon

called diffusiophoresis. It is present in many natural situations or can be designed into the

23

system to our advantage. Another example where an internally generated electric field causes

electrophoresis is the autonomous motion of Pt-Au catalytic nanomotors.

4. Electro-osmosis: Fluid near a fixed, charged surface also moves under the application of an

electric field (electro-osmosis). Electro-osmosis is used as a mechanism to drive fluid flow or

mixing in microfluidic channels since it offers better control and mixing compared to pressure

driven flows.

5. Adsorption: When a surfactant, protein, polymer molecule or organic matter is adsorbed to a

surface, the zeta potential of the substrate can change in both magnitude and sign. The zeta

potential changes only up to the point of complete coverage. Therefore, zeta potential

measurements are often used to determine the extent of adsorption of a molecule on to a

substrate.

2.3 Classical theory of the EDL

2.3.1 Spatial distribution of charge – the Poisson-Boltzmann equation

The classical electrokinetic theory includes a description of the spatial distribution of potential or in other words, a description of the electrical double layer. The Poisson-Boltzmann (PB) equation gives the spatial distribution of the surface potential, 0. For a Z:Z electrolyte, it is written as :

푍푒 푍푒 2 ( 0) = 2 sinh 0 (1) 푘푇 푘푇

The Debye-Hückel parameter,  is a constant in this equation and depends mainly on the solution concentration, C∞. This is a very important parameter that we will come across repeatedly in this dissertation.

The inverse of  is the Debye length, 휅−1, which is the length scale over which the potential decays in the electrical double layer. For a Z:Z electrolyte, it is given as :

24

2 푍2푒2퐶 휅2 = ∞ (2) 휀푘푇

-19 -10 Where, e = charge on an electron = 1.602 x 10 C,  = dielectric permittivity of the medium = 6.9 x 10

C2/Nm2 for water at 298 K, k = Boltzmann constant, 1.38 x 10-23 J/K and T = temperature.

As can be seen from equation 2, the Debye length is proportional to the square root of concentration. Typical values of the Debye length span from ~ 100 nm in de-ionized water to 9.65 nm in 1 mM KCl, to 0.965 nm in 100 mM KCl.

2.3.2 Relationship between charge density and surface potential – the Gouy-Chapman model

In addition to a description of the spatial distribution of potential, the classical electrokinetic theory also describes the relationship between the surface charge density and surface potential. An approximate form of this relationship is obtained by using an approximate form of the PB theory - the linearized Poisson-

Boltzmann equation (the Debye- Hückel model) :

 = 0 (3)

푘푇 where  = surface charge density (C/m2). This approximation is only valid for small  , i. e.  ~ = 0 0 푒

25 푚푉. For large 0, solving the full PB equation gives the Gouy Chapman (GC) model of the EDL :

 푘푇 푍푒 4 푍푒  = 0  [2 sinh ( 0) + 푡푎푛ℎ ( 0)] (4) 푍푒 2푘푇 푎 4푘푇

In using these equations with experimental data, the surface potential, 0 is often approximated as the  potential which is a parameter more relatable to the experiments.

25

2.4 Methods of zeta potential measurement

 potential is usually measured using either microelectrophoresis, streaming potential or electroacoustics.

While the definition of  potential itself is ambiguous because the Stern layer cannot be located experimentally, these 3 techniques measure the same potential12.

2.4.1 Electrophoresis based techniques

Electrophoresis is the translation of charged particles when subjected to an electric field, with a velocity directly proportional to the  potential (see section 2.6.1). Micro-electrophoresis is used to measure the zeta potential of particles by measuring the electrophoretic velocity of individual particles using optical microscopy. This was accomplished using the Rank Brother’s instrument13 popular 2 decades ago and by its modern counterparts available commercially. The more popular modern measurement method uses light scattering instead of optical microscopy to observe the velocity of millions of particles in a few minutes, using (DLS) found in instruments manufactured by Malvern and Brookhaven.

Usually a known electric field is applied across an electrophoretic cell and the  potential is calculated using

Eq 5 or the more detailed model of O’ Brien and White14. While DLS offers quick measurements for even nano-sized particles, electrophoresis using microscopy offers the ability to directly observe the particle in motion, allowing easy and unambiguous identification of erroneous behavior such as low stability, or changes in the particles appearance with time. Moreover, DLS requires knowledge of the scattering intensity of the surface and poses a problem when a single particle consists of 2 broadly different scattering intensities, as is the case with metal/polystyrene Janus particles discussed in chapter 3.

26

2.4.2 Streaming potential

When a known pressure gradient is applied across 2 ends of a fixed surface exposed to water, the pressure driven flow carries with it ions which leads to the generation of an electric current, called the . Measurement of this current by short circuiting the 2 ends of the surface externally allows estimation of the  potential.

2.4.3 Electroacoustics

When a sound wave in the MHz range is applied to a concentrated suspension, it sets the particles and fluid into motion. But since the particles are denser than the fluid, there is relative motion between particles and the fluid. This causes a charge separation because the Stern layer moves with the particles whereas the diffuse layer moves with the fluid. As these dipoles oscillate with the same frequency as the sound wave, an alternating current called the Colloid Vibration Current (CVI) is generated which is measured by externally short circuiting the two ends of the suspension. The magnitude of CVI is directly proportional to

 potential15. In this technique, the signal is enhanced by higher solids concentration and higher density difference between particles and the suspension. Thus, this technique is ideal for concentrated dispersions of high density contrast particles such as mineral particles, but very difficult to use for polymer latex particles or dilute suspensions.

Today, with the availability of much better video microscopy techniques and particle tracking algorithms, microelectrophoresis using microscopy is the answer to many challenging systems and this is the technique which will be used primarily in this work. There are several reasons for choosing this technique over the other two methods mentioned above. While the streaming potential technique is very popular for measurements on rock surfaces, it suffers from 2 distinct disadvantages – (1) the measurement relies heavily on accurate knowledge of the geometry of the surface which is not always possible or easily controlled, and

(2) its application for particles is not straightforward. The electroacoustic technique is not suited for dilute

27 systems which limits its use for active materials such as Janus motors which usually suffer from very low yields. Moreover, electroacoustic measurements are hard to interpret for polydisperse or heterogeneous systems. For example, water from shale gas reserves contains a variety of minerals such as silica, alumina, calcium carbonate, etc. which may also be of different sizes and densities. Under such conditions, only an average value may be obtained using electroacoustics which is usually not enough to predict the system stability.

2.5 Colloidal stability

Formulating stable colloidal suspensions is a classical problem in colloid and interface science. By stable suspension, we mean a suspension where the particles do not aggregate with each other because if they did, the settling rate would increase and the particles will fall out of suspension. Usually, this is not desirable.

For example, acrylic paints are concentrated colloidal suspensions and stability is necessary for a long shelf life of the paint. Other examples where making a stable formulation is an important industrial problem is in drug formulation, food, and cosmetics.

In some other cases, aggregation is desirable. For example, in order to separate milk into curd and whey, an acid such as lime or vinegar is added to a milk (a stable colloidal suspension) causing the suspended proteins to aggregate (or coagulate) and fall out of solution. Another example is the Salting-Quenching-

Fusion (SQF)16 technique of assembling colloidal doublets, developed in the Velegol lab.

The question then is why are particles attracted or repelled by each other, and how can we control this behavior? The answer lies in an understanding of the colloidal forces and energy. The sign, magnitude and spatial range of each of the colloidal forces controls the overall interaction of particles with each other or with other surfaces. In this section, I will introduce the DLVO theory of colloidal stability and the forces included in its description – electrostatic and van der Waals forces. A detailed description of the forces can be found in the classic book by Israelachvili17 or in the pedagogic book by Velegol.

28

2.5.1 DLVO theory

The DLVO theory of colloidal stability introduced by Derjaguin and Landau (1941) and Verwey and

Overbeek (1948) says that the total interaction between any two surfaces is given by the sum of van der

Waals forces (always attractive) and electrostatic forces (repulsive between surfaces with like charge). The total energy can be plotted as a function of distance between the particles in a DLVO plot, an example of which is shown in Figure 2-2. A positive energy on this plot indicates repulsive interaction and a negative energy represents attraction. The plot shows the sum of electrostatic and van der Waals energies as the solid blue line. In the example shown, the total energy is non-monotonic, going through a secondary minimum, a maximum and a primary minimum at negative infinity. The reason for such behavior lies in the different length scales over each the repulsive electrostatic and attractive van der Waals forces act.

The van der Waals forces between two surfaces are always attractive and relatively long range – the energy scales as 1/ for colloidal particles where  is the separation between particles (<< particle diameter. Also note this is unlike the behavior of van der Waals energy between atoms which scales at 1/6). The range of the electrostatic forces however, depends on the Debye length and acts up to about 3 Debye lengths away from the particle’s surface. Thus, the total interaction varies as a function of distance between the particles.

29

Figure 2-2: A sample DLVO plot. The parameters are :  potential = 100 mV, Debye length = 1 nm, particle diameter 2 µm, Hamaker constant = 1.4 x 10-19 J.

2.5.2 Rapid aggregation

When the DLVO plot looks like the one shown above, particles can aggregate into a secondary minimum if they come close enough. In this state, the particles can be separated by putting some energy into the system such as sonication for a few minutes. Thermodynamically, they would like to aggregate at the primary minimum where the energy is lowest, but the energy barrier, max (the height of the maxima) may be too high for them to do so in a practical time-frame. If the particles cannot attain the primary minimum, the suspension remains kinetically stable.

Particles do have the ability to overcome small energy barriers due to their thermal energy, given by kT.

So, if max~ kT, the particles can in fact overcome the barrier and go to the primary minimum within some

30

finite time-frame. When max  3kT, particles aggregate each time they collide and this is called the regime of rapid aggregation. For a suspension with particles of radius a and volume fraction , the rapid aggregation time scale is given by

푎3  = (5) 2푘푇

Thus, even for a suspension where there is no repulsive energy barrier, the volume fraction controls how long it takes to aggregate.

2.6 Electrokinetics

The study of the transport of particles or of fluid under the influence of an electric field is termed electrokinetics. A brief description of 2 electrokinetic phenomena – electrophoresis and diffusiophoresis is presented below. As we will see below, the electric field could be either externally applied or be generated internally.

2.6.1 Electrophoresis and electro-osmosis

2.6.1.1 Electrophoresis

When a particle with a  potential, p is subjected to an electric field, E it moves with an electrophoretic velocity given by the Smoluchowski equation18:

 퐸 푢 = 푝 (6) 푒푝 

Where u = electrophoretic velocity, and  = viscosity of the fluid. This equation is valid if the  potential is ~ kT/e = 25 mV and a >> 1. While this equation was derived for a spherical particle, it was shown by

Morisson19 that it is valid for a particle of arbitrary shape. For higher values of  potential, O’Brien &

31

White20 showed that the particle double layer undergoes polarization, lowering the electrophoretic velocity.

Thus, the electrophoretic velocity actually undergoes a maxima as  increases rather than the monotonic velocity increase with  as predicted by Eq 2.

2.6.1.2 Electroosmosis

When a fixed surface is exposed to a high dielectric constant protic solvent such as water, it similarly attains a  potential, w which leads to the formation of the EDL. In this case, the application of an electric field causes the motion of ions in the double layer called electroosmosis or electroosmotic flow (EOF), given by18

 퐸 푢 = − 푤 (7) 푒표 

2.6.1.3 Electrophoresis in a closed capillary

When an electric field is applied to a suspension in a closed capillary, the EOF from the glass surface causes a back-pressure at the closed end of the capillary which returns as parabolic back flow (Figure 2-3).

Figure 2-3: Fluid and particle velocity when an electric field is applied on fluid in a closed capillary.

32

Bowen21 has calculated the expression for velocity of a particle in a rectangular capillary as a function of height, z when the view is centered on the width of the capillary

32 cos(휋푧⁄2푏) 1 cos(휋푧⁄2푏) 푏2 [1 − ( − + ⋯ )] − 푧2 (8) −휀퐸 3 휋3 cosh(휋푎⁄2푏) 33 cos(휋푎⁄2푏) 푢(0, 푧) = [(1 − { })  −  ] 2 192푏 휋푎 1 3휋푎 푤 푝 휂 2푏 1 − (tanh + tanh + ⋯ ) 휋5푎 2푏 35 2푏

Where a = half width of the capillary, and b = half height of the capillary

Eq 4 is linear in p and w. When the u vs. z profile is measured experimentally, it can be used to obtain unique best-fit values of p and w. Another way often used to measure p is to simply measure the velocity at the stationary plane, also called Komagata plane where the fluid velocity is 0 in the cell. The velocity at this plane is entirely due to the particle’s motion. Though simpler in calculation, the accuracy of zeta potential determined by this method depends on the accurate location of the Komagata plane, and slight errors in locating the plane can result in large errors in the measured zeta potential.

2.6.2 Diffusiophoresis and diffusioosmosis

Diffusiophoresis is the chemically-driven transport of colloidal particles driven by a concentration gradient of solute, without the application of any outside force.22 There are two general types of diffusiophoresis: electrolyte and non-electrolyte. Electrolyte diffusiophoresis causes particle transport by generating spontaneous electric fields (electrophoresis)23 and pressure fields in the Debye layer (chemiphoresis)24,25.

Non-electrolyte diffusiophoresis (NEDP) also works by creating a pressure gradient across the particle due to particle-solute interactions, such as repulsive steric exclusion26,27 or attractive van der Waals interactions28,29, and this also is a type of chemiphoresis. Diffusiophoresis has been discussed in detail in this thesis in Appendix 3 which is based on a review article I co-authored.

33

2.7 References

(1) Lyklema, J. Water at Interfaces: A Colloid-Chemical Approach. J. Colloid Interface Sci. 1977, 58,

242–250.

(2) Kar, A.; Chiang, T.-Y.; Ortiz Rivera, I.; Sen, A.; Velegol, D. Enhanced Transport into and out of

Dead-End Pores. ACS Nano 2015, 9, 746–753.

(3) Roger, K.; Cabane, B. Why Are Hydrophobic/water Interfaces Negatively Charged? Angew. Chemie

- Int. Ed. 2012, 51, 5625–5628.

(4) Garg, A.; Cartier, C. A.; Bishop, K. J. M.; Velegol, D. Particle Zeta Potentials Remain Finite in

Saturated Salt Solutions. Langmuir 2016, acs.langmuir.6b02824.

(5) Behrens, S. H.; Christl, D. I.; Emmerzael, R.; Schurtenberger, P.; Borkovec, M. Charging and

Aggregation Properties of Carboxyl Latex Particles : Experiments versus DLVO Theory. 2000,

2566–2575.

(6) Evenhuis, C. J.; Guijt, R. M.; Macka, M.; Marriott, P. J.; Haddad, P. R. Variation of Zeta-Potential

with Temperature in Fused-Silica Capillaries Used for Capillary Electrophoresis. Electrophoresis

2006, 27, 672–676.

(7) Das, S.; Garg, A.; Campbell, A. I.; Howse, J.; Sen, A.; Velegol, D.; Golestanian, R.; Ebbens, S. J.

Boundaries Can Steer Active Janus Spheres. Nat. Commun. 2015.

(8) Oncsik, T.; Trefalt, G.; Csendes, Z.; Szilagyi, I.; Borkovec, M. Aggregation of Negatively Charged

Colloidal Particles in the Presence of Multivalent Cations. Langmuir 2014, 30, 733–741.

(9) Reynolds, E. C.; Wong, A. Effect of Adsorbed Protein on Hydroxyapatite Zeta Potential and

Streptococcus Mutans Adherence. Infect. Immun. 1983, 39, 1285–1290.

(10) Meyer, E. E.; Lin, Q.; Hassenkam, T.; Oroudjev, E.; Israelachvili, J. N. Origin of the Long-Range

Attraction between Surfactant-Coated Surfaces. Proc. Natl. Acad. Sci. U. S. A. 2005, 102, 6839–

34

6842.

(11) Borkovec, M.; Behrens, S. H.; Semmler, M. Observation of the Mobility Maximum Predicted by

Standard Electrokinetic Model for Highly Charged Amidine Latex Particles. Langmuir 2000, 16,

5209–5212.

(12) Delgado, a. V.; González-Caballero, F.; Hunter, R. J.; Koopal, L. K.; Lyklema, J. Measurement and

Interpretation of Electrokinetic Phenomena (IUPAC Technical Report). Pure Appl. Chem. 2005, 77.

(13) Bangham, A. D.; Flemans, R.; Heard, D. H.; Seaman, G. V. F. An Apparatus for

Microelectrophoresis of Small Particles. Nature 1958, 182, 642–644.

(14) O’Brien, R. W.; White, L. R. Electrophoretic Mobility of a Spherical Colloidal Particle. J. Chem.

Soc. Faraday Trans. 2 1978, 74, 1607.

(15) Dukhin, A. S.; Goetz, P. J. Characterization of Liquids, Nano- and Microparticulates, and Porous

Bodies Using Ultrasound; Elsevier, 2002.

(16) Yake, A. M.; Panella, R. a; Snyder, C. E.; Velegol, D. Fabrication of Colloidal Doublets by a Salting

out-Quenching-Fusing Technique. Langmuir 2006, 22, 9135–9141.

(17) Israelachvili, J. N. Intermolecular and Surface Forces: Third Edition; Academic Press, 2011.

(18) Hunter, R. J. Zeta Potential in Colloid Science: Principles and Applications; Academic Press, 1981.

(19) Morrison, F. . Electrophoresis of a Particle of Arbitrary Shape. J. Colloid Interface Sci. 1970, 34,

210–214.

(20) O’Brien, R. W.; White, L. R. Electrophoretic Mobility of a Spherical Colloidal Particle. J. Chem.

Soc. Faraday Trans. 2 1978, 74, 1607.

(21) Bowen, B. D. Effect of a Finite Half-Width on Combined Electroosmosis - Electrophoresis

Measurements in a Rectangular Cell. J. Colloid Interface Sci. 1981, 82, 574–576.

(22) Anderson, J. Colloid Transport By Interfacial Forces. Annu. Rev. Fluid Mech. 1989, 21, 61–99.

35

(23) Lin, M. M.-J.; Prieve, D. C. Electromigration of Latex Induced by a Salt Gradient. J. Colloid

Interface Sci. 1983, 95, 327–339.

(24) Lechnick, W. J.; Shaeiwitz, J. A. Measurement of Diffusiophoresis in Liquids. J. Colloid Interface

Sci. 1984, 102, 71–87.

(25) Lechnick, W. J.; Shaeiwitz, J. A. Electrolyte Concentration Dependence of Diffusiophoresis in

Liquids. J. Colloid Interface Sci. 1985, 104, 456–470.

(26) Staffeld, P. O.; Quinn, J. A. Diffusion-Induced Banding of Colloid Particles via Diffusiophoresis: 1.

Electrolytes. J. Colloid Interface Sci. 1989, 130, 69–87.

(27) Staffeld, P. O.; Quinn, J. A. Diffusion-Induced Banding of Colloid Particles via Diffusiophoresis 1.

Electrolytes. J. Colloid Interface Sci. 1989, 130, 69–87.

(28) Sharifi-Mood, N.; Koplik, J.; Maldarelli, C. Molecular Dynamics Simulation of the Motion of

Colloidal Nanoparticles in a Solute Concentration Gradient and a Comparison to the Continuum

Limit. Phys. Rev. Lett. 2013, 111, 184501.

(29) Sharifi-Mood, N.; Koplik, J.; Maldarelli, C. Diffusiophoretic Self-Propulsion of Colloids Driven by

a Surface Reaction: The Sub-Micron Particle Regime for Exponential and van Der Waals

Interactions. Phys. Fluids 2013, 25, 12001.

36

Chapter 3

Particle Zeta Potentials Remain Finite in Saturated Salt Solutions1

3.1 Abstract

The zeta potential of a particle characterizes its motion in an electric field and is often thought to be negligible at high ionic strength (several moles per liter) due to thinning of the electrical double layer (EDL). Here, we describe zeta potential measurements on polystyrene latex (PSL) particles at monovalent salt concentrations up to saturation (~ 5 M NaCl) using electrophoresis in sinusoidal electric fields and high-speed video microscopy. Our measurements reveal that the zeta potential remains finite at even the highest salt concentrations. Moreover, we find that the zeta potentials of sulfated PSL particles continue to obey the classical Gouy-Chapman model up to saturation despite significant violations in the model’s underlying assumptions. By contrast, amidine-functionalized

PSL particles exhibit qualitatively different behaviors such as zero zeta potentials at high concentrations of NaCl and KCl and even charge inversion in KBr solutions. The experimental results are reproduced and explained by Monte Carlo simulations of a simple lattice model of the

EDL that accounts for effects due to ion size and ion-ion correlations. At high salt conditions, the model suggests that quantitative changes in the magnitude of surface charge can result in qualitative

1 This chapter has been adapted and reprinted with permission from A. Garg, C. Cartier, K. Bishop, D. Velegol, Particle Zeta Potentials Remain Finite in Saturated Salt Solutions. Langmuir, 32, 11837. Copyright 2016 American Chemical Society.

37 changes in the zeta potential – most notably, charge inversion of highly charged surfaces. These findings have important implications for electrokinetic phenomena such as diffusiophoresis within salty environments such as oceans, geological reservoirs, and living organisms.

3.2 Introduction

The zeta potential () of a colloidal particle determines the particle’s stability in solution and its motion in an electric field. Physically, this potential difference derives from the separation of charge bound to the particle’s surface (physically or chemically) from that of mobile counter-ions in solution. Charge separation over a finite length scale is driven by thermal fluctuations and gives rise to the so-called electric double layer (EDL). In the classical Gouy-Chapman (GC) model, this

−1 2 1/2 length scale is identified as the Debye screening length,  = (푘퐵푇/2푒 푛0) , which depends on the dielectric permittivity (  ), thermal energy ( 푘퐵푇 ), elementary charge ( 푒 ), and salt concentration (푛0) (here assuming a 1:1 electrolyte). Despite their widespread application and effectiveness, the GC model and its derivatives make a variety of assumptions that break down at high salt concentrations1 that approach saturation. Notably, these continuum descriptions neglect the finite size of ions and molecules2–4, which are actually larger than the predicted screening length in concentrated electrolytes (Figure 3-1b,c). Furthermore, the mean-field approximations used to describe electrostatic interactions fail to account for ion-ion correlations4–6 that ultimately guide crystallization from saturated solutions. At high salt concentrations, the predicted screening length

2 becomes smaller than the Bjerrum length,  = 푒 /4푘퐵푇, over which ion-ion correlations are significant7 (Figure 3-1b). For particles in salty media, one might therefore anticipate that the structure of the EDL and the corresponding zeta potentials will differ significantly and qualitatively from expectations built largely on the study of dilute electrolytes. It is critical to identify and understand these differences in order to predict the behaviors of colloids in salty environments such

38 as human blood ( 150 mM), seawater ( 600 mM), wastewater ( 2M during reverse osmosis), and geological reservoirs ( 5 M).

The standard methods used to determine zeta potentials of particles at low ionic strengths are often inapplicable at high salt concentrations. Methods based on electrophoresis measure the velocity of particles in an applied electric field. Application of such fields within concentrated electrolytes results in large electric currents that polarize electrodes and rapidly heat the sample. The resulting temperature variations can induce significant convective flows and complicate the estimation of fluid properties needed to infer particle zeta potentials (e.g., dielectric constant and viscosity). Due to such challenges, few studies reliably report zeta potentials for colloidal particles at high ionic strength. Dilute latex particle dispersions have been examined by phase angle light scattering

(PALS) for salt concentrations up to 3 M8,9 and by a combination of optical tweezers and high- speed microscopy up to 1 M10. Concentrated inorganic mineral particle systems have been examined using the electroacoustic method11–14 at salt concentrations as high as 3 M. These studies found finite zeta potentials for particles in 1-3 M monovalent salt solutions as well as concentration- dependent shifts in the isoelectric point12,14. However, we know of no reports that investigate particle zeta potentials at salt concentrations at or near saturation (5.4 M for NaCl) where finite size effects and ion-ion correlations are most pronounced. In strongly interacting charged systems (e.g., multivalent electrolytes), ion-ion correlations are known to drive counterintuitive phenomena such as charge inversion (also called charge reversal), whereby an excess of counter-ions binds to a charged surface to reverse its polarity15. Similar effects have been hypothesized for monovalent electrolytes near saturation. Additionally, it has been suggested that the EDL can never be thinner than the finite size of the counter-ions, leading to the failure of GC predictions at high ionic strength16. Within such thin double layers, changes in the structure of interfacial water may also contribute to charge separation and the associated zeta potential – even for uncharged surfaces13.

39 Investigating these and other interesting hypotheses requires experimental data on particle zeta potentials at saturated salt conditions.

Here, we describe a method for measuring particle mobilities up to saturated salt conditions and present data for polystyrene latex particles in several monovalent salt solutions (KCl, KBr, and

NaCl). We use electrophoresis with a sinusoidal electric field (300 Hz) to drive oscillatory particle motions, which are captured and quantified using high-speed video microscopy. An apparent zeta potential is then inferred from the measured mobility using the Smoluchowski equation. For negatively charged, sulfated polystyrene latex particles (sPSL), we find finite zeta potentials of

  − 20 mV at saturated salt conditions. Moreover, the measured zeta potentials and their dependence on salt concentration are found to be consistent – within experimental uncertainty – with predictions of the GC model despite violating many of its foundational assumptions. By contrast, positively charged, amidine-functionalized PSL (aPSL) exhibit more complex behaviors including charge reversal in KBr, LiCl and NaCl solutions. To guide the interpretation of our experimental results, we investigate a lattice model of the EDL that incorporates effects due to finite ion size and ion-ion correlations. By comparing the results of this model to those of the standard GC model, we explain how the latter can make successful predictions at even the highest salt concentrations for weak ion-ion coupling but breaks down when such coupling is strengthened.

Moreover, we show that breakdown of the GC model is accompanied by charge inversion, in qualitative agreement with experimental observations for aPSL. Our results indicate that electrokinetic phenomena – particularly diffusiophoresis17–20 – are present even in saturated salt solutions and may therefore contribute to the many transport processes occurring in oceans, geological reservoirs, and living organisms. 

40

Figure 3-1: (a) Schematic illustration of the experimental setup showing polystyrene latex (PSL) particle in a glass capillary subject to an oscillating electric field. The electrolyte solution is characterized by its viscosity (), conductivity (ke), permittivity (), and salt concentration (n0). (b) The Debye length (-−1) decreases as a function of NaCl concentration becoming smaller than the Bjerrum length () at 0.3 M and the ions themselves (ca. 0.3 nm) above 2 M. (c) An illustration of the EDL at two different salt concentrations highlighting the relevant length scales such as ion size, Debye length (shown on left and right edges for low and high salt, respectively), ion-ion spacing, and surface charge spacing (red semi-circles).

41 3.3 Methods and Materials:

3.3.1 Measuring Zeta Potentials

We used the sinusoidal response of particles subject to a sinusoidal electric field to measure the electrophoretic mobility and thereby infer the zeta potential. In a typical experiment (see Figure

3-1a), a dilute suspension of polystyrene latex microparticles (diameter 푎  3 μm) in an aqueous electrolyte was flowed into a glass capillary. The particles were subject to an oscillating electric field of magnitude 퐸0 and frequency 푓0 = 휔0/2휋 directed along the length of a capillary, 퐸(푡) =

퐸0 sin(휔0푡). The motion of a single particle was captured using high-speed video microscopy, and its dynamic trajectory 푥(푡) was reconstructed using particle tracking algorithms (see below). For particles much larger than the screening length (휅푎 ≫ 1), the electrophoretic particle velocity 푢 is related to the applied field by the Smoluchowski equation21

 푢 = 휇퐸 = 푠푚 퐸, (1) 

where 휇 is the electrophoretic mobility, 휂 is the fluid viscosity, and 푠푚 is the Smoluchowski zeta potential. Integrating this equation, we obtain the following expression for the anticipated particle trajectory

푠푚 푥(푡) = − 퐸0 cos(휔0푡). (2) 휔0

The reported zeta potentials were obtained by linear regression of particle trajectories using equation (2).

42 3.3.2 Experimental Details

We collected data on three different types of surfactant-free, polystyrene latex particles (density,

 = 1055 kg/m3): sulfated PSL (radius, a =2.9 m; surface charge density,  = −9.7 C/cm2), carboxylated PSL(a = 3.2 m;  = −11.9 C/cm2), and amidine functionalized PSL (a = 3.3 m;

 = 34.9 C/cm2) henceforth denoted sPSL, cPSL, and aPSL, respectively. The particles were suspended at low volume fractions (10-5) into freshly prepared, aqueous solutions of ACS grade

NaCl, KCl, KBr, LiCl and CsCl. Salt solutions were freshly prepared and their conductivity was measured using a Hach H170 pH and conductivity meter. Conductivity measurements are shown in Figure 3-4 d. The measured current i, conductivity ke, and cross-sectional area of the cell A were used to determine the applied electric field as E = i/keA.

The suspension was drawn into an RCA-1 cleaned 1 x 1 x 50 mm glass capillary (Vitrocom) positioned on a glass slide and gold electrodes (ca. 1 cm of 16 strands of 0.003 inch diameter wire,

A M Systems) were inserted at each end. The capillary was sealed using Bondic UV glue and was immediately relocated to the microscope (Zeiss Axio Imager A1m upright microscope) stage for imaging. A 50x long working distance objective was used to focus in the central plane of the capillary (scale of 0.25 µm per pixel). A sinusoidal voltage was applied using a function generator

(Keithley Model 3390) and a voltage amplifier (Trek Model 2205); current through the circuit was monitored (Keithley 2612 A) using a LabVIEW interface (schematic in Figure 3-2).

The sinusoidal voltage had a frequency 푓0 = 300 Hz (Figure 3-2) and was applied for for 0.5 – 2 s. The measured current i, conductivity ke, and cross-sectional area of the cell A were used to determine the applied electric field as E = i/keA (E = 300 – 1000 V/m). Simultaneous to applying the voltage, the motion of a single particle was observed with a microscope using a 50x objective focused near the center of the capillary (scale of 0.25 μm per pixel). A high-speed camera (Phantom

43 v310) synchronized with the voltage source recorded the particle’s motion at 6000 frames per second. Gravitational forces induced particle motions perpendicular to the applied field; however, the particles remained in focus for at least 1 s – even in the densest electrolyte (4 M CsCl,  = 1498 kg/m3).

Our experimental setup was designed to mitigate a variety of challenges that arise when measuring zeta potentials at high salts. The use of AC fields was essential to avoid the rapid polarization of the electrodes at high salt concentrations, at which the electric current density can exceed 30 mA/mm2. To prevent oscillatory electroosmotic flows in the center of the channel, the applied frequency was chosen to be much faster than the rate of momentum diffusion over the width 푊 of the capillary – that is, 휔 ≫ 휈/푊2~1 s−1 where 휈 is the kinematic viscosity22. At the same time, the frequency was chosen to be small enough that field-induced particle displacements were clearly distinguishable from those due to Brownian motion. The cell was not thermostated; however, the temperature increase was mitigated by applying the voltage for short times to reduce heat generation and by using a narrow capillary to increase the rate of heat transfer to the surroundings.

At the highest power input of 1.3 W, the measured current and hence the conductivity increased by

3.9% in 2 s, which corresponds to a temperature increase of 1.9°C using a typical coefficient of conductivity variation23. Further details of the experimental set-up and problems encountered are discussed below.

44 3.3.2.1 Heat generation and convective flows

The cell was not thermostated; however, heat generation was minimized by applying the voltage for short times of 0.5-2 s and by using a cell with high surface to volume ratio. During a 2 s time period, we measured a change in current and hence conductivity of 3.87% for the highest electrical power of 1.3W. This corresponds to a temperature change of 1.9 ◦C using a typical coefficient of conductivity variation = 2 %/◦C.

Alternatively, we could estimate the temperature increase assuming that all heat dissipated

 by the ionic current contributes to heating the electrolyte in the cell: ∆T = iEtexp=ρcp = 12.2 C, where i is the measured current density (29.9 mA/mm2), E is the applied electric field (978 V/m),

3  texp is the measurement duration (2s), ρ is the density (1133 Kg/m ), and cp (4186 J/Kg/ C) is the specific heat. This gives a much higher value since it ignores the heat dissipated to the surroundings through the capillary walls. Even for short experiments and small temperature increases, natural convection cannot be avoided at the highest salt concentrations. As discussed in Section 2.4 of the

SI, thermally driven flows are responsible for steady particle motions and for those at two times the driving frequency. We emphasize that these effects do not impact electrophoretic particle motions at the driving frequency.

3.3.2.2 Gravity

We should emphasize that the electric field is horizontal while settling is vertical. Thus, our measurements are affected by gravity only to the extent that it causes the particles to go out of focus.

The polystyrene latex particles used have a density of 1055 kg/m3. Even for the densest suspension we used (4 M CsCl, ρ =1598 kg/m3), the particles rose at a rate of 2.3 µm/s, which kept them in focus for 1 s (with a numerical aperture of 0.55, the depth of field was 2.5 µm).

45 3.3.2.3 Electroosmotic flows

By contrast to thermal flows, electroosmotic flows are potentially problematic as they occur at the frequency of the applied voltage and could therefore interfere with measurements of electrophoretic particle motions. We prevented electroosmotic flows in the center of the channel by operating at sufficiently high frequencies. Oscillatory flows originating at the walls of the channel decay over a length (ν/)1/2, which is much smaller than the width of the channel. Thus, while EOFs are likely present near the channel walls, they are not present in the center of the channel where particle motion is measured. This reasoning is further supported by the agreement between particle zeta potentials measured at low salt conditions using our method and a commercial instrument.

3.3.2.4 Electrode polarization

The large flow of electrons produced for the high conductivity solutions used here can cause electrode polarization that can further lead to formation of bubbles. This is avoided by using a relatively high frequency alternating field (300 Hz). Also electrode surface area is increased by using bundles of 16 strands of gold wire as electrodes. We used bundled electrodes over blackened

Pt electrodes to increase surface area because this was a lot quicker than blackening, that requires an overnight operation.

46

Figure 3-2: Electronics and instrumentation. A function generator generated a sine wave that was amplified by the amplifier. The voltage from the amplifier was connected across the sample cell, and a Keithley source-measure-unit recorded the current through a labview interface. The high speed camera was synchronized with the function generator through an external trigger port.

3.3.3 Particle tracking and analysis

Dynamic particle trajectories x(t) along the length of the capillary were reconstructed from the recorded images at 10 nm resolution using the Trackmate plugin in Fiji24. Low frequency translational motions were removed by subtracting a smoothing spline fit to the noisy trajectory

(Figure 3-3).

47

Figure 3-3: Data analysis and filtering. (a)Fourier transform of the current data i(!) shows a peak only at the applied frequency. This observation suggests that particle motions at frequency 2!0 cannot be attributed to effects that depend linearly on the current. (b) Raw position vs. time for sPSL, showing that 0f0 velocity is so high that all other information is hidden due to it. (c) The filtered position, obtained by subtracting cubic splines fitted to the data split into 24 adjacent sections. This removes information only at low frequencies, f ≤ 24 Hz. The filtered position was used to calculate the fourier transforms shown in Figure 3-5 c,d. All data are for sPSL in 1 M NaCl, f0 = 300Hz and E = 829 V/m. We then decomposed the filtered particle trajectories into Fourier series of the form, 푥(푡) =

𝑖푡 ∑ 푋()푒 , and examined the (complex) Fourier components 푋() to confirm the validity of equation (2) (see below). The mobility was obtained by performing a least squares fit of the filtered position data directly to equation (2) (Figure 3-4 a). In deriving the zeta potential from the electrophoretic mobility, it is important to recognize that the viscosity and the dielectric constant of concentrated electrolytes differ significantly from that of pure water. For example, the viscosity of saturated NaCl is 70% higher than that of water while the relative permittivity (휀푟 = 휀/휀0 where

휀0 is the permittivity of free space) is 50% lower. Failure to account for these changes leads to errors in the predicted zeta potential of more than 300%. We used literature data for the

48 concentration-dependent viscosity and dielectric constant for monovalent electrolytes summarized below (Figure 3-4 b,c) to best approximate the Smoluchoswski zeta potential from the measured mobilities (Figure 3-4 a).

Figure 3-4: (a) Mobility (u/E) vs. salt concentration for sulfated PSL in various salts. Error bars are obtained from 90% confidence intervals. (b), (c) data compiled for (b) viscosity25–27] and (c) relative permittivitys28–32; at T = 25 ◦C. (d) Measured conductivity for each of the salt solutions at T = 25 ◦C. For each condition (i.e., particle surface chemistry, salt, and concentration), the reported zeta potentials were obtained by averaging at least six measurements taken on different particles. The

90% confidence intervals reported in Figure 3-9 represent the larger of two errors: that obtained

49 from linear regression on each particle and that derived from the distribution of zeta potentials measured for various particles. Variations in the zeta potential from particle to particle were typically smaller than the 90% confidence interval for each particle, indicating good repeatability.

3.3.4 Method Validation

To confirm the validity of equation (2), we examined the Fourier components 푋() to assess their dependence on the applied frequency and field strength. Equation (2) implies that 푋() should be finite and real only at the driving frequency, corresponding to a pure cosine component at 0. At low salt concentrations (10 mM NaCl), the trajectories appear roughly sinusoidal, and the dominant

Fourier component was indeed that of the driving frequency 푓0 = 300 Hz (Figure 3-5 a,c). By contrast, at high salt concentrations (5 M NaCl), the sinusoidal motion of the particle is hardly visible above the noise; however, the Fourier component at the driving frequency can still be clearly resolved (Figure 3-5 b,d). We further confirmed (i) that the phase of 푋(0) was near 0 or ±π

(Figure 3-6), (ii) that the particle velocity 0푋(0) was linearly proportional to the field strength

(Figure 3-7), and (iii) that the inferred zeta potential was independent of the driving frequency

(Figure 3-7).

50

Figure 3-5: Filtered position x(t) vs. time for a 2.9 m sPSL particle in (a) 10 mM NaCl for E = 1120 V/m and (b) 5.4 M NaCl for E = 733 V/m. Magnitude of the Fourier components for particle velocity |X()| as a function of frequency for (c) 10 mM and (d) 5.4 M NaCl, respectively. The red dot highlights the peak at the applied frequency f0 = 300 Hz. At high salt concentrations, an additional peak is seen at 600 Hz and attributed to field-induced thermal convection (section 3.4.1).

Figure 3-6: Phase of Fourier transform of position at f0 for (a) sulfated and (b) amidine- functionalized polystyrene latex particles. A phase of 0 implies ζ < 0 and a phase of ±π implies ζ > 0. For situations where the phase was neither of ±π or 0, the measurement was considered as noise.

51

Figure 3-7: If the frequency of electric field is low (∼ 1 Hz), the ζ measured would be a function of the frequency22 due to electro-osmotic flow (EOF) from the capillary walls being transmitted to the center. (a) We show here that the measured ζ potential of sulfated PSL particles does not depend on frequency for f0 ≥ 300 Hz. Lower frequencies were not applied because electrode degradation led to strong flows. (b) The particle velocity at the applied frequency is linear with E (i.e., mobility does not depend on electric field). This validates that we are operating in the linear response regime.

Interestingly, an additional component at two times the driving frequency 2푓0 was observed at high salt concentrations. The magnitude of this component increased linearly with the electric power supplied, indicating a second order dependence on the applied field (Figure 3-8 a,b). As detailed in the results section 3.4.1, we attribute these observations to thermally-induced convective flows, which depend on the rate of Joule heating within the capillary. Importantly, these effects are second order in the field and do not contribute to the first order, electrophoretic motions of interest here

(nonlinear electrophoretic effects are third order in the field as required by symmetry33).

Finally, we performed an independent validation of our measurement technique by comparing its predictions to those of a commercial PALS-based instrument (Malvern Zetasizer Nano ZS90) at lower salt concentrations. For 2.9 m sPSL particles in 1 mM NaCl, our setup gave a zeta potential of −110 ± 2.4 mV compared with −107 ± 4.8 mV from the commercial instrument.

52 3.3.5 Monte Carlo Simulations of the EDL

We performed Monte Carlo simulations of the EDL at high salt concentrations using the lattice restricted primitive model, which accounts for effects due to ion size and ion-ion correlations

(figure 3-11 a). In the model, ions are represented by point charges (±푒) positioned within a uniform dielectric medium onto discrete lattice sites separated by a distance, ℓ. Physically, the lattice spacing mimics effects due to ion size, and the parameter ℓ can be interpreted as the diameter of a solvated ion (typically, 0.2-0.5 nm). Initially, M surface charges, M counterions, and N ion pairs are distributed onto a rectangular simulation cell of dimensions 퐿1×퐿2× 퐿3. The surface charges are distributed at random within the 푥3 = 0 plane (figure 3-11 a) to create an average surface charge density of 휎 = 푒 푀/2퐿1퐿2 (the factor of two corrects for the fact that the surface is bounded by the electrolyte on two sides and not just one, as in experiment). Similarly, the 2푁 +

푀 ions are distributed at random onto bulk lattice sites (푥3 ≠ 0) to give a nominal salt concentration of 푛푐 = 푁/퐿1 퐿2 (퐿3 − ℓ) within the cell.

The positions of the ions are equilibrated using the Metropolis Monte Carlo algorithm34; those of the surface charges are fixed throughout the simulation. During each Monte Carlo move, two sites are selected at random and their contents swapped. If the move lowers the electrostatic energy of the system, it is accepted unconditionally. If it raises the energy of the system, it is accepted with probability p(Accept) = exp(−훥푈/푘푇) where 푘푇 is the thermal energy, and 훥푈 is the energy increase accompanying the change in ion configuration. For lower salt concentrations, several such lattice swaps are conducted during each Monte Carlo move to achieve an acceptance frequency of around 50%. During each simulation, the system is equilibrated for 5×105 attempted moves; the resulting distribution is then sampled over the course of an additional 5×105 moves. Each condition is simulated twelve times using different realizations of the surface charge distribution;

53 the final results are obtained by averaging over these realizations. The electrostatic energy 푈 is computed using the Ewald method34, which is described in detail in Appendix 1

The behavior of this model is fully specified by three dimensionless parameters: (i) the salt

−3 concentration within the periodic cell 푛푐 (scaled by ℓ ), (ii) the surface charge density 휎 (scaled

2 2 by 푒/ℓ ), and (iii) the Bjerrum length, 휆 = 푒 /4휋휀푘퐵푇 (scaled by ℓ), which characterizes the strength of ionic interactions. At room temperature (25 oC),  ranges from 0.71 nm in pure water to ~1.3 nm in 5 M NaCl, which is ca. 1 to 4 times the characteristic ion diameter (ℓ = 0.4 nm).

3.4 Results & Discussion

3.4.1 Particle Motions at Frequency 20

At high salt conditions (salt concentration ≥ 500 mM), we observe a second peak in the power spectrum at two times the driving frequency - that is, at 20 (Figure 3-5). We note that this peak is not seen in the current data (Figure 3-3). We hypothesize that this effect is due to convective flows that arise from electrically induced thermal gradients within the cell. Here, we present an analysis of these flows and provide additional experimental data that supports our hypothesis. We approximate the experimental cell as a long channel between two parallel boundaries separated by a distance 2H (Figure 3-1 a). Application of an electric field E(t) = E0 cos(0t) down the length of the channel results in a steady current density j(t) = keE(t), where ke is the constant conductivity.

Joule heating due to the electric current induces thermal gradients within the cell as described by the following equation for thermal transport

(3)

54 where ρ, cp, and k are the constant density, specific heat, and thermal conductivity of the electrolyte, and v is the convective velocity. The boundaries of the channel are modeled as

(4) where h is a heat transfer coefficient. Assuming there are no fluid flows (v = 0), this differential equation can be solved analytically to obtain the transient temperature distribution within the cell

(5) where Bi = hH/k is the dimensionless Biot number. Note that there are two contributions: one steady due to the time averaged heating within the cell; the other oscillatory with frequency 20. Both contributions have a characteristic temperature gradients of order , which increases with increasing electric field and with increasing conductivity (salt concentration). This temperature gradient can induce buoyancy-driven convection as captured by the so-called Boussinesq approximation35

(6) where ρ0 is the fluid density at some reference temperature T0, P is the dynamic pressure, g is the acceleration due to gravity, β is the thermal expansion coefficient, and ν is the kinematic viscosity.

In particular, we consider the case in which gravity acts in the z-direction (parallel to the channel but perpendicular to the field). The characteristic flow velocity can then be approximated by balancing buoyant and viscous forces (last two terms) to obtain

(7)

-6 2 -4 For 500 mM NaCl in water (ν = 1.0 × 10 m /s, β = 2.1 × 10 1/K, ke = 4.2 S/m) in an H = 500 µm

55 channel and subject to an E0 = 400 V/m field, this approximation predicts flow velocities of order

10 µm/s. As the flow velocity depends linearly on the temperature gradient, we anticipate both steady flows and oscillatory flows at frequency 20. Importantly, as these effects are second order in the applied field, they do not contribute to our measurement of the ζ potential, which is first order in the field. Finally, we note that the buoyancy-driven flows are not expected to impact the temperature gradients within the cell; convective transport with velocity v over the channel width

-2 H is slow compared to thermal conduction, vHρcp/k ∼ 10 << 1. This observation justifies our neglect of convection in equation 3 above.

Our putative explanation for particle motions based on field-induced, buoyancy-driven flows is further supported by experimental observations. First, the motion observed at 20 goes linearly with the square of electric field, or electrical power (Figure 3-8 a,b). Second, these effects are only observed at high salt concentrations where the electrical conductivity is large such that Joule heating becomes significant. This observation rules out explanations based on second-order electrokinetic effects such as induced-charge electrophoresis, which should also be present at lower salt concentrations. Third, oscillatory particle velocities at frequency 20 are of the order of magnitude suggested by the above analysis and correlate with steady particle velocities at 00

(Figure 3-8 c,d). This observation suggests that motions at 00 and 20 derive from a common origin - namely, a second-order effect whereby velocities depend on the square of the applied field.

Finally, we note that particle motions at 20 are quite variable from particle to particle, which likely reflects the spatial heterogeneity of flows within the cell.

56

Figure 3-8: (a) and (b) X(20) plotted against electrical power for sPSL and aPSL for the salt concentrations where the 2 non-linear effect was statistically significant (1-5 M). The X(20) peaks are roughly linear with power (E ). (c) and (d)

X(20) plotted against the velocity at zero frequency for sPSL and aPSL in NaCl for the same concentrations, showing that the two are correlated as can be expected from an E2 effect. (e) Phase for each peak at 20 showing that X(20) is in phase with the applied electric field, which is also a sine wave.

57 3.4.2 Zeta potential measurements of sPSL and aPSL

Our experimental zeta potential measurements are summarized in Figure 3-9 for negatively charged, sPSL particles (Figure 3-9 a) and for positively charged, aPSL particles (Figure 3-9 b) dispersed in a series of monolavent electrolytes. The measured zeta potentials of sPSL particles are negative at all salt concentrations and remain finite up to saturated NaCl concentration of 5.4 M (Figure 3-9 a).

Apart from fluctuations due to experimental uncertainty, the zeta potentials are roughly constant for each salt beyond 1 M (Table 1), ranging from ca. −20 mV for KCl to −40 mV for LiCl and comparable to the thermal potential of 푘퐵푇/푒 = 25 mV. The maximum salt concentration of 5.4

M represents the solubility of NaCl but is considerably lower than that of LiCl (13 M). We attempted measurements in saturated LiCl; however, the signal was overwhelmed by noise due to the ca. 13-fold decrease in the ratio / and thereby the particle mobility. A tabulated summary of all of our zeta potential measurements can be found in Appendix 1.

58

Figure 3-9: (a) Zeta potentials of 2.9 m sPSL particles vs. concentration in five different monovalent salts. (b) Zeta potentials of 3.3 m aPSL particles vs. concentration in four different monovalent salts. Only data for high salt concentrations  100 mM are shown; see Figure 3-10 for a larger salt concentration range. The error bars represent 90% confidence intervals. The green curves show the prediction of the GC model assuming different surface charge densities. The solid curves are fits obtained for data in NaCl.

59 For comparison, Figure 3-9 a also shows the zeta potentials predicted by the standard GC model, which relates the zeta potential to the surface charge density 휎 as

휎 2푘푇 푒 = sinh ( ) ≈  , (8) 휀 푒 2푘퐵푇

21 where the approximate equality is appropriate for small zeta potentials ( < 푘퐵푇/푒) . The charge density is estimated to be σ = −9.7 C/cm2 based on independent conductometric titrations36; however, this method is known to overestimate σ as it neglects ion conduction in the Stern layer37,38.

Therefore, we derived an additional estimate of σ = −5.1  0.16 C/cm2 for NaCl by fitting the GC model to the experimental data using nonlinear least squares regression. Only zeta potentials obtained at salt concentrations larger than 10 mM were used in the fitting process, as the assumption of constant surface charge is known to fail at lower concentrations for “hairy” particles39 such as

PSL. Even at 5 M salt concentrations, at which the predicted Debye length −1 = 0.12 nm is smaller than the ions themselves (e.g., 0.37 nm for Na), the continuum GC model provides a reasonable description of the experimental data. GC model fits for aPSL, sPSL and cPSL particles over a larger range of salt concentrations can be found below (Figure 3-10).

60

Figure 3-10 ζ potential plotted against NaCl salt concentration on a logarithmic scale, for each of the 3 particles tested (circles) along with chi-squared fits for each particle (solid lines). The number below each particle name is the best fit value of σ. Table 3-1: Zeta potentials of sPSL particles averaged over 1 M to 5.5 M for each salt. The uncertainties are calculated as root mean square of the 90% confidence intervals for the measurement for each salt concentration, for each salt.

Salt KCl KBr CsCl NaCl LiCl

avg (mV) −18.8  7.0 −29.5  7.4 −28.9  7.6 −30.4  8.0 −40.1  7.4

By contrast to sPSL, zeta potentials of positively charged, aPSL particles exhibit qualitatively different behaviors at high salt concentrations (Figure 3-9 b). For each of the salts investigated

(KCl, KBr, NaCl and LiCl), the zeta potentials decrease to zero at high concentrations despite following closely the predictions of the classic electrokinetic model at low salt concentrations38.

Additionally, for KBr, LiCl, and NaCl, the zeta potentials change sign to become negative at concentrations above ~1M. Comparing the different salts, charge inversion is significantly more pronounced for KBr than KCl, suggesting that Br− counter-ions have higher affinity for the aPSL surface that Cl−. This trend is consistent with the Hofmeister series, which orders ions based on their ability to precipitate proteins40 and correlates with the extent of ion hydration41. Interestingly,

61 the co-ions also appear to influence the degree of charge inversion, with the magnitude of the effect increasing as Li+ > Na+ > K+. These observations indicate that the detailed structure of the EDL at high salt depends on the chemical identity of all ions present in the electrolyte. By contrast, the zeta potentials at low salt depend only on the ionic strength (independent of chemical identity) and perhaps the identity of some potential determining ions21 (e.g., hydroxide).

We do not think that the salt-specific effects could be explained by pH of the salts. The solutions are expected to have pH in the range 5-7 and the amidine group is a strong base (pka ~ 10-11) so that pH changes in the range 5-7 are not expected to vary the ionization state of these groups.

3.4.3 Stout and Khair zeta potential

Our reported zeta potentials are based on the Smoluchowski equation; however, we also considered the more rigorous treatment of Stout and Khair2 that accounts for ion steric effects3 and ion-ion correlations42 since we are in a regime where the Debye length (κ-1) is on order of the Bjerrum length (λB). This treatment accounts for ion steric effects and ion-ion electrostatic correlations. The relevant results from their work are presented in Figure 2 of their paper where the dimensionless mobility, M̃ e = 3ηMe/2ζSK is plotted on the y axis, against κR on the x axis, for different values of the dimensionless correlation length δc = κℓc. Here ζSK is the zeta potential calculated from the Stout and Khair treatment, ℓc is the correlation length, and R is the particle radius.

For our conditions, κR falls between 1578 (100 mM) to 15000 (4 M). For these values of κR, M̃ e has already reached an asymptote and does not depend on κR. The value of δc depends on the choice of ℓc which is bounded such that ℓi ≤ ℓc ≤ λB. Here ℓi is the size of an ion. Through κ and through changes in viscosity and dielectric constant, δc also depends on the salt concentration.

62

From these relationships, we find that at 100 mM, 0:3 ≤ δc ≤ 0:75, and at 4 M, 0:25 ≤ δc ≤ 1. For these values, we find that at 100 mM, 1:1 ≤ ζSK/ζsm ≤ 1:5, and at 4 M, 1:1 ≤ ζSK/ζsm ≤ 2.0.

Using the above analysis we find that the Stout and Khair zeta potential (SK) is consistently larger than the Smoluchowski zeta potential (sm). For example, in 4M NaCl, SK can be up to two times larger than sm depending on the choice of the correlation length. Importantly, we are operating in a regime where the Stout and Khair model predicts a significant mobility correction but no mobility reversal as observed in experiment. We chose to present our results in terms of the Smoluchowski zeta potential rather than SK or the raw mobilities for several reasons. First, the Smoluchowski equation normalizes the mobilities by a concentration-dependent factor / that provides numbers of comparable magnitude for a broad range of salt concentrations. Second, sm is considerable simpler than the Stout and Khair zeta potential and involves no unknown quantities such as the correlation length. Finally, the surface charge density implied by the zeta potential (equation 8) can be compared to that obtained by titration to provide additional information on the arrangement of ions in the EDL. Those readers who prefer to consider the measured mobilities directly can do so

(Figure 3-4).

3.4.4 Monte Carlo simulation of the EDL

63

figure 3-11: (a) schematic illustration of the lattice model of a charged surface in a symmetric electrolyte. (b) electric potential within the simulation cell averaged over the 1 and 2-directions. the solid markers represent the average of 12 independent monte carlo simulations; the solid curves are predictions of the gc model; the x markers denote the zeta potentials. the simulation parameters are charge density, 휎 = 0.0625푒/ℓ2 and bjerrum length, 휆 = 2ℓ within a cell of −3 −3 dimensions 퐿1 = 퐿2 = 8ℓ and 퐿3 = 32ℓ; the salt concentrations are 푛푐 = 0.005ℓ (low salt) and 푛푐 = 0.32ℓ (high salt). (c) zeta potential vs. concentration for different three different  corresponding to conditions similar to experiments on spsl. the markers are results of the mc simulations; solid curves are the gc predictions of equation (3). simulation parameters are the same as in (b) unless stated otherwise. To better interpret our experimental findings, we performed Monte Carlo simulations of the EDL at high salt concentrations using the lattice restricted primitive model (see Methods and figure 3-11 a). Despite its simplicity, this model offers insights as to how the continuum GC model breaks down (or not) at high salt concentrations due to finite ion size and ion-ion correlations. We first consider the case of small surface charge densities (휎 < 4휀푘퐵푇/푒ℓ) characteristic of sPSL particles

64 (figure 3-11 b,c). At low salt concentrations, when the Debye screening length is greater than the ionic radius (휅−1 > ℓ/2), the electric potentials predicted by the lattice model agree well with those of the continuum GC model (figure 3-11 b, low salt). At high salt, however, the lattice model presents a qualitatively different picture of the EDL, whereby an excess of counter-ions accumulates at the surface resulting in damped oscillations in the potential (figure 3-11 b, high salt).

Nevertheless, the zeta potentials predicted by the two models are remarkably similar. Here, the zeta potential is defined – somewhat arbitrarily – as the electric potential at 푥3 = 0.5ℓ, the plane separating the surface charge from the first layer of counter-ions. figure 3-11 c shows these zeta potentials as a function of salt concentration for three different values of the Bjerrum length  spanning the range found in experiment. Notably, there is strong agreement between the predictions of the MC simulations and the continuum GC model – even at concentrations above the critical value where 휅−1 = ℓ/2 (black x’s). Intuitively, one might guess that the zeta potential could never fall below  = 휎ℓ/2휀, which corresponds to all counter-ions located in the single layer closest to the surface. In fact, the over-adsorption of counter-ions in this layer leads to an even smaller zeta potential – one more closely in line with predictions of the GC model,  = 휅휎/휀.

These simulation results provide one possible explanation for the surprising success of the GC model in describing the experimental findings for sPSL particles.

For highly charged surfaces (휎 > 4휀푘푇/푒ℓ) such as that of aPSL particles, the zeta potentials predicted by the lattice model at high salt concentrations are significantly greater than the thermal potential (Figure 3-12 a). Moreover, the electric potential within the plane of the first counter-ion layer shows significant heterogeneity, with peaks and valleys separated by many times the thermal potential (Figure 3-12 b). Consequently, it is reasonable to expect that the ions in this layer will not move freely in the direction parallel to the surface upon application of an external field. These ions may instead form a tightly bound Stern layer that does not contribute to electrophoretic motion.

65

Figure 3-12: (a) Electric potential near a highly charged surface averaged over the 1 and 2-directions. The solid markers represent the average of 12 independent Monte Carlo simulations; the solid curves are predictions of the GC model; the 2 −3 x markers denote the zeta potential. The simulation parameters are 휎 = 0.0188푒/ℓ , 휆 = 4ℓ, and 푛푐 = 0.32ℓ within a cell of dimensions 퐿1 = 퐿2 = 8ℓ and 퐿3 = 32ℓ. (b) Electric potential in the plane 푥3 = ℓ for a particular realization of the surface charge distribution; other parameters correspond to those in (a). (c) Effective zeta potentials 2 vs. concentration for three different  corresponding to conditions similar to experiments on aPSL. The markers are results of the MC simulations; solid curves are the GC predictions of equation (3). Simulation parameters are the same as in (a) unless stated otherwise.

To approximate this scenario, we introduce a second potential denoted 2 and defined as the electric potential at 푥3 = 1.5ℓ, the plane separating the first and second ion layers. In qualitative agreement with experiments on aPSL particles, 2 is comparable to the thermal potential and changes sign upon increasing the salt concentration (Figure 3-12 c). This effect is more pronounced for larger Bjerrum lengths, which corresponds to stronger electrostatic interactions and stronger ion-ion correlations. These results suggest that electrostatic effects alone – neglecting

66 contributions due to solvation and interactions – can provide a suitable mechanism for the inversion of highly charged surfaces at high salt concentrations. At the same time, this simple model cannot explain differences among specific salts (except indirectly via the characteristic ion size, ℓ and the permittivity, 휀).

As noted above, the specific salt effects observed in experiment are largely consistent with expectations based on the Hofmeister series. Poorly hydrated counter-ions are expected to interact more strongly with hydrophobic surfaces such as PSL15,40 thereby reducing the magnitude of the zeta potential. Consequently, the zeta potentials of sPSL particles increase in magnitude as LiCl >

NaCl > KCl at high salt concentrations – in order of increasing ion hydration. Notably, the results for CsCl deviate from this trend. A similar argument can be made to explain the behavior of aPSL particles at high salt. The adsorption of an excess of counter-ions (Cl− or Br−) results in a negatively charged surface to which poorly hydrated co-ions adsorb more strongly (K+ > Na+ > Li+).

Furthermore, we note that charge inversion of positively charged, hydrophobic particles has been observed previously at lower salt concentrations for polyanions43 and for monovalent anions such as ClO4− and SCN−15. The salt concentrations required to induce charge reversal increases as SCN

− 9,15 (200 mM), ClO4− 15 (400 mM), and Br− (1 M, this study) – again consistent with the Hofmeister series.

Given the extreme thinness of the double layer at high salt concentrations, it is remarkable that continuum approximations to the dielectric permittivity and the liquid viscosity are so effective. In both the GC model and the lattice model, it is assumed that the dielectric constant in the EDL is same as that in the bulk, despite inevitable changes in water structure near the solid surface.

Nevertheless, we note that the characteristic electric field within the double layer is too small to induce significant alignment of water molecules therein.44 In relating the electrophoretic slip velocity and the zeta potential, it is assumed that the continuum equations of hydrodynamics are

67 applicable with a constant viscosity equal to that of the bulk liquid. This assumption is supported by previous experiments that reveal that the viscosity of a fluid approaches its bulk value over molecular dimensions45. The continuum hydrodynamic treatment also assumes the validity of the no slip boundary condition at the particle surface despite some evidence to the contrary46. One fortunate consequence of the thin double layer is that theories developed for flat plates – such as the present lattice model – provide an excellent approximation for the relatively large spherical particles studied here. Thus, the Smoluchowski equation should provide results very close to that of the model proposed by O’Brien and White that includes double layer polarization47.

Our observation of negative zeta potentials for positive aPSL particles at high salt can be explained by a combination of two effects: (1) the over-adsorption of counterions at the charged surface, (2) the displacement of the shear plane to enclose some of these counterions. Based on the MC simulations, the first effect is attributed to ion-ion correlations, which are increasingly significant at high salt concentrations. We attribute the second effect to the high surface charge of aPSL particles, which reduces the mobility of the most closely bound counterions. However, differences in the behavior of sPSL and aPSL particles may also result from differences in their “hairiness”39 due to dangling polyelectrolytes at the particle surface. For hairy particles, the shear plane encloses counterion charge even at lower salt concentrations resulting in a discrepancy between surface charge estimates from titration and from zeta potential measurements37,48–50. In our experiments, these discrepancies are more significant for aPSL particles than for sPSL particles. Consequently, the over-adsorption of counterions “behind” the shear plane at high salt concentrations could result in zeta potentials of opposite sign for aPSL (more “hairy”) but not for sPSL (less “hairy”). The

Ohshima model51 describes the electrokinetics of soft (i.e., hairy) particles; however, there are two key difficulties in applying this model to explain our experiments. First, we do not have a way to independently measure the frictional coefficient  of the polyelectrolyte layer, which is necessary

68 for a quantitative comparison. Second, the model relies on a mean-field treatment of that does not allow for charge reversal as observed in our experiments for aPSL particles.

3.5 Summary and Conclusions

We have measured finite electrophoretic mobilities of negatively charged, sulfated PSL and positively charged, amdine-functionalized PSL particles up to saturation in NaCl, KCl and KBr.

Our measurements were made by applying sinusoidal electric fields in a disposable cell using high- speed microscopy to capture the particle motion. As a simple way for normalization of the mobilities, we have presented our results in terms of Smoluchowski zeta potentials. Equipped with zeta potential measurements, we attempted to test the validity of the classical mean field continuum

GC model at high salt concentration, approaching saturation.

At the outset we hypothesized that the continuum picture might break down as the electrostatic screening length, -1 becomes smaller than the hydrated ion size and the Bjerrum length, .

However, experiments indicated that the classical picture – no change in sign of zeta potentials - is still qualitatively valid for sPSL particles. MC simulations reproduced these experimental trends for surfaces with relatively low surface charge density. Interestingly, over-adsorption of counter- ions within the EDL allows the effective screening length to be smaller than the ionic size in agreement with the continuum model.

Experiments also indicate that the more strongly charged aPSL particles undergo charge inversion in KBr, LiCl and NaCl at salt concentrations ~ 1 M. In this case, the MC simulations show that the potential in the first layer of counter-ions is large and heterogenous, suggesting that the layer is likely immobile. By shifting of the plane of shear to enclose the first layer of counter-ions, the model reproduces the inversion of the zeta potential at high salt concentrations.

69

Our experiments indicate that these zeta potentials, while too low to stabilize dispersions (~kBT/e), are large enough to cause electrokinetic transport even in high ionic strength systems, notably geological reservoirs, sea water, and blood. In these systems, spontaneous electric fields may result from salt concentration gradients, which cause diffusiophoresis at charged surfaces17,52. Thus, diffusiophoresis might yet be a useful tool to affect transport in otherwise unreachable places, as long as a concentration gradient exists or can be created.

3.6 Author Contributions

I along with Dr. Darrell Velegol came up with the question and a way to do the measurement. I collaborated with Dr. Charles A. Cartier to create the set up for measurements. I carried out the measurements and analysis. Dr. K. J. M. Bishop designed and carried out the simulations. The manuscript was written through contributions of all authors.

3.7 Copyright Notice

Reprinted (adapted) with permission from (Garg, Astha, et al. "Particle Zeta Potentials Remain

Finite in Saturated Salt Solutions." Langmuir 32.45 (2016): 11837-11844. Copyright (2016)

American Chemical Society.

3.8 Acknowledgements

This work was funded in part by Penn State Materials Research Science and Engineering Centers under National Science Foundation grant DMR-1420620. D.V. would like to acknowledge support from NSF CBET 1603716. C.A.C. acknowledges support from the National Science

70 Foundation under Award CBET-1351704. K.J.M.B. acknowledges support from the Center for

Bioinspired Energy Science, an Energy Frontier Research Center funded by the US Department of Energy, Office of Science, Basic Energy Sciences, under Award DE-SC0000989.

ABBREVIATIONS sPSL, aPSL and cPSL – sulfated, amidine-functionalized and carboxylated Polystyrene Latex respectively, EDL - Electrical Double Layer, GC model - Gouy-Chapman model, PALS – Phase

Angle Light Scattering

3.9 References

(1) Carnie, S. L.; Torrie, G. M. The Statistical Mechanics of the Electrical Double Layer. Adv.

Chem. Phys. 1984, 546, 141–253.

(2) Stout, R. F.; Khair, A. S. A Continuum Approach to Predicting Electrophoretic Mobility

Reversals. J. Fluid Mech. 2014, 752, R1.

(3) Bikerman, J. J. Structure and Capacity of Electrical Double Layer. Philos. Mag. 1942, 33,

384–397.

(4) Storey, B. D.; Bazant, M. Z. Effects of Electrostatic Correlations on Electrokinetic

Phenomena. Phys. Rev. E 2012, 86, 56303.

(5) Quesada-Pérez, M.; González-Tovar, E.; Martín-Molina, A.; Lozada-Cassou, M.; Hidalgo-

Álvarez, R. Overcharging in Colloids: Beyond the Poisson-Boltzmann Approach.

ChemPhysChem 2003, 4, 234–248.

71 (6) Bazant, M.; Kilic, M.; Storey, B.; Ajdari, A. Towards an Understanding of Induced-Charge

Electrokinetics at Large Applied in Concentrated Solutions. Adv. Colloid Interface

Sci. 2009, 152, 48–88.

(7) Israelachvili, J. N. Intermolecular and Surface Forces: Third Edition; Academic Press,

2011.

(8) Quesada-Pérez, M.; Martín-Molina, A.; Galisteo-González, F.; Hidalgo-Álvarez, R.

Electrophoretic Mobility of Model Colloids and Overcharging: Theory and Experiment.

Mol. Phys. 2002, 100, 3029–3039.

(9) Oncsik, T.; Trefalt, G.; Borkovec, M.; Szilágyi, I. Specific Ion Effects on Particle

Aggregation Induced by Monovalent Salts within the Hofmeister Series. Langmuir 2015,

31, 3799–3807.

(10) Semenov, I.; Raafatnia, S.; Sega, M.; Lobaskin, V.; Holm, C.; Kremer, F. Electrophoretic

Mobility and Charge Inversion of a Colloidal Particle Studied by Single-Colloid

Electrophoresis and Molecular Dynamics Simulations. Phys. Rev. E 2013, 87, 22302.

(11) Rowlands, W. N.; O’Brien, R. W.; Hunter, R. J.; Patrick, V. Surface Properties of

Aluminum Hydroxide at High Salt Concentration. J. Colloid Interface Sci. 1997, 188, 325–

335.

(12) Kosmulski, M.; Dahlsten, P. High Ionic Strength Electrokinetics of Clay Minerals. Colloids

Surfaces A Physicochem. Eng. Asp. 2006, 291, 212–218.

(13) Dukhin, A.; Dukhin, S.; Goetz, P. Electrokinetics at High Ionic Strength and Hypothesis of

the Double Layer with Zero Surface Charge. Langmuir 2005, 21, 9990–9997.

72 (14) Kosmulski, M.; Rosenholm, J. B. High Ionic Strength Electrokinetics. Adv. Colloid

Interface Sci. 2004, 112, 93–107.

(15) Calero, C.; Faraudo, J.; Bastos-Gonzalez, D. Interaction of Monovalent Ions with

Hydrophobic and Hydrophilic Colloids : Charge Inversion and Ionic Specificity. J. Am.

Chem. Soc. 2011, 133, 15025–15035.

(16) Vinogradov, J.; Jaafar, M. Z.; Jackson, M. D. Measurement of Streaming Potential Coupling

Coefficient in Sandstones Saturated with Natural and Artificial Brines at High Salinity. J.

Geophys. Res. 2010, 115, B12204.

(17) Velegol, D.; Garg, A.; Guha, R.; Kar, A.; Kumar, M. Origins of Concentration Gradients

for Diffusiophoresis. Soft Matter 2016, 12, 4686–4703.

(18) Guha, R.; Shang, X.; Zydney, A. L.; Velegol, D.; Kumar, M. Diffusiophoresis Contributes

Significantly to Colloidal in Low Salinity Reverse Osmosis Systems. J. Memb. Sci.

2015, 479, 67–76.

(19) Kar, A.; Mceldrew, M.; Stout, R. F.; Mays, B. E.; Khair, A.; Velegol, D.; Gorski, C. A. Self-

Generated Electrokinetic Fluid Flows during Pseudomorphic Mineral Replacement

Reactions. Langmuir 2016, 32, 5233–5240.

(20) Staffeld, P. O.; Quinn, J. A. Diffusion-Induced Banding of Colloid Particles via

Diffusiophoresis 1. Electrolytes. J. Colloid Interface Sci. 1989, 130, 69–87.

(21) Hunter, R. J. Zeta Potential in Colloid Science: Principles and Applications; Academic

Press, 1981.

73 (22) Minor, M.; van der Linde, A. J.; van Leeuwen, H. P.; Lyklema, J. Dynamic Aspects of

Electrophoresis and Electroosmosis : A New Fast Method for Measuring Particle Mobilities.

J. Colloid Interface Sci. 1997, 189, 370–375.

(23) Robinson, R. A.; Stokes, R. H. Electrolyte Solutions; Academic Press, 1959.

(24) Jaqaman, K.; Loerke, D.; Mettlen, M.; Kuwata, H.; Grinstein, S.; Schmid, S. L.; Danuser,

G. Robust Single-Particle Tracking in Live-Cell Time-Lapse Sequences. Nat. Methods 2008,

5, 695–702.

(25) Mao, S.; Duan, Z. The Viscosity of Aqueous Alkali-Chloride Solutions up to 623 K, 1,000

Bar, and High Ionic Strength. Int. J. Thermophys. 2009, 30, 1510–1523.

(26) Isono, T. Density, Viscosity, and Electrolytic Conductivity of Concentrated Aqueous

Electrolyte Solutions at Several Temperatures. J. Chem. Eng. Data 1984, 29, 45–52.

(27) Ozbek, H.; Fair, J. A.; Phillips, S. L. Viscosity of Aqueous Sodium Chloride Solutions from

0-150C. In Americal Chemical Society 29th Sountheast Regional Meeting; Tampa, FL, 1977.

(28) Chen, T.; Hefter, G.; Buchner, R. Dielectric Spectroscopy of Aqueous Solutions of KCl and

CsCl Dielectric Spectroscopy of Aqueous Solutions of KCl and CsCl. J. Phys. Chem. A

2003, 107, 4025–4031.

(29) Chandra, A. Static Dielectric Constant of Aqueous Electrolyte Solutions: Is There Any

Dynamic Contribution? J. Chem. Phys. 2000, 113, 903–905.

(30) Wei, Y.-Z.; Sridhar, S. Dielectric Spectroscopy up to 20 GHz of LiCl/H2O Solutions. J.

Chem. Phys. 1990, 92, 923–928.

74 (31) Wei, Y.-Z.; Chiang, P.; Sridhar, S. Ion Size Effects on the Dynamic and Static Dielectric

Properties of Aqueous Alkali Solutions. J. Chem. Phys. 1992, 96, 4569–4573.

(32) Hasted, J. B.; Ritson, D. M.; Collie, C. H. Dielectric Properties of Aqueous Ionic Solutions.

Parts I and II. J. Chem. Phys. 1948, 16, 1–21.

(33) Dukhin, A. S.; Dukhin, S. S. Aperiodic Capillary Electrophoresis Method Using an

Alternating Current Electric Field for Separation of Macromolecules. Electrophoresis 2005,

26, 2149–2153.

(34) Frenkel, D.; Smit, B. Understanding Molecular Simulation : From Algorithms to

Applications; 2nd ed.; Academic Press: San Diego, 2002.

(35) Deen, W. M. (William M. Analysis of Transport Phenomena; Oxford University Press, 2012.

(36) S37495 Certificate of Analysis for Lot 1129648, 2013.

(37) Zukoski IV, C. F.; Saville, D. A. The Interpretation of Electrokinetic Measurements Using

a Dynamic Model of the Stern Layer. II. Comparisons between Theory and Experiment. J.

Colloid Interface Sci. 1986, 114, 45–53.

(38) Borkovec, M.; Behrens, S. H.; Semmler, M. Observation of the Mobility Maximum

Predicted by Standard Electrokinetic Model for Highly Charged Amidine Latex Particles.

Langmuir 2000, 16, 5209–5212.

(39) Seebergh, J. E.; Berg, J. C. Evidence of a Hairy Layer at the Surface of Polystyrene Latex

Particles. Colloids Surfaces A Physicochem. Eng. Asp. 1995, 100, 139–153.

(40) Parsons, D.; Boström, M.; Lo Nostro, P.; Ninham, B. Hofmeister Effects: Interplay of

75 Hydration, Nonelectrostatic Potentials, and Ion Size. Phys. Chem. Chem. Phys. 2011, 13,

12352–12367.

(41) Kunz, W. Specific Ion Effects in Colloidal and Biological Systems. Curr. Opin. Colloid

Interface Sci. 2010, 15, 34–39.

(42) Bazant, M.; Storey, B.; Kornyshev, A. Double Layer in Ionic Liquids: Overscreening versus

Crowding. Phys. Rev. Lett. 2011.

(43) Gillies, G.; Lin, W.; Borkovec, M. Charging and Aggregation of Positively Charged Latex

Particles in the Presence of Anionic Polyelectrolytes. J. Phys. Chem. B 2007, 111, 8626–

8633.

(44) Yeh, I.-C.; Berkowitz, M. L. Dielectric Constant of Water at High Electric Fields: Molecular

Dynamics Study. J. Chem. Phys. 1999, 110, 7935–7942.

(45) Israelachvili, J. N. Measurement of the Viscosity of Liquids in Very Thin Films. J. Colloid

Interface Sci. 1986, 110, 263–271.

(46) Joly, L.; Ybert, C.; Trizac, E.; Bocquet, L. Hydrodynamics within the Electric Double Layer

on Slipping Surfaces. Phys. Rev. Lett. 2004, 93, 257805.

(47) O’Brien, R. W.; White, L. R. Electrophoretic Mobility of a Spherical Colloidal Particle. J.

Chem. Soc. Faraday Trans. 2 1978, 74, 1607–1626.

(48) Zukoski, C. F. I. Studies of Electrokinetic Phenomena in Suspensions, PhD Thesis,

Princeton University. Princeton, New Jersey, 1984.

(49) Zukoski IV, C. F. An Experimental Test of Electrokinetic Theory Using Measurements of

76 Electrophoretic Mobility and Electrical Conductivity. J. Colloid Interface Sci. 1985, 107,

322–333.

(50) Zukoski IV, C. F.; Saville, D. A. The Interpretation of Electrokinetic Measurements Using

a Dynamic Model of the Stern Layer: I. The Dynamic Model. J. Colloid Interface Sci. 1986,

114, 32–44.

(51) Ohshima, H. Electrophoresis of Soft Particles. Adv. Colloid Interface Sci. 1995, 62, 189–

235.

(52) Anderson, J. L. Colloid Transport by Interfacial Forces. Annu. Rev. Fluid Mech. 1989, 21,

61–99.

77

Chapter 4

Relative Roles of Kinetics, Transport and Thermodynamics in Pseudomorphic Mineral Replacement

4.1 Abstract

Pseudomorphic mineral replacement of KBr crystals in saturated KCl has been studied before extensively, but the reverse reaction has been barely studied. He we report the rate, extent and composition of the replaced layer for the reverse reaction – replacement of KCl crystal in saturated

KBr. We find that full replacement of the crystal can be achieved provided the solution to solid mole ratio is >= 0.2 , contrary to previous reports of the replacement stopping after an hour. This shows that kinetic and transport effects can lead to complete replacement in spite of ‘positive relative volume change’ predicted based on thermodynamic equilibria. By comparing the composition of the replaced solid for KBr in KCl and KCl in KBr, we find that the gradient for KBr in KCl is far gentler than for KCl in KBr, indicating that internal transport is far slower for the latter. Thus, we find that the much slower rate of replacement of the latter than the former may be explained simply by the internal transport limitation. Lastly, we find that mixing does not increase the rate of replacement of KCl in KBr in spite of the transport limitation, because it is the internal and not external transport that is limiting in this case, and because external mixing is not expected to create any additional fluid flow internally over and above fluid diffusion. These findings provide insights into the relative roles of fluid transport, kinetics and thermodynamics of the replacement reaction, moving us one step closer to making a theoretical model for this and other more complicated replacement systems.

78 4.2 Introduction

Pseudomorphic mineral replacement reactions (pMRRs) are mineral replacement reactions where the space occupied by the parent mineral comes to be occupied by a guest material, but the macroscopic shape is preserved. pMRRs have been observed broadly in nature in a variety of low and high solubility systems1,2 ranging from carbonates to chloride. The time scale for these transformations is often several orders of magnitude lower than possible by solid-state diffusion alone. Therefore, the presence of a fluid phase and porosity in the guest phase are considered to be pre-requisites for these reactions. The most widely accepted mechanism for this reaction is that pMRR proceeds through dissolution and precipitation reactions, coupled through the concentration at the fluid-solid interface3. The porosity, created during the process of replacement due to a difference in the volume of material dissolved and precipitated, provides a pathway for the fluid to keep penetrating deeper into the crystal. pMRRs are of interest because they occur in geological processes1,4, have been used as a route to synthesize templated, porous materials5–7, play a role in corrosion processes, and can be used to create porosity in geological reservoirs for enhanced oil recovery. In order to interpret natural pMRRs or to use them for technological advancements, it is desirable to accurately model the complex interplay of thermodynamics, kinetics and fluid transport that lead to pMRRs. In this light, replacement in the KBr-KCl system has been studied extensively as a model system to understand the mechanism of replacement because it happens at a reasonable time-scale (minutes to hours) and the thermodynamics of the system are well-known. However, even though thermodynamic implications have been explored in-depth for pMRR in this model system, much about the role of fluid transport and kinetics remains either unexplored or misunderstood. The objective of this work is to gain insights into the role of internal and external fluid transport, reaction kinetics and thermodynamics by experimentally determining the fluid concentration in the crystal

79 and the rate of replacement under various conditions. This would help us upgrade our understanding of the mechanism of pMRR and take us one step closer to making a quantitative model for this system, and eventually for more complicated systems. It will also help us reconcile sometimes contradictory experimental observations as well as the existing theoretical understanding of dissolution, precipitation and transport.

Here we show that full replacement of the crystal can be achieved with either configuration - KBr crystals in KCl solution, or vice versa in spite of ‘positive relative volume change’ predicted based on thermodynamic equilibria in the latter case. The rate and extent of replacement depends on the molar ratio of KBr to KCl, but not on external mixing.

A typical experiment with the KBr-KCl system is observing a small, transparent KBr crystal in a saturated KCl solution. It is observed that the crystal turns white on the outside, and when the crystal is removed and sectioned, a white rim is observed around a transparent inner core. This replacement reaction proceeds the other way as well – KCl crystal in saturated KBr solution – but

~10 times slower than the former8. Though the former reaction has been studied extensively through a variety of innovative techniques including interferometry9, isotope tracing10 and in-situ optical microscopy11, the latter reaction has been largely ignored. The slow progress of the replacement of KCl crystals in KBr solution has been interpreted as ‘stopping’8 of the reaction after an initial period of fast exchange. This has been attributed to a positive relative volume change (i.e. lack of porosity) calculated based on the assumption that the entire system is in thermodynamic equilibrium. A key theme of this work is that for understanding experiments at realistic time- scales of hours, we must recognize the fact that the system may not be in overall equilibrium, even as local equilibria exist. Therefore, the porosity must be determined by not only thermodynamic factors - the lattice size and amount of the appropriate solid phase that precipitates

- but also the interfacial fluid concentration and threshold supersaturation, that depend on fluid

80 transport and reaction kinetics. Here we gain insights into both the reactions by making quantitative measurements of the solid composition through EDS (Electron Dispersive

Spectroscopy). These are supplemented by measurements of the rate of replacement for various times and solid/solution ratios, along with qualitative information from electron- microscopy images.

4.3 Background

4.3.1 Steps in the replacement process

The preservation of parent shape as well as features at various length scales ranging from nanometers to meters suggests that there must be a robust and broadly applicable mechanism for pMRR reactions. One of the earliest hypothesis was that these reactions occur through a solid- diffusion mechanism, but this hypothesis has been discarded as the rates are much higher than could be obtained through solid dissolution, and because it has been observed that the presence of fluid is essential to the replacement reaction1. The most widely accepted theory now is that the crystallographic information is conveyed through the synchronization of dissolution and precipitation reactions that are coupled through the concentration at the fluid-solid interface3.

Another more complicated hypothesis given by Merino12 et al says that crystallographic information is transferred mechano-chemically through pressure solution and force of crystallization, however this has been contradicted recently by some researchers13.

Let’s take a look at the steps in the replacement of KBr crystal in a saturated KCl solution.

81 1. Dissolution: When a KBr crystal is exposed to a saturated KCl solution, KBr begins to

dissolve at the interface. This is because the saturated KCl solution is not in equilibrium

with solid KBr, meaning that it is under-saturated with respect to KBr14.

2. Precipitation and porosity generation: As soon as dissolution happens, the solution

becomes supersaturated with respect to a mixed KBr,Cl solid which precipitates out on the

surface. It precipitates onto the surface and not else-where because nucleation is easier on

the surface, and because supersaturation is highest close to the surface where dissolution

happens. The precipitate however does not occupy the entire space cleared up by

dissolution. This has been confirmed through experiments and can be understood by

considering that KBr crystal has a lattice size of 6.598 Å while KCl has a lattice constant

of 6.292 Å. Moreover, 1 mol of dissolved KBr may not lead to 1 mol of precipitated KCl,Br.

Rather, this would depend on the fluid composition on the Lippmann diagram (section

4.3.2).

3. Fluid transport: The regions not filled by the precipitate act as pores to hold the fluid and

to further cause replacement mediated by the fluid. However, during the process of

dissolution-precipitation, KBr concentration in the fluid increases due to dissolution and

KCl concentration reduces due to precipitation. Diffusion of Br- and Cl- into and from the

bulk solution tries to counter these concentration gradients. Thus, the interface fluid

concentration is a function of the diffusion rate that depends on the existing fluid path as

well as time and position in the crystal.

If one wanted to model this system, one would be faced with the problem of concentration boundary conditions at the fluid-solid interface within the pores. Since high solubility systems have fast kinetics, the usual practice in such a case is to assume a solution saturated with respect to the solid. But because we have solid KCl and a saturated KBr solution, we can never get to

82 ‘equilibrium’ – a saturated KCl solution, which would require getting rid of all KBr from the

solution. Therefore, the interfacial fluid concentration at the dissolution front will be

somewhere between saturated KCl and saturated KBr on the Lippman curve (explained in the

next section) which shows the concentrations of all solutions that are saturated, with respect to

a certain mixed solid composed only of KBr and KCl. The interfacial concentration would

depend on the kinetics, the geometry of the pore (transport) and on threshold supersaturation,

where the onset of crystallization happens. It would effectively depend on the thickness of the

layer of solid that participates in the equilibrium, as elucidated further by Pollok et al15. Since

the kinetics and extent of supersaturation are not known exactly for these mixed solutions,

we would like to measure the interfacial fluid concentration experimentally.

4.3.2 Thermodynamics - Lippmann Diagrams

KBr and KCl can undergo complete solid solution, meaning that a whole range of solids can precipitate from a saturated solution, spanning from pure KBr to pure KCl. The actual composition of the precipitated solid depends on the solution concentration. The solid-solution equilibria for this system can be represented using a Lippmann diagram16,17 as shown below (Figure 4-1). The y- axis shows the sum of solubility products  = [K+]([Br-] + [Cl-]) where the quantities in square brackets are activities.

83

Figure 4-1: Lippmann diagram for the KCl-KBr system.The y axis is the sum of activity products of KBr and KCl and - the x-axis shows both xKBr (mole fraction of KBr in solid) and Br,aq (aqueous activity fraction of Br ). The plot consists of data generated by using a mathematical formulation of the thermodynamics as explained by Pollok et al15(p 217-220). Mathematica was used to calculate the curves.

- The x-axis indicates both the activity fraction of Br in the aqueous phase (Br,aq) and the mole fraction of KBr in the solid phase (xKBr). The solidus curve corresponds to xKBr and the solutus curve corresponds to the Br,aq. Saturated KCl solution lies on the left on this plot and saturated KBr solution lies at the rightmost point. These are also the points where the solidus and solutus curves meet, indicating that the precipitating solid has the same concentration as the solution at these points, i.e. the pure end members. However, there is another such point called the alyotropic point, analogous to an azeotrope in 2-component vapor liquid equilibria18.

When the  for a solution at a given Br,aq falls below both the curves, the solution is under- saturated and no precipitation takes place. On the other hand, if the solution lies above both the curves, the solution is super-saturated and precipitation of a solid will take place. An important point here is that the solution may be supersaturated with respect to a range of solid compositions

84 on the solidus curve. The thermodynamics, as represented by the Lippmann diagram could be used to determine the solid composition with the highest thermodynamic driving force for precipitation using the concept of ‘maximum supersaturation’14,19. The actual composition of the precipitating solid may be decided not just by the thermodynamics but also the kinetics, often resulting in precipitation of a less stable but kinetically favorable precipitate. However, for a system with fast kinetics as is typical of systems with a high solubility, we might expect the precipitating solid to be the thermodynamically stable one.

In order to actually use the Lippmann diagram to interpret experimental results or predict outcomes, it is convenient to consider the solubility phase diagram developed by Pollok et al15,20 where the experimental solubility data has been used to go from solution activity fraction to solution mole fraction on the x-axis, and from the sum of activity products to the sum of concentrations of Br-

(aq) and Cl- (aq) on the y-axis. This is shown in the figure below (Figure 4-2). The solid and solution concentrations in equilibrium are connected by horizontal tie-lines.

85

Figure 4-2: Solubility phase diagram for the KBr-KCl system. The top curve represents solid concentrations and the bottom cure, corresponding solution concentrations, connected by horizontal tie-lines. The curves were calculated using the mathematical formulation of the thermodynamics as explained by Pollok et al15 (p 217-220).

4.3.3 Porosity generation

A key feature of the mineral replacement reaction is that a porosity is observed in the replaced phase. In fact, it has been found that the presence of porosity is essential to the progress of mineral replacement1. Porosity in the precipitated phase must result from the difference in the volume of the material dissolved vs. the material precipitated. This difference depends on two factors – the molar volume of the precipitating phase, and the moles of material precipitated. Each in-turn depends on the fluid concentration at the interface, and the process thermodynamics and kinetics.

For the case of KCl and KBr, we can use the Lippmann diagram to predict the precipitating phase for a given interfacial fluid concentration. Going a step further, Pollok et al calculated the porosity

86 of the precipitating phase for various molar ratios of KBr to KCl by assuming either an overall system equilibrium, or a strictly local equilibrium only. Their calculations predict that KCl cannot be replaced in KBr solutions due to no porosity, contrary to our finding that even mm scale crystals can be fully replaced. Therefore, we sought to observe the crystal porosity qualitatively using

FESEM under various conditions.

4.4 Questions

Below is a list of the existing assumptions of the replacement process, our questions to delve further into them, and our hypotheses. We will come back to this list towards to end to reconcile what we found. Our goal here was to question to the assumptions that have kept us from gaining a deeper understanding of the replacement process.

4.4.1 Passivation

Existing picture: KCl replacement stops after a period of initial growth due to

‘passivation’ at the surface.

Question: Does a KCl crystal in KBr solution form a passivation layer and stop

replacement always, never, or only under certain conditions? Does the rate and extent

of replacement depend on the molar ratio of KCl in the solid phase to KBr in the

solution phase (solid to solution molar ratio)? If so, what is this dependence?

87 4.4.2 Solution to solid ratios

Existing picture: KCl and KBr replacement extents depend on the solution to solid

ratio. The rate of replacement may depend on this ratio as well.

Question: What is the dependence of rate and extent of replacement on the molar ratio

n n of KCl (KBr) in the solid phase to KBr (KCl) in the solution phase ( Br( aq) or Cl( aq) nKCl(s) nKBr(s)

or simply solution to solid molar ratio)?

4.4.3 Porosity

Existing picture: KCl replacement in KBr is accompanied by a positive relative

volume change, due to which there is no porosity in the parent phase. KBr in KCl is

accompanied by a negative relative volume change due to which there is porosity and

replacement proceeds all the way. Relative volume changes, based on the assumption

of overall system equilibrium can be used to predict actual porosity.

Question: What is the porosity of the rim of replaced KCl crystals vs. the rim of

replaced KBr crystals? Does the porosity of the rim vary with distance from the outside

of the crystal? What does the porosity look like qualitatively?

4.4.4 Fluid transport

Existing picture: If mixing changes the rate of replacement, the reaction is transport

limited. Otherwise the reaction is not transport limited. Since KCl and KBr have a high

solubility, their replacement reactions must be limited by fluid transport.

88 Question: Does the rate of KCl replacement in KBr depend on external mixing? Does

the internal fluid have concentration gradients in the absence of mixing, indicating

limitations on internal fluid transport? Do these internal fluid concentration gradients

persist even in the presence of mixing?

4.4.5 Interfacial concentration

Existing picture: An unknown amount of the solid is dissolved into an unknown amount of solution. The ratio of these amounts results in the precipitation of a certain concentration and amount of solid. The knowledge of these amounts can help us infer the reaction kinetics and transport.

Question: What is the solid composition as a function of position within the crystal? How do these change over time? What can these tell us about the ratio of thickness dissolved to diffusion thickness?

4.5 Methods and Materials

4.5.1 Replacement experiments

Optical grade KBr and KCl crystals (International Crystals) were cut to roughly 3x3x6 mm size

using a blade and a 3D printed tool for sizing. The crystals were then soaked in hexane for 5

minutes to remove organic contaminants and dried using Kim wipes. Saturated solutions of

KBr and KCl were prepared by dissolving excess ACS grade powders (Sigma Aldrich), heating

up the solution to about 50 C with stirring, and then letting it cool back to 25 C. Experiments

were carried out in a water bath on top of a thermostated multi-position stir plate (Ika RT 5)

89

maintained at 26 C. A temperature sensor dipped in the bath indicated that the temperature

was maintained at 25.2  1.2 C. The solutions were equilibrated in the bath for at least 12

hours before an experiment was started. 0.1 g of Sulfo-Rhodamine B sodium salt (SRB) (Sigma

Aldrich) dye powder was added per 100 mL of solution in order to observe the porosity using

confocal microscopy. The replacement was carried out by dropping a crystal into the

equilibrated solution in a 20 mL glass vial for low solid to solution ratios or a 1.7 mL centrifuge

tube for high solid to solution ratios. After the experiment, the crystal was pulled out using a

tweezer and patted dry immediately using kim wipes. The sample was left to dry in air at room

temperature overnight.

푛 For the purpose of solution to solid ratio calculation ( 퐵푟 (푎푞) = moles of Br (aq) divided by 푛퐶푙 (푠)

푛 moles of Cl(s) for KCl in KBr experiments, or 퐶푙 (푎푞) for KBr in KCl experiments), the crystal 푛퐵푟 (푠)

size was assumed to be exactly 3x3x6 mm. It was found that the rim width for KCl in KBr

experiments varied substantially in different trials (~20%), probably due to the slow dynamics

of temperature changes. Therefore, all experiments corresponding to the data points reported

on a single plot were all obtained in a single trial.

4.5.2 Fluorescence optical microscopy

The dried crystals were cut using a razor blade perpendicular to the long end on two sides such

that we obtained a cross-section with a transparent center (unreplaced crystal) and a white outer

replaced layer. This was placed on a glass slide on the stage of a Nikon TE 300 inverted

microscope in fluorescence mode with green light from a mercury lamp corresponding to SRB

(excitation/emission maxima 565/586). The dye penetrated the replaced layer along with the

90 fluid and only the rim appeared bright. Grayscale images of the crystal observed with a 4x

objective were taken using a QImaging CCD camera at a resolution of 1392x1040 pixels which

resulted in a calibration of 1.73 m/pixel. The images were analyzed using Fiji ImageJ to obtain

rim widths. The side with the greatest width was chosen for analysis of each crystal, since there

were non-uniformities resulting from low mass transfer at the side facing the bottom. The

averages and standard deviations reported are all results from a single side for each crystal.

4.5.3 Electron Microscopy

Cross-sections cut above using a razor blade were stuck onto carbon tape and mounted on an

aluminium stub. The samples were sputter-coated with iridium (Emitec) to prevent charging.

There were observed using a Zeiss Sigma Field-Emission Scanning Electron Microscope (Zeiss

SIGMA VP-FESEM) at 5-10 kV with the secondary electrons (SE2) detector at magnifications

of 50x – 250x.

4.5.4 Elemental analysis

Blade-cut crystals were further smoothed using a Leica UC6 ultramicrotome. Smoothed crystals were sputter coated with iridium to prevent charging. An EDS detector was used to capture composition data along with FESEM images on the Zeiss SIGMA VP-FESEM. A working distance of 7.5 mm was maintained for EDS measurements, 15 kV accelerating voltage and 60 mm aperture was used for each measurement which resulted in a dead-time of 30-50%. The data acquisition and processing was handled by Aztec (Oxford) software.

The beam was calibrated thrice before the start of each session by collecting data from a piece of copper tape. K, Br and Cl were standardized using unreplaced optical grade KBr and KCl crystals

91 used for replacement. This was done by collecting spectrum from a smooth surface of pure KBr and KCl. The software also identified up to 10% of C and O but this was specified to be not present in the sample. The standardization data was saved into a calibration file which was used to calculate compositions for the replaced samples.

A count rate of at least 1 million per area selection was obtained for each spectrum reported. EDS spectral data was collected from rectangular selections in smooth regions, avoiding the porosity.

Distance from the edge for each spectrum was calculated as the center of mass for each rectangular region. Image analysis for this was done using ImageJ.

4.6 Results and Discussion

4.6.1 Replacement rate of KCl vs KBr

We carried out a typical replacement experiment starting with both end-members for the same molar ratio of solid to solution, for 30 minutes. FESEM images of replaced cross-sections from this experiment are shown in Figure 4-3. The inner, relatively smooth region in each case shows the unreplaced crystal core while the porous rims show the replaced layers. The overall crystal shape was preserved in each case. As previously reported, the replacement front for KCl is much sharper than for KBr. We measured the rim widths to be 0.15 mm for replacement in KCl and 0.54 mm for replacement in KBr. The measured rim width for KBr is 3.6 times that for KCl, much less than 10 times, as reported previously15.

Also, there appear to be obvious qualitative differences in the texture and porosity of the replaced layer in each case. The replaced layer in KCl appears to be composed of two distinct parts – the outer one is a continuous, porous layer while the inner one looks similar to a column of pebbles

92 thrown in together, with porosity in between individual pebbles. The replaced layer in KBr on the other hand is always continuous with a gradient of porosity, and the pores always grow inward, perpendicular to the crystal boundaries.

Figure 4-3: Electron microscopy images of partially replaced cross sections of (left) KCl crystal in saturated KBr 푛 푛 1 solution for 퐵푟( 푎푞) = 50 and (right) KBr crystal in saturated KCl solution for 퐵푟( 푎푞) = , both at t = 30 mins. The 푛퐾퐶푙(푠) 푛퐾퐶푙(푠) 50 scale bars are both 100 m. We wanted to see if the replacement of KCl was simply slow, or had it actually stopped? Upon letting the reactions run for long times in excess solution (ratio of solution moles to solid moles =

50), we found that a 3x3x6 mm crystal of KCl was replaced completely in less than 18 hours

(between 5 hrs to 18 hrs) while a KBr crystal of the same size was replaced completely in about 1 hour. Hence, contrary to previous work with this system, we found that under favorable conditions

(excess solution) and long times, KCl could indeed be replaced at a mm scale. Since porosity is a pre-requisite for these reactions to proceed, we can conclude that no passivating layer had formed due to re-precipitation of KCl that might prevent fluid access of the internal crystal surface.

93 4.6.2 Varying solution to solid ratios

We hypothesized that our results were different from those of previous researchers due to differences in the solid to solution ratios. While our experiments so far were done under an excess solution condition, previously reported experiments may have been done with solution to solid molar ratio of close to 1, though the conditions have not been reported15. We tested the replacement of both KBr and KCl for different solution to solid ratios. A dye dissolved in the solution during replacement penetrated the crystal and illuminating the rim under fluorescent light. The results from these experiments are summarized in Figure 4-4. Note that we chose 30 mins for KBr crystals and

5 hours for KCl crystals in order to see incomplete replacement in each case. We find that KBr replacement does not depend on the solution to solid ratios for the range tested, but KCl replacement depends strongly on it. KCl replacement reduces drastically with decreasing amount of solution in the system. This observation helps us reconcile our results with those of previous researchers.

Figure 4-4: Rim width vs. solution to solid ratio. Fluorescence optical microscopy images of dried cross-sections of partially replaced KBr (top row) and KCl (bottom row) crystals, at times 30 mins and 5 hours respectively, for varying solution to solid ratio shown at the bottom left corner for each image.

94 Figure 4-5 below shows the modified Lippmann diagram given by Pollok et al – a solubility phase diagram - annotated with the overall mole fraction of Br (XBr,total ) for each of the experiments tried.

The green lines correspond to experiments with KBr crystal in KCl solution ranging from 0.02 

XBr,total  0.5 and the red lines correspond to experiments with KCl crystals in saturated KBr solution ranging from 0.5 XBr,total  0.98. We would like to point out that the same ratio of moles in solution to solid can lead to different outcomes for the KBr and KCl starting crystals (in satd. KCl and KBr solutions respectively) owing to asymmetry of the Lippmann diagram16 (section 4.3.2 ). The most striking example is the case of XBr,total = 0.5 which was testing starting with both end members and obviously leads to different results. We hypothesize that if we carried out a KBr in KCl experiment at an even higher fraction of XBr,total, we might see a dependence on the solution to solid ratios as thermodynamic limitations kick-in.

95

Figure 4-5: Solubility phase diagram for KBr-KCl annotated with the solution to solid ratios teted in experiments. The green lines correspond to KBr crystal in KCl experiments, and red lines to KCl crystal in KBr. n Our next question was – for KCl replacement in low solution to solid ratio ( Br( aq) = 1, bottom nKCl(s) rightmost image in Figure 4-4) – is the replacement just proceeding slowly, or has it actually stopped? We sought to answer this question by measuring the replacement rim width at 0.5 hrs, 2 hrs, 5 hrs, 18 hrs and 30 hrs for high to low solution to solid molar ratios. This data is shown in

Figure 4-6.

96

Figure 4-6: Replaced rim thickness for KCl crystals in saturated KBr solution for various times and solution to solid ratios. Note that the 1.75 mm rim thickness point corresponds to the maximum rim thickness possible since it corresponds to complete replacement. Note that the first time-point corresponds to 30 mins. n Compared to the higher Br( aq) ratios (5-50), both the rate of rim growth and the final end point nKCl(s)

n seems to be lower for the lower ratio ( Br( aq) = 1). This difference does not show at 30 mins, but nKCl(s) shows for all subsequent points. Since the rim width doesn’t change much between 2 hours and 30 hours, it appears that an overall thermodynamic equilibrium has been attained. The case for a thermodynamic equilibrium is also supported by KCl replacement experiments for varying exposed surface areas, keeping the total crystal volume and solution to solid ratios constant (Figure 4-7).

The experiment shows that for a high solution to solid ratio, higher surface area leads to higher mass replaced as the rim width remains constant – this points to a diffusion limited reaction that keeps the length penetrated constant. But for the low solution to solid case, we see that a higher surface area leads to lower rim width in order to keep the mass constant. This suggests that there is a thermodynamic limitation in this case. The equilibrium point can be estimated from Lippmann diagrams following the methodology of Pollok et al15, and can be confirmed through XRD

97 measurements that are out of the scope of this work. However, there still remains the question of whether the replacement is slow because of a ‘closing off’ of the crystal. We will address this question later in this chapter, when we look at FESEM images of internal cross-sections of replaced crystals.

Figure 4-7: Effect of exposed surface area on KCl replacement, while keeping the volume of crystal constant.

n The replacement rate of KCl in KBr depends only weakly on the Br( aq) ratio for values from 5 to nKCl(s)

50. One way this might happen is that for each case, we get the same porosity meaning that we get the same solution concentration (at least approximately). The rate of replacement does depend on time for each case, potentially due to a diffusion limitation. We can estimate a diffusion time based on the rim width L and ion diffusivity, D (t ~ L2/D) which comes out to about 19 minutes - about

100 times the actual time taken for replacement! This suggests either that the reaction has kinetic limitations as well, or the diffusion length is ~10 times higher than the rim width, on account of the tortuosity. Such a high tortuosity, if that is the case, points to a very low porosity medium.

98 4.6.3 Effect of external transport

We also carried out one set of KCl replacement experiments with mixing for the solution excess

n case ( Br( aq) = 50). While we recovered and measured crystals for 0.5, 2 and 5 hours, we could not nKCl(s) recover crystals at 18 hours and 30 hours. The crystals were suspended on a nylon string initially and the crystal fell down either because it was weak, or due to dissolution. For the 3 points measured,

n the rim thicknesses lie between those marked for Br( aq) = 50 and 12.5 in Figure 4-6. Thus, external nKCl(s) mixing does not substantially alter the rate of replacement, at least for the first 5 hours. This is somewhat surprising because dissolution in high solubility systems is generally transport limited.

But the observation is well justified because in this case, it must be the internal transport and not external transport that must be rate limiting. We expect that external mixing will not cause mixing within the internal fluid because of the small length scales. It also provides us a pointer if we need to speed up a replacement reaction that is transport limited – a mixing method that could function even at small length scales, such as electrokinetic fluid flow. Our result seems to be in contradiction with Kar et al21 who showed that KBr replacement in saturated KCl solution increased when the solution was flowed rather than stagnant, but since adding the flow also effectively increased the amount of solution present, their effects may be attributed entirely to a higher solution to solid ratio.

4.6.4 Replaced layer composition for replaced KCl

Previous researchers have hypothesized that the KCl replacement stops when the evolving solution composition going from saturated KBr to more KCl rich reaches a point on the solubility phase diagram where there’s a positive relative volume change. For KCl in saturated KBr solution, Pollok et al calculated this transition from negative to positive relative volume change to be at XBr = 0.4

99 (solid composition). We sought to see if this was indeed that case – had replacement stopped when

XBr reached a value of 0.4 in the replaced layer?

Figure 4-8: Top and middle: Secondary electrons (SE) images of the cross section of a replaced KCl crystal before (top) and after (middle) it was smoothed using ultra-microtomy. The squares in the middle image represent areas from which

100 the spectrum was averaged to obtain each of the plotted points. Bottom: Solid composition of the rim of a partially replaced KCl crystal obtained through quantitative EDS measurements. The dark area to the right is the unreplaced KCl crystal. Solid lines are guide to the eye. In order to test this hypothesis, we performed quantitative EDS measurements on a cross-section of a partly replaced KCl crystal in saturated KBr solution under excess solution conditions (t = 5

n hours, Br( aq) = 50) following the procedure described in section 4.5.4 above (see Methods section). nKCl(s)

Figure 4-8 shows the electron image and corresponding EDS compositions calculated by taking spectra at various positions. The darker area to the right is the unreplaced KCl surface and the EDS data shows no Br here. Also, we get 50% K for each region, as expected from a solid composed entirely of KBr and KCl. Looking close to the crystal edge (left-most points in Figure 4-8), we find that the precipitate is very Br-rich. This makes sense, since a small amount of crystal equilibrates with a large amount of KBr solution outside. However, as we go within the crystal, we see a steep reduction in the Br content for the first 100 m, followed by a region of almost constant composition for the next 300 m. Then the last 100 or so microns again see a steep reduction in Br concentration until we get to pure KCl. We would like to point out that the composition measurements from this analysis suffer from some additional uncertainty due to the drying process

– replacing crystals are simply removed from solution and patted with kim wipes, which might not remove all the fluid within the crystal pore space. The drying of this solution will affect the measured composition from dried crystals.

Comparing the top and middle images in Figure 4-8, we find that the blade-cut crystal before smoothing (top image) reveals additional texture compared to the middle image obtained after smoothing, providing hints as to the way material precipitated within the crystal. The top figure clearly shows the presence of 4 zones from left to right in the partially replaced crystal

1. the leftmost, smooth, continuous zone corresponding to almost pure KBr from the EDS

data,

101 2. then an abrupt change-over to haphazardly arranged smaller crystals creating porosity

among individual crystals that corresponds to the region with very gentle concentration

change seen in the EDS data,

3. then a rough, low porosity region corresponding to the steep concentration gradient in

the EDS data, and

4. lastly, the rightmost smooth, unreplaced KCl crystal.

This visual correspondence to the zones seen in the EDS data is helpful in getting an idea of replaced layer compositions for KCl crystals replaced under a variety of conditions without actually doing quantitative EDS measurements each time. We can use this idea to analyze the images of partly replaced KCl crystals under low and high solution to solid ratios, and with mixing, at 3 different times shown in Figure 4-9.

All the images shown in Figure 4-9 show each of the four zones enumerated above, except the bottom right image. But we often see that the smooth KBr layer often separates out from the replaced crystal during sample preparation, so it is likely that the first layer may have just fallen off in this case during mixing. The presence of all these zones at all times shows that the surface is both texturally and compositionally re-equilibrated as time progresses. Thus, it may be justified to assume that the solution is in equilibrium with the solid at this point, and this information can be used along-with the phase solubility diagram to determine the solution concentration in the crystal at each point, as we do in the next section.

Secondly, it also suggests that during the growth of the rim, it is primarily the middle section that becomes broader, with the innermost and outermost concentrations staying fixed. This is important information for determining the boundary conditions on the solution concentration – the innermost concentration is decided by the excess unreacted KCl solid, while the outermost concentration is set by the excess solution outside the crystal. Also, coming back to a question we

102 had raised in section 4.6.2 while looking at replacement under conditions of low solution to solid ratios – is the replacement slow because of a ‘closing off’ of the crystal? The first column of images in Figure 4-9 clearly show that the first zone is just as porous as in other cases even at 15 hours, So the reaction may have slowed down because of low thermodynamic driving force and not because of closing off of the crystal due to a passivating layer. Moreover, this smooth zone 1 is not the last layer to form, rather it is perhaps the first layer to form.

Figure 4-9: Electron images of blade-cut cross-sections of partly replaced KCl crystals at various solution to solid ratios and times.

4.6.5 Replaced layer composition for KBr

We carried out a quantitative EDS analysis similar to the above for partly replaced KBr crystals as

n well (t = .5 hours, Br( aq) = 50). In sharp contrast to observations for a KCl crystal, we see that the nKCl(s) replaced layer concentration varies smoothly throughout the rim. The gentle slope in solid

103 concentration is accompanied by a continuous reduction in the porosity of the rim. The last point corresponds to unreplaced KBr and that comes abruptly because the replacement front is sharp, as reported previously.

Figure 4-10: (Top) SE2 image of the cross section of a replaced KBr crystal smoothed using ultra-microtomy. Bottom: Solid composition of the rim of a partially replaced KBr crystal obtained through quantitative EDS measurements. The non-porous area to the very right is the unreplaced crystal. The white boxes show regions from which spectral data was averaged to get each point on the plot.

104 4.6.6 Accuracy of the EDS composition measurements

Conventional quantitative EDS analysis on a flat, homogeneous and polished surface can provide accuracies as good as 1-2%22. However, given the compositionally heterogeneous and porous nature of the replaced samples on which the EDS analysis was performed, the measurements inherently suffer from certain inaccuracies. In our samples care was taken to smoothen the surface to a roughness of 1 m or lesser using ultra-microtomy, but the sample is porous rendering the x- rays from lower lying regions susceptible to absorption at elevated regions in the sample. Secondly, the sample likely has variation in composition and porosity in all 3 dimensions, though we expect the composition variation to be continuous rather than abrupt owing to the smoothing effects of solid-solution equilibria and fluid diffusion. While we avoided large pores in selecting surface areas for EDS sampling, the possibility of pores within the interaction volume under the surface is not ruled out.

The region sampled by the electron beam is a 3D interaction volume that depends on the beam energy as well as the atomic weight of the material being probed which can be estimated using

Monte Carlo simulations23. A simpler way of describing the volume is a length scale called the electron range that is the radius of the hemisphere centered at the surface that encompasses 90% of the interaction volume22. An expression for the electron range R is given by22

0.0276 퐴 푅 (푚) = 퐸1.67, 푍0.89 0

3 where A is the atomic weight (g/mol), Z is the atomic number,  is the density (g/cm ), E0 is the beam energy (keV), and R is calculated in micrometers. Using the average values of Z and A for

KCl and KBr along with the respective densities and 15 keV energy we obtain electron ranges of

3.6 m and 2.9 m for KCl and KBr respectively. Thus, this is roughly the penetration depth of the x-rays used for EDS analysis.

105 To get meaningful data in the presence of various sources of error, we obtained each spectrum by averaging the x-rays collected from an area that is approximately 35 m x 50 m and presumably is up to 4 m deep based on the analysis above. This was done in order to get an average value from the area as opposed to sampling a tiny area that may or may not represent the compositional average for that area.

To further verify the consistency of our results, we zoomed in on an area (500 x instead of the usual

250 x magnification) and acquired EDS spectra from three 16x16 m square regions spaced 100

m apart (Figure 4-11). All the squares were approximately the same distance from the edge and were centered on a line parallel to the edge, hence expected to have approximately the same composition. The compositions from the 3 squares yielded 49.4%  0.5% K, 14.9%  0.8% Cl and

35.7%  0.3% Br where the uncertainties noted are standard deviations from the 3 square regions.

The low uncertainties ascertain that the measurements are consistent over the large area chosen for each spectrum, and each measurement is likely a good approximation of the average composition in that region. Additionally, the composition data shown in Figure 4-8 and Figure 4-10 is reasonable in that the composition varies monotonically from high percentage of the guest (KBr for Figure 4-8) at the edge to a low percentage of the guest at the replacement front.

106

Figure 4-11: A zoomed in section of the crystal shown in Figure 4-9 The crystal is oriented perpendicularly to that in Figure 4-9 so that the squares numbered 36-38 lie at the same distance from the replacement front. EDS data was acquired from the regions marked Map Data 36, 37 and 38.

107 4.6.7 Solution composition as a function of distance

Figure 4-12: Circles represent bromide mole fractions in solid phases (considering only Br and Cl, excluding K) as a function of distance from the edge of crystal for replaced KCl and KBr crystals shown in Error! Reference source not found. and Figure 4-9 respectively, and marked as KCl, solid and KBr, solid respectively. Squares represent corresponding solution concentrations calculated using solubility phase diagram by assuming solid-solution local equilibria. All lines are guides to the eye. As discussed above in Section 4.6.4 above, images in Figure 4-9 show that re-equilibration of the replaced layer takes place over time. The replacement is mediated through solution, so its only fair to assume that solution concentrations cause this re-equlibration, and that the solution is in equilibrium with the solid. Therefore, making use of the EDS data above and the solubility phase diagram shown in Figure 4-2, we calculated the corresponding equilibrium solution compositions

108 for both the KCl crystal and KBr crystals. These are shown along with the solid compositions in

Figure 4-12. Note that we do not show corresponding solution compositions for the last solid compositions in each case, because these correspond to pure (or almost pure) crystals not in contact with any solution.

The KCl curves, shown in red show that the solid and solution compositions are close together as in the solubility diagram in Figure 4-2. The KBr curves, shown in blue on the other hand are farther apart, again tracing back to similar behavior on the solubility diagram. Both solutions show the presence of concentration gradients but the gradient is far steeper in the KCl crystal than in the KBr crystal – pointing to the lower porosity and correspondingly lower diffusivity in the KCl crystal.

Also, it is worth noting that close to the outside edge, the KBr crystal has solution that has equal amounts of Cl and Br, while for the KCl crystal, this solution contains almost no KCl. This shows how severe the internal transport limitation is for the KCl crystal, compared to KBr crystal. The absence of KCl near the edge of the KCl crystal provides another reason why mixing does not change the replacement – mixing can only increase the Br content of that solution but is it already at the maximum possible Br. However, for KBr in KCl, we might expect mixing to change the composition of the solid precipitate as it might make the solution concentration close to the edge more Cl rich, and slow down the gradient even further. This would in-turn reduce the porosity gradient and keep the porosity high, making the replacement go even faster for KBr.

The high KBr content in the KCl replacement rim also provides a reason for why KCl rim thickness is so low for low solution to solid ratios, even as the rim width of KBr crystal seems not to be affected by it. Due to high KBr precipitation in the case of KCl, relatively high amounts of KBr are lost during the initial replacement causing it to replace only to a small extent.

Lastly, the abrupt change-over from zone 1 (continuous KBr layer) to zone 2 (haphazard mixed crystals) as enumerated in section 4.6.4 in the rim of replaced KCl crystals suggests that zone 1 has

109 relatively high porosity compared to zone 2 and 3, in direct contradiction to previous works that identify this layer as a low porosity ‘passivating layer’, perhaps due to its closed appearance.

4.7 Conclusions

Above, we have performed many new experiments to better understand the role of fluid transport, thermodynamics and kinetics in the pseudomorphic mineral replacement in the KBr-KCl system.

Data on the rates of replacement with varying solution to solid ratios and mixing, the rim composition from quantitative EDS measurements and qualitative data from electron-microscopy images help us consolidate past knowledge as well as gain a much better understanding of this system. The key conclusions are that KCl crystals can be replaced even on a mm scale in saturated

KBr solutions albeit about an order of magnitude lower than KBr crystals in saturated KCl solutions.

The replacement rate is much slower for KCl due to lower internal porosity but it does not stop due to closing of the porosity, as previously thought. Lastly, both these replacement processes are limited by transport – KCl replacement is limited mainly by internal transport, and KBr replacement is limited by both internal and external transport. Conclusions in more detail on each of the questions listed previously are summarized below.

4.7.1 Passivation

Existing picture: KCl replacement stops after a period of initial growth due to

‘passivation’ at the surface.

Question: Does a KCl crystal in KBr solution form a passivation layer and stop

replacement always, never, or only under certain conditions? Does the rate and extent

110 of replacement depend on the molar ratio of KCl in the solid phase to KBr in the

solution phase (solid to solution molar ratio)? If so, what is this dependence?

New knowledge: What looks like a passivating layer is in-fact a continuous, porous

KBr rich layer that is not responsible for stopping replacement. Previous observations

of stopped KCl replacement may be attributed to low solution to solid ratios.

4.7.2 Solution to solid ratios

Existing picture: KCl and KBr replacement extents depend on the solution to solid

ratio. The rate of replacement may depend on this ratio as well.

Question: What is the dependence of rate and extent of replacement on the molar ratio

n n of KCl (KBr) in the solid phase to KBr (KCl) in the solution phase ( Br( aq) or Cl( aq) nKCl(s) nKBr(s)

or simply solution to solid molar ratio)?

New knowledge: The rate of KCl replacement has a weak dependence on solution to

solid ratios for ratios between 5 to 50, observed for 30 hours. But the extent and

consequently, rate is lower for the ratio = 1. The extent (and presumably rate) of KBr

replacement is invariant with the solution to solid ratio between 1-50. Based on EDS

solid composition data, we think that the extent of KCl replacement is lowered for low

solution amounts (but not so for KBr) because the KCl rim has a high KBr content (but

the KBr rim has relatively lower KCl content). Thus, we think it is a thermodynamic,

not kinetic effect.

111 4.7.3 Porosity

Existing picture: KCl replacement in KBr is accompanied by a positive relative

volume change, due to which there is no porosity in the parent phase. KBr in KCl is

accompanied by a negative relative volume change due to which there is porosity and

replacement proceeds all the way. Relative volume changes, based on the assumption

of overall system equilibrium can be used to predict actual porosity.

Question: What is the porosity of the rim of replaced KCl crystals vs. the rim of

replaced KBr crystals? Does the porosity of the rim vary with distance from the outside

of the crystal? What does the porosity look like qualitatively?

New knowledge: The rim of replaced KBr looks far more porous than the rim of

replaced KCl, but nevertheless, both are porous as evidenced by full replacement in

each case (1.7 mm rim width). The EDS data indicate that KCl concentration close to

the unreplaced crystal boundary corresponds to the region of ‘positive relative volume

change’ as given by Pollok et al15, but replacement still progresses through it. The

porosity varies as a function of distance from the rim in each case. For the rim of a

replaced KBr crystal, the porosity decreases continuously as we go from edge of the

crystal inwards. The porosity for KCl has 3 different zones qualitatively – a continuous

KBr-rich zone, a haphazardly arrange zone of smaller mixed crystals with porosity

between them and lastly a low-porosity zone composed of KCl rich precipitates.

112 4.7.4 Fluid transport

Existing picture: If mixing changes the rate of replacement, the reaction is transport

limited. Otherwise the reaction is not transport limited. Since KCl and KBr have a high

solubility, their replacement reactions must be limited by fluid transport.

Question: Does the rate of KCl replacement in KBr depend on external mixing? Does

the internal fluid have concentration gradients in the absence of mixing, indicating

limitations on internal fluid transport? Do these internal fluid concentration gradients

persist even in the presence of mixing?

New knowledge: The rate of KCl replacement does not change significantly with

mixing. Upon comparing FESEM images of replaced surfaces we see no significant

differences between KCl replaced in excess solution with and without mixing, for

times 0.5, 2 and 5 hours. We do find the KBr-rich outermost layer missing at 15 hours

in the case with mixing, but this can be attributed to falling off of this layer due to

mixing.

Looking at the EDS composition data as well as qualitative porosity images, we

see large solution concentration gradients in the pore space of a partly replaced KCl

crystal which must be a result of transport limitation. However, it is the internal

transport and not external that is rate limiting in this case. In fact, the solution close to

the edge of a replaced KCl crystal is already almost pure KBr, and so mixing cannot

change it further, nor can it create convective flows within the crystal due to the small

length scales. However, for a replacing KBr crystal, the concentration close the edge

has Cl and Br in the ratio of approximately 1:1, so we expect mixing to further speed

up the replacement of KBr. A replacing KBr crystal has relatively high rate of internal

113 transport as evidenced by the much gentler solution concentration gradient, and a

substantial amount of Br in the solution close to the surface.

Additionally, the monotonically varying concentration profile points to diffusional

transport. However, the second last point in Figure 4-8 (close to the replacement front)

seems to suggest an inflection point. This could potentially be because the replacement

front passes through the sampled area resulting in greater compositional heterogeneity.

It could also be because of convective transport in this area close to the rim. Greater

accuracy of EDS measurements, especially zooming in on this area can help resolve

this question.

4.7.5 Interfacial concentration

Existing picture: An unknown amount of the solid is dissolved into an unknown

amount of solution. The ratio of these amounts results in the precipitation of a certain

concentration and amount of solid. The knowledge of these amounts can help us infer

the reaction kinetics and transport.

Question: What is the solid composition as a function of position within the crystal?

How do these change over time? What can these tell us about the ratio of thickness

dissolved to diffusion thickness?

New knowledge: The solid composition changes as a function of distance from the

rim due to internal transport limitation. For a replacing KCl crystal, the solution close

to the replacement front has a very steep concentration gradient and is quite KCl rich

owing to slow internal transport. The concentration gradient depends on the kinetics of

dissolution, internal transport and precipitation. A more accurate value of the

114 concentration may be determined by doing quantitative EDS measurements at a larger

magnification close to the replacement front.

A replacing KBr crystal shows a gentler concentration gradient close to the

replacement front. This must be due to the faster internal transport caused by a higher

porosity, and not because of difference in the kinetics of precipitation and dissolution

because both the systems must be very similar in that respect.

4.8 Contributions

Ibrahim Al’Abri performed the preliminary replacement experiments. Alex Wu and I performed the controlled replacement experiments and image analysis from those reported here. Rest of the work reported here was performed by me and guided by Prof. Darrell Velegol and Prof. Christopher

Gorski.

4.9 Acknowledgements

Funding for this work was provided in part by Halliburton Energy Services, by NSF under the

Active Matter Transport projects and by NSF MRSEC IRG2.

I would like to thank the staff at Materials Characterization Laboratory at Penn State, specifically

Julie Anderson for help with EDS measurements and Eric Golden and Katherine Crespin for help with sample preparation. Missy Hazen and John Cantolina at Huck Institute of Life Sciences at

Penn State are thanked for their help with sample preparation and FESEM/EDS measurements. I am thankful to my committee members, especially Prof. Jim Adair, Prof. Manish Kumar and Prof.

Ali Borhan for their critical examination of the EDS data and their comments to improve upon it.

115 4.10 References

(1) Putnis, A. Mineral Replacement Reactions: From Macroscopic Observations to

Microscopic Mechanisms. Mineral. Mag. 2002, 66, 689–708.

(2) Putnis, A. Mineral Replacement Reactions. Rev. Mineral. Geochemistry 2009, 70, 87–124.

(3) Ruiz-Agudo, E.; Putnis, C. V.; Putnis, A. Coupled Dissolution and Precipitation at Mineral-

Fluid Interfaces. Chem. Geol. 2014, 383, 132–146.

(4) Putnis, a. Mineral Replacement Reactions. Rev. Mineral. Geochemistry 2009, 70, 87–124.

(5) Beberwyck, B. J.; Surendranath, Y.; Alivisatos, A. P. Cation Exchange: A Versatile Tool

for Synthesis. J. Phys. Chem. C 2013, 117, 19759–19770.

(6) Xia, F.; Brugger, J.; Ngothai, Y.; O’Neill, B.; Chen, G.; Pring, A. Three-Dimensional

Ordered Arrays of Zeolite Nanocrystals with Uniform Size and Orientation by a

Pseudomorphic Coupled Dissolution -Reprecipitation Replacement Route. Cryst. Growth

Des. 2009, 9, 4902–4906.

(7) Xia, F.; Zhou, J.; Brugger, J.; Ngothai, Y.; O’Neill, B.; Chen, G.; Pring, A. Novel Route to

Synthesize Complex Metal Sulfides: Hydrothermal Coupled Dissolution-Reprecipitation

Replacement Reactions. Chem. Mater. 2008, 20, 2809–2817.

(8) Pollok, K.; Putnis, C. V.; Putnis, a. Mineral Replacement Reactions in Solid Solution-

Aqueous Solution Systems: Volume Changes, Reactions Paths and End-Points Using the

Example of Model Salt Systems. Am. J. Sci. 2011, 311, 211–236.

(9) Putnis, C. V.; Tsukamoto, K.; Nishimura, Y. Direct Observations of Pseudomorphism:

116 Compositional and Textural Evolution at a Fluid-Solid Interface. Am. Mineral. 2005, 90,

1909–1912.

(10) Putnis, C. V.; Mezger, K. A Mechanism of Mineral Replacement: Isotope Tracing in the

Model System KCl-KBr-H2O. Geochim. Cosmochim. Acta 2004, 68, 2839–2848.

(11) Raufaste, C.; Jamtveit, B.; John, T.; Meakin, P.; Dysthe, D. K. The Mechanism of Porosity

Formation during Solvent-Mediated Phase Transformations. Proc. R. Soc. A 2011, 467,

1408–1426.

(12) Merino, E.; Dewers, T. Implications of Replacement for Reaction-Transport Modeling.

Science (80-. ). 1998, 209, 137–146.

(13) Kondratiuk, P.; Tredak, H.; Ladd, A. J. C.; Szymczak, P. Synchronization of Dissolution

and Precipitation Fronts during Infiltration-Driven Replacement in Porous Rocks. Geophys.

Res. Lett. 2015, 42, 2244–2252.

(14) Astilleros, J. M.; Pina, C. M.; Fernández-Díaz, L.; Putnis, A. Supersaturation Functions in

Binary Solid Solution-Aqueous Solution Systems. Geochim. Cosmochim. Acta 2003, 67,

1601–1608.

(15) Pollok, K.; Putnis, C. V.; Putnis, A. Mineral Replacement Reactions in Solid Solution-

Aqueous Solution Systems: Volume Changes, Reactions Paths and End-Points Using the

Example of Model Salt Systems. Am. J. Sci. 2011, 311, 211–236.

(16) Lippmann, F. Phase Diagrams Depicting Aqueous Solubility of Binary Mineral Systems;

1980; Vol. 139.

117 (17) Prieto, M. Thermodynamics of Solid Solution-Aqueous Solution Systems. Rev. Mineral.

Geochemistry 2009, 70, 47–85.

(18) Themis, M. Fundamentals of Chemical Engineering Thermodynamics; Prentice Hall, 2012.

(19) Prieto M., Putnis A, F.-D. L. Crystallization of Solid Solutions from Aqueous Solutions in

a Porous Medium: Zoning in (Ba, Sr)SO4. Geol. Mag. 1993, 130, 289–299.

(20) Pollok, K. Crystal Growth Patterns in Solid Solution Systems: Case Studies on Oscillatory

Zoning and Mineral Replacement Reactions, 2004.

(21) Kar, A.; Mceldrew, M.; Stout, R. F.; Mays, B. E.; Khair, A.; Velegol, D.; Gorski, C. A. Self-

Generated Electrokinetic Fluid Flows during Pseudomorphic Mineral Replacement

Reactions. Langmuir 2016, 32, 5233–5240.

(22) Goldstein, J. I.; Newbury, D. E.; Echlin, P.; Joy, D. C.; Lyman, C. E.; Lifshin, E.; Sawyer,

L.; Michael, J. R. Scanning Electron Microscopy and X-Ray Microanalysis; 3rd ed.;

Springer Science + Business Media: New York, 2003.

(23) Hovington, P.; Drouin, D.; Gauvin, R. CASINO: A New Monte Carlo Code in C Language

for Electron Beam Interaction -Part I: Description of the Program. Scanning 19967, 19, 1–

14.

118

Chapter 5

Chemical microfracking - Enhancing calcite porosity through pseudomorphic mineral replacement

5.1 Abstract

Hydraulic fracking is a massive but low yield (~5 %) operation to extract oil and gas. A higher yield from existing reservoirs can create huge environmental savings especially since operating a reservoir is a water intensive operation. We propose that fracking yields could be increased by changing the fracking fluid chemistry to create additional porosity chemically through pseudomorphic mineral replacement reactions

(pMRRs). In this paper, we study the extent of replacement and porosity created in calcite in 2 M phosphate solutions, as a function of pH in the range 4 – 8 and at 50 and 80 C, relevant to many oil and gas reservoirs.

While mm scale calcite crystals can be replaced completely at 200 C for pH ~ 8, replacement at the lower temperatures is negligible at that pH. Here we report that a small but measurable extent of replacement (20-

100 m) can be obtained in just 3 days by lowering the pH for lower temperatures. This also results in porosity as hypothesized - at pH 5, we find a 10-fold increase in porosity of 2-4 mm calcite crystals after replacement at 50 C for 6 days giving a porosity of 7.5% measured with mercury porosimetry. Moreover, we find using dye penetration experiments with a laser scanning confocal microscope that the porosity created through replacement can open up access to pre-existing internal fractures or cause their preferential replacement. This makes the porosity even more useful for oil and gas extraction by accessing gas even deeper in the reservoir. The extent of replacement is non-monotonic with pH for both 50 C and 80 C. This complex dependence of the replacement rate on the pH stems from the pH dependence of calcite dissolution rate and calcium phosphate solubility and kinetics.

119 5.2 Objectives

Calcite is a common mineral found in geological reservoirs that contain oil and gas. Typically, fracking – the mechanical fracturing of rocks by pumping high pressure fluid – is used to enhance oil recovery by enhancing the porosity. The focus of the present work is to further enhance this porosity chemically, using pseudomorphic mineral replacement reactions (pMRRs), also called coupled dissolution – reprecipitation

(CDR) reactions, as hypothesized in the schematic of Figure 5-1.

Figure 5-1: Schematic showing how an enhanced porosity from mineral replacement could increase gas yields. Higher the surface area exposed to the replacing fluid, greater the additional porosity generated through replacement.

pMRRs are mineral replacement reactions where the space occupied by a parent mineral comes to be occupied by a guest mineral, such that the external morphology of the parent mineral is preserved1. pMRRs reactions proceed through the mechanism of dissolution, transport and reprecipitation and are characterized by porosity in the replaced phase2–4. Additionally, textural information is also preserved on a broad range of length scales ranging from meters to nanometers1,4, depending upon the degree of coupling of the rates of dissolution and precipitation.

120 Dissolution alone is used in well-acidizing operations5 to clean up scale during oil extraction. This process employs HCl mixed with water to obtain pH in the range 2-7 depending on the formation. Porosity generation for enhanced oil recovery using pMRRs rather than dissolution alone provides two major advantages : (1) CDR reactions would presumably provide a more mechanically stable end product by creating a more controlled porosity and (2) during pMRR, dissolution rate can be enhanced by precipitation6,7 (perhaps through stress of crystallization) which might result in a higher porosity. For enhanced oil recovery, we are interested in open porosity, both the amount of porosity generated and its spatial distribution. The same porosity – the ratio of fluid accessible volume to solid volume – is more valuable when it extends out into the crystal than when all the porosity is close to the surface.

Thus, our objectives are:

1. to find solution composition conditions under which calcite can be replaced in phosphate

solutions under reservoir relevant temperature and pH (50-80 C, pH 5-8).

2. to characterize the increase in porosity so generated in comparison to bare crystals.

We approach both our objectives using mm scale calcite crystals as a proxy for the surface of calcite rocks in a geological reservoir. To find the conditions necessary to get porosity at these low temperatures, it would be ideal to have a comprehensive model that accounts for all the factors that go into replacement

– thermodynamics, kinetics and transport. However, the process of mineral replacement is poorly understood, and the kinetics of calcite dissolution, as well as calcium phosphate precipitation are quite complicated. Therefore, in this work, we set out to map unexplored territory of these reactions under conditions that we are interested in.

Previous studies of this replacement process carried out using 2 M phosphate at pH ~ 8 found no replacement at T <= 80 C8. It was also suggested previously that the replacement process at T >= 120 C was dissolution limited8. Based on these, we hypothesized that lowering the pH to increase the dissolution rate may permit the replacement reaction to proceed even at lower temperatures. Therefore, we investigated

121 the extent of replacement of calcite in phosphate solutions in the pH range 4-8 at 50 and 80 C. Here we find that calcite can indeed be partially replaced at 50 C between pH 4-5, and at 80 C between pH 5-7 in

2 M phosphate solutions. We further studied the effect of mixing, phosphate concentration, time and the addition of additional Ca+2 ions in solution on the extent of replacement and textures observed, using electron microscopy.

To characterize the porosity, we have used mercury porosimetry to measure the pore volume and estimate their size. Additionally, we soaked replaced crystals in a dye solution which penetrated into the pore spaces inside the crystal through diffusion. These crystals, probed in all 3 dimensions through confocal microscopy, revealed that after 12 days of replacement the newly generated pores connected to pre-existing fractures within the crystal showing that the porous channels went deep into the crystal.

5.3 Background

5.3.1 pMRR of calcium carbonate

Most previous studies of pMRR reactions of calcium carbonate in phosphate solutions report replacement primarily under hydrothermal (high T and P) conditions (T  120 C)8–11 barring 2 studies that studied surface changes only with nacre powders (a biomineral composed primarily of aragonite). All these studies are summarized in Table 5-1: Summary of previous and current work done on dissolution- precipitation of calcium carbonate in phosphate solutions. A summary of other work done at even higher temperatures can be found in Yoshimura et al11.Table 5-1. In the study by Yoshimura et al11, the replacement of a calcite

 powder (<10 µm in size) was carried out at T = 120 C in 60 mM H3PO4 solution at pH 7 for 6 hours. Their testing of the reaction at 80 C resulted in no reported changes. Putnis and coworkers8–10 on the other hand, carried out the replacement of calcite single crystals (~ 1 mm in size) in much higher phosphate

122

  concentrations (2 M (NH4)2HPO4), at 120 C > T > 200 C, at pH close to neutral (pH 7.8) for 1 week to 2 months. Both the studies report that no replacement was observed at 80 C – the upper limit of temperatures we are interested in.

Table 5-1: Summary of previous and current work done on dissolution- precipitation of calcium carbonate in phosphate solutions. A summary of other work done at even higher temperatures can be found in Yoshimura et al11. pH time T (C) Calcite form Ref ~8 2-21 180, 200 ~ 1 mm size crystals of Iceland spar and carrara Jonas et al8 days marble, focusing on internal cross-sections (2013) 7.7- 2hr-4 120-160 Calcite powders size ~100 m, ~ 1 mm size Iceland Kasioptas et al10 8.06 days spar, focusing on internal cross-sections and porosity (2011) ~7 6-30 120, 180 Calcite powders size ~10 m, complete conversion Yoshimura et hrs reported al11 (2004) 6.0- 1-12 37 Nacre powders, only outside surfaces observed Guo et al12 8.0 days (2007) 7.4 5-14 25 Nacre powders, only outside surfaces observed Ni et al13 (2003) days 4.0- 1-12 50, 80 ~ 1 mm size crystals of Iceland spar, focusing on Present work 8.0 days internal cross-sections and porosity

5.3.2 Porosity measurement

14 Recently, Pedrosa et al estimated the open porosity of calcite crystals fully replaced by fluorite (CaF2) using BET surface area. BET measurements have two drawbacks for our purposes here. First, it does not give any information about the length of channels, and second, it operates in a limited pore diameter range

123 of ~0.3 – 300 nm15. Pedrosa et al14 also supplemented BET measurements with calculating the total porosity based on the analysis of electron microscopy (FESEM) images. FESEM/EDS gives textural and compositional information at a high resolution but it does not differentiate between open and closed porosity.

Moreover, the analysis of surface images to get porosity works for a homogeneously replaced crystal but we find that our system has large scale heterogeneities even at the end of 12 days.

5.4 Methods and Materials

5.4.1 Crystals

Natural calcite crystals (Iceland spar) were obtained from American Educational Products LLC

(Figure 5-2). Smaller crystals for replacement reactions were prepared in one of 2 ways – fracturing with a hammer or cutting with a saw. Fracturing was done by lightly beating a large crystal placed in a plastic bag using a hammer, causing it to fracture naturally at the cleavage planes (Figure 5-2). The broken crystals were sorted by sieving with stainless steel sieves. The experiments were carried out using crystals that passed through the 4 mm sieve but not throughs the 2 mm sieve, resulting in crystals of cross-section 2-4 mm, and the third dimension > 1 mm (Fig S1). Saw cutting was done using an automated dicer (Accutom

50) fitted with a diamond saw that cut the crystals into 3x3x3 mm cubes. The crystals were sonicated in DI water for 2 minutes, and then rinsed using de-ionized (DI) water (Millipore Corporation Milli-Q system, specific resistance = 18 MΩ cm, pH ~5.5 in equilibrium with atmospheric CO2). The crystals were dried by placing on kim wipes to absorb the water.

124

Figure 5-2: Iceland spar calcite crystal (left), calcite crystals after hammering, sieving and washing (middle), sample S01 (see Table 5-3) calcite crystals after replacement for 6 days (right).

5.4.2 Buffer solutions

Powdered, ACS grade chemicals (Sigma Aldrich/Acros Organics) and de-ionized (DI) water were used to prepare the buffers summarized in Table 5-2. The DI water had pH ~5.5, in equilibrium with atmospheric carbon dioxide.

The pH was measured using a Hach H170 pH and conductivity meter a few hours after buffer preparation.

The ammonium phosphate buffers (B11-14) were especially slow to equilibrate. The pH was also measured after the replacement for buffers B11-16, and was found to be within 0.1 pH unit for each case, indicating that the buffering action was adequate.

Table 5-2: Buffer compositions used for replacement experiments. * mono refers to ammonium phosphate monobasic as the source of phosphate.

Phosphate Sno pH Acid type Acid (M) Base type Base (M) (M)

B16 4.18 0.82 Sodium 0.18 2 (mono*) Acetic B15 5.06 0.30 Acetate 0.71 2 (mono*) B11 5.27 1.76 0.2426 2 B14 6.49 Ammonium 1.06 Ammonium 0.944 2 Phosphate Phosphate B12 6.73 Monobasic 0.840 Dibasic 1.160 2 B13 7.97 0.135 1.87 2

125

P1 8.11 - 0 2.0 2.0 P2 5.02 1.66 Ammonium 2.2 2.2 Hydrochloric Phosphate P3 4.98 0.57 0.5 0.5 acid Dibasic P4 5.07 0.137 0.133 0.133

5.4.3 Replacement experiments

Experiments to probe the effect of pH on replacement were carried out using the buffers B11-B16 (Table

5-2). 3 saw-cut calcite crystals (~0.06 g each) were placed in a 15 mL centrifuge tube soaked in 3 mL of the buffer, and lowered into a thermostated silicone bath at the desired temperature for 3 days.

Further experiments at 50 C probing the effect of time, phosphate concentration, mixing and Ca+2 concentration, were carried out with buffers P1-P4 (Table 5-2). These experiments were carried out with 4 g of washed and dried 2-4 mm calcite crystals in 60 mL solution, inside 80 mL polypropylene containers.

For experiments without mixing, the crystals were placed in an even layer on a polypropylene mesh fabric

(mesh size 600 micron, manufactured for water filtration by The Cary Company) and were covered completely with solution. For experiments with mixing, the polypropylene mesh containing the crystals was sealed using a heat sealer (Figure 5-3). The bag containing the crystals was placed about an inch from the bottom of the container, to provide clearance for a magnetic stirrer. The bag stayed jammed in its place due to the stiffness of the fabric. Again, the solution completely covered the crystals.

Table 5-3: Conditions for experiments reported. All experiments were carried out with fractured crystals. Crysta Solutio Soak Temperatur Original S (NH ) HP l n Buffer time 4 2 pH Mixing e sample No O4 conc weigh volume (days) (C) number t S01 P1 6 2.0 M 8.11 0 0.202 5 mL 200 S200 S02 P1 6 2.0 M 8.11 0 4 g 60 mL 50 S62

126

S03 P2 1-12 2.2 M 5.02 0 4 g 60 mL 50 S104(6), S76 500 4 g 60 mL 50 S04 P2 6, 12 2.2 M 5.02 S77 rpm S05 P3 6 0.54 M 4.98 0 4 g 60 mL 50 S89 500 4 g 60 mL 50 S06 P3 6 0.54 M 4.98 S103 rpm S07 B12 6 2.0 M 6.73 0 4 g 60 mL 80 N23 S08 P1 6 2.0 M 8.11 300 0.2 g 50 mL 200 S23 rpm 500 0.5 g 60 mL 50 S09 P2 6 2.2 M 5.11 S95 rpm S10 P2 6 2.2 M 5.11 500 15 g 210 50 rpm (5g x mL S90 3 bags)

Figure 5-3: (Left) Multi position heater-stirrer used for the replacement experiments, (Middle) calcite crystals in the reaction vessel, placed on a polypropylene net have turned white after replacement, (Right) calcite crystals in a heat-sealed polypropylene net bag for replacement experiment with mixing. The crystals have turned white and joined together due to extra precipitation. To see how deep the replacement rim could proceed in buffer B2, we carried out an experiment with time where 2 crystals were withdrawn from the same bottle every few days. Since the total weight of crystals

127 was much larger than the weight of 2 crystals, we expect that the surface area exposed did not change significantly throughout the experiment.

Multiple reactions without mixing were carried out simultaneously by placing them in a thermostated silicone oil bath ( 3 C), that allowed for 1 reactor to be mixed in the center using a magnetic stir bar.

Further mixing experiments were carried out on a multi-position hot-plate stirrer (Figure 5-3) set to a temperature such that 50 C was maintained within the reaction containers. As a control, we carried out the same experiment in both the set ups and obtained good agreement. After the experiment, the crystals were rinsed with DI water 3 times, patted dry using kim wipes and left to dry at room temperature overnight.

2 experiments were carried out at 200 C and 30.6 bar using a high P-T Parr 50 mL stainless steel reactor with a Teflon liner (Figure 5-4) to reproduce results from previous researchers. 5 mL of buffer P1 was added along with 3 saw-cut crystals and reacted for 6 days in the reactor. One of the experiments was carried out with overhead stirring by a stainless-steel stirring-rod in the reactor rotated at 300 rpm during the experiment at 200 C.

Figure 5-4: Parr High P and T cell we used for calcite replacement. The vessel can fill upto 50 mL and can go upto 5000 psi and 350 0C. The reactor was pressurized with air.

128 5.4.4 Electron Microscopy and Elemental Mapping

Replaced crystals were characterized using a field emission scanning electron microscope (FESEM). The crystals were cleaved using a carbon steel blade, by applying pressure parallel to the cleavage plane. The results were independent of the axis chosen. The cleaved crystals were placed on double sided carbon tape which was glued to an aluminium stub. The crystals were iridium coated using a Quorom EMS 150 Sputter

Coater and imaged under a Zeiss SIGMA VP-FESEM (accelerating voltage = 5-8 kV, SE2 lens). Electron dispersive spectroscopy (EDS or EDX) for elemental analysis was also carried out with the same machine

(accelerating voltage = 10 kV), using an EDS detector fitted on the microscope.

5.4.5 Image processing for rim width measurement

Rim width was measured from FESEM images. For each sample, the rim width varied within a crystal for different sides because some sides were placed facing another side, and hence did not have good surface exposure to the solution. There was no way to pick these sides apart before electron microscopy, so we chose the side with the highest replacement thickness for image analysis from each experiment.

Additionally, it was found that the rim was uneven in thickness even within a single image. Therefore, instead of making a few thickness measurements within the image, we measured its thickness throughout the image using Fiji ImageJ.

The raw images were rotated with bilinear interpolation such that the direction of rim growth appeared vertical in the image. Using the wand tool, the uneven rim was selected from the image and the rest was deleted. Next, the image was smoothed and thresholded, resulting in a binary image. The image was saved as a text file that listed a value of either 0 or 255 for each pixel of the image. The rim width for each column was calculated by counting the number of pixels with a value of 0. This list of widths was used to obtain

129 the average and standard deviation for each experiment. The pixel size was calibrated using the scale bar embedded into the image by the image acquisition software and the actual rim widths were reported.

5.4.6 Mercury porosimetry

Porosimetry was carried out using a MicroActive AutoPore V 9600 1.02 system (Micromeritics) by

Materials Characterization Laboratory staff at Penn State. Since mercury is non-wetting, it enters pores only upon application of an external pressure. The pressure P required to force it inside depends on the pore diameter D (assuming a cylindrical pore) as 퐷 = −4 cos  /푃 where  is the surface tension,  is the contact angle (130 ) and P is the applied pressure.

Pore diameters are reported as equivalent cylindrical pores. No correction was applied, which might result in an error of 1% of the stem volume used (for 31 % stem volume used, and 7.7% reported porosity, the actual porosity may be between 7.45% to 7.94%). The measurement was done in set-time equilibration mode with a hold of 10s for each pressure. Only the intrusion curve was recorded and used for analysis.

Void volumes were corrected for and are not included in the porosity data. The diameters reported are those corresponding to the neck of the pore even if the actual diameter may be larger (ink bottle effect).

Micromeritics recommends that at least 25 % of the stem volume must be used. This was done for all the reported intrusion curves for replaced calcite samples (25 – 60 % of the stem volume was used) but for the unreplaced calcite (control) runs, only 9 % and 11 % of the stem volumes was used in 2 runs, resulting in porosities of 1.5 % and 0.7 % respectively. Only the second result was used because it was done with a much larger calcite weight.

130 5.4.7 Dye penetration using confocal microscopy

0.2 g Rhodamin B dye powder (Sigma Aldrich) was dissolved in 50 mL DI water to prepare a stock solution.

A 10 times diluted solution of the stock was used to soak a replaced calcite crystal in a petridish sealed using parafilm. After soaking for 24 hours, the crystals were padded dry using kim wipes and then left to air dry for an hour. Since replaced crystals are opaque from the outside, they were cleaved on two opposite sides to reveal a transparent center with a white rim. The crystals were imaged using a Leica TCS SP5 laser scanning confocal microscope (LSCM, Leica Microsystems) with illumination from a HeNe543 laser, a 5× objective, a 70.77 μm pinhole and emission recorded in the range 580 nm – 650 nm. Each crystal was imaged at 15 – 30 horizontal planes depending upon its thickness, with a z step size of 20.01 µm. Each of the images shown in Figure 5-17 is a z projection of the entire z-stack for each crystal which shows the maximum intensity recorded for each pixel. The images are a visualization of the maximum dye penetration observed at each position without any regard to the thickness of that particular channel.

5.4.8 Calculations using VMinteq

Calcium concentrations at equilibrium at 50 C and 80 C were calculated with ionic strength set to 2 M, and pH varied using the sweep function. To simulate equilibrium, an infinite solid was specified and equilibrium was allowed to be reached. The Specific ion interaction theory (SIT) model as described by Grenthe et al16 was used for activity correction. However, it should be pointed out that the actual system has higher ionic strengths of > 6 M in some cases which is beyond the range (up to 4 M) of the activity model used. Thus, the calculations are meant to give only rough estimates of what we might expect.

131 5.5 Results and discussion

5.5.1 Effect of temperature on replacement at pH 8.11, 200 C vs. 50 C

As a starting point of our investigation, we tested the replacement of calcite crystals prepared by hammering

(method 1) in 2 M ammonium phosphate dibasic (buffer P1 from Table 5-2, pH 8.11) at 200 C, as demonstrated previously by Putnis and others10. We found that after soaking the crystal for 6 days, the initially transparent crystals turned white but retained their shape as previously reported. FESEM imaging of a cleaved cross-section revealed a thick, porous replacement rim extending into the crystal. In fact most of the mm scale crystal was replaced, with only a small ~0.2 mm region in the center unreplaced. Next we carried out the same experiment at 50 C. In this case, we macroscopically observed that the crystals became slightly translucent, but stayed smooth. FESEM imaging of the cross-section revealed that the boundary had been disturbed by the solution somewhat but no clear rim was observed. These 2 experiments are summarized in Figure 5-5.

Figure 5-5: FESEM images of calcite crystals soaked in 2 M phosphate (buffer P1, pH 8.11) for 6 days at (a) 50 C and (b) 200 C To understand why replacement is observed at 200 C but not at 50 C, let us look at the temperature dependence of the underlying processes – dissolution, transport and precipitation. For replacement to begin,

132 calcite must dissolve and a calcium phosphate phase should nucleate on the dissolved surface. For replacement to continue, the fluid should continue to be in contact with the dissolution front by means of a porous replacement layer. The image at 50 C shows that replacement had begun and perhaps even continued, but very slowly.

Unlike most minerals, the solubility of both calcite and calcium phosphates goes down with increasing temperature. However, the change in solubility is small, and the solubility stays within the same order of magnitude for temperatures in the range 50 – 200C. However, in spite of the less favorable thermodynamics, the dissolution rate increases with temperature17. For example, the dissolution rate constant of Iceland spar at pH 8.4 in 0.7 M KCl solutions goes from 0.1 cm/s at 84 C to 0.001 cm/s at 2

C. As temperature goes from 200 C to 50 C, all 3 processes slow down, but dissolution and precipitation slow down far more than transport18. Thus, one plausible hypothesis for the lack of replacement at 50 C at pH 8.11 is that the dissolution and/or precipitation is too slow.

However, an explanation based on transport as the rate controlling step is not ruled out. This is because transport is controlled by the porosity of the replaced phase that depends on the molar volume and solubility of the precipitated phase19. Both of these quantities are affected by temperature and to the best of our knowledge, this dependence is not known for temperatures greater than 37 C. A supporting fact for this hypothesis is that qualitative changes in porosity of the replaced phase have been observed by Jonas et al8 for this system, for temperature change from 160 C to 140 C.

5.5.2 Effect of pH on observed replacement

Previous studies have conjectured that dissolution rate is the rate limiting step for this replacement at higher temperatures10,20. Since the rate of calcite dissolution increases with pH for pH<5.521, we hypothesized that

133 reducing the pH could speed up the rate limiting dissolution step at low temperatures. Accordingly, we prepared a range of buffers with pH from 4-8 to test this hypothesis.

Figure 5-6 shows FESEM images of crystals reacted for 3 days in the various buffers listed in Table 5-2 at

50 and 80 C. The images reveal that a distinct and measurable replacement rim can be obtained under all conditions except at pH 8 at 50 C. The rim thickness, waviness and thickness of overgrowth vary with both temperature and pH indicating that the replacement behavior is a complex function of the pH.

In this study, we have used the rim thickness as a simple proxy for extent of replacement normalized by the exposed surface area since the rim width is effectively the ratio of replaced volume and exposed surface area. In contrast, previous studies use mass % converted which depends on the exposed surface area, and renders is difficult to compare data from powder runs vs. larger crystals. However, rim width is an adequate metric only if the system is limited by kinetics which is the case for our system, and not thermodynamics.

Figure 5-6: Rim width vs. pH at 50 and 80 C. FESEM images of cross-sections of calcite crystals after replacement for 3 days in various buffers described in Table 1. Regions of greater rim width in a wavy rim perhaps indicates that surface reaction is faster than transport in those areas, resulting in a ‘replacement instability’. Given that the rim is particularly wavy at pH 4, and

134 quite straight at pH 6.49, it is reasonable to conclude that the waviness is not simply a result of crystal defects which should be present under both the conditions tested. The waviness is most likely a result of a faster surface dissolution rate than transport. This explanation also allows for the non-monotonic variation in waviness with pH observed here since both the rates vary with pH – the surface dissolution rate generally decreases with pH while the transport rate depends on the porosity that depends on pH through the molar volume of the phosphate species precipitated.

Figure 5-7 shows a plot of rim width vs. pH for these crystals, using the method outlined in section 5.4.5.

The error bars are 1 standard deviation and serve as an indicator of the waviness. The rim thickness is non- monotonic with pH for both the temperatures tested unlike the expected dissolution rate of calcite, as discussed below. This indicates that the rate controlling step is not dissolution throughout the pH range.

Perhaps the replacement rate controlling step changes as a function of pH, and it could be either of dissolution, nucleation, precipitation or transport for each case.

Figure 5-7: Mean and maximum rim thickness observed for calcite crystals as a function of pH, in 3 days.

135 5.5.2.1 Effect on dissolution rate with pH pH affects both dissolution and precipitation in complex ways. The solubility of calcite increases substantially with reducing pH. For example, the plot in Figure 5-8 shows calcium concentrations at saturation as a function of pH obtained using Visual Minteq. Notice that the y axis is on a log scale and that the Ca+2 concentration goes from ~1 M at pH 4 to ~2 mM at pH 8 at 80 C with 2 M phosphate. The solubility is affected far more by pH and ionic strength than by temperature, other things being constant.

Figure 5-8: Calcite solubility with pH using Visual Minteq. The simulation was done using a pH sweep at fixed temperature and total phosphate concentration, along with calcite specified as an infinite solid allowed to reach a saturation in solution. However, the dissolution rate does not always correlate directly with the solubility. In extensive controlled studies of the dissolution rate of calcite with pH by Sjoberg and coworkers21–23 and more recently by

Dolgaleva et al24, 3 regimes have been identified for the dissolution rate of calcite with pH.

136

1. pH dependent regime with transport control only. pH < 4 at 25 C.

2. Complex H+ dependence regime. 4

dependence varies with pH, hence complex.

3. Mixed kinetics regime. pH >5.5. The rate depends on both kinetic rates and transport of the

products from the diffusion boundary layer, but not on the pH.

All three regimes are displayed for a wide range of temperatures23 (tested from 1 – 62 C) but the threshold pHs listed above shift towards lower pH. Thus at 50 and 80 C, we might well be in a regime where the dissolution rate is independent of pH for most of the regime barring perhaps only pH = 4. However, we would like to point out that the mechanism, rate and solubility of calcite dissolution is not fully reconciled yet. The present mechanism is thought to be mixed kinetics21 for the third regime. This would predict a rate directly proportional to the solubility, which we saw depends on the pH, so it is not clear how the rate stays independent of pH in the third regime.

5.5.2.2 Effect on precipitation rate with pH

The precipitation step is affected by pH, since the pH controls the speciation of phosphate (Figure

5-9). This in-turn affects the phase diagram25 of calcium phosphate that dictates the most stable phase under each condition. Also, the overall solubility of calcium phosphates increases with reducing pH. Owing to the complexity in the precipitation of calcium phosphates, the exact kinetics are not well-known but two factors must come into play into the precipitation reaction:

1. Increased solubility means a lower saturation index (SI). SI controls mode of growth - island

nucleation versus incorporation at existing steps18.

2. The least stable phase generally precipitates first (Ostwald ripening)25. The least and most stable

phase changes with pH in accordance with the speciation of phosphate25.

137

Figure 5-9: Speciation diagram of phosphate plotted by doing a pH sweep in Visual Minteq with 2 M phosphate at 50 C. The plot at 80 C (not shown here) shifts by less than half a pH point.

5.5.2.3 Interdependence of rates

Lastly, the fluid transport is affected by the porosity which is in-turn controlled by both dissolution and precipitation rates as well as on the fluid composition each of which depends on the pH. Mixed kinetic of the dissolution rate in the third regime dictates that the surface concentration of Ca+2 ions will increase with increasing solubility with reducing pH. Since precipitation happens at the surface, this would change the supersaturation level for calcium phosphate precipitation. Additionally, precipitation of dissolved products serves to remove the product out of solution, thus further speeding up the dissolution process. These are

138 some of the reasons why we might have a complex, non-monotonic dependence of the replacement behavior with pH.

5.5.3 Replacement at pH 5 at 50 C

Given the complex behavior with pH, we chose to investigate one relevant case in more detail, as a function of time, mixing and phosphate concentrations. For this, we chose the condition of pH 5 at 50 C. Instead of the acetate-buffer we used in the experiment above for pH 5, we added HCl to the solution to get the pH down to 5 (solution P2), since HCl is a common acid already in use for acidization of wells. Whereas in acidization, the pH may be as low as pH 1-2 resulting in plain dissolution which might hollow out the reservoir and collapse it, here the pH is much higher. It should be pointed out that in our HCl-phosphate solutions, the solution starts out at pH 5 but the pH could change over the course of the experiment because the solution is not buffered at pH 5. That said, we found that qualitative features of the replaced rim and the average rim width were the same for HCl-phosphate solution at pH 5 as well as acetate-buffered phosphate solution at pH 5.

With this solution, we found that the calcite crystals turned white on the outside within a day at 50 C. A replacement rim could be identified from its relative roughness compared to the smooth unreacted calcite surface. The rim width at 3 days was very close to that obtained with the acetate buffer.

Furthermore, we characterized cleaved cross-sections of these crystals under FESEM along with an EDS analysis, as shown in Figure 5-10. We were able to easily identify the presence of phosphorous and absence of carbon in the rough rim. In the smooth, unreplaced region, we detected carbon but no phosphorous, as expected from a pure calcium carbonate phase. Note that the red dots in Figure 5-10 (b) in the smooth

139 calcite region are at the background noise level. Calcium and oxygen maps (not shown here) show that these elements are present in both phases, suggesting that the replaced phase is a type of calcium phosphate.

Figure 5-10: EDS map of sample S03 (buffer P2, pH 5). (a) Electron image and (b), (c) elemental maps of phosphorous and carbon respectively overlaid on the electron image, for a cleaved calcite crystal from S03 removed at 6 days.

5.5.4 Replacement with time

To see how deep the replacement rim could proceed in buffer B2, we carried out an experiment with time.

Figure 5-11 shows a plot of the rim width with time, where each time point is the average rim width (see section 5.4.5). The error bars are standard deviations from the same image. The standard deviations are inherently high since the replacement front is wavy as can also be seen in Figure 5-10.

The rim width increases with time, but the replacement growth slows down significantly after the first day.

A sample with buffer B15 (acetate buffer, pH 5.06, 2.0 M phosphate) was also reacted for 12 days at 50 C and the rim width was found to be similarly low for that case. This is surprising, since the replacement at day 1 should be the hardest (and slowest) from the point of view of dissolution and nucleation. That the replacement is even slower after day 1 suggests that a limitation is imposed by internal transport within the crystal18, perhaps resulting from a very low porosity. This is doubly surprising because for sparingly soluble minerals such as calcite, generally it is the surface reaction and not transport that is the rate limiting step.

140

Figure 5-11: Plot of rim thickness with time for S3 – calcite in pH 5 at 50 C, with and without mixing. The error bars are standard deviations calculated based on the variations in thickness from a single image.

5.5.5 Effect of external transport

Based on the observation of slowing down of rim growth after day 1, we ran the replacement experiment at pH 5 with external mixing at 600 rpm in the vessel. Under these conditions, the crystals themselves did not appear very different from those without mixing. Figure 5-12 a and b show FESEM images of cleaved cross-sections of samples reacted without and with mixing respectively. The average rim widths were almost equal for both the samples at 6 days and 12 days respectively as can be seen on the plot in Figure

5-11. Even though the rim width does not depend on the external transport, we argue that this does not rule out the possibility of transport being the rate limiting step after day 1. This is because external mixing may only change the concentration right outside the crystal without affecting the internal transport significantly.

Our argument is contrary to that of Xia et al26 who concluded that transport was not the rate limiting step in their system based on their observation that the rate decreased under conditions where external transport should have increased. In fact, we find a similar trend for the replacement of KCl in KBr (manuscript in

141 preparation) where the rate data points to a transport limited reaction but the rate does not increase with external mixing. We would like to point out that in our setup the shear field is very complex and is not easy to quantify, and that a well-defined necessary to accurately account for the effect of external transport.

Figure 5-12: FESEM images of cleaved calcite crystals showing the effect of phosphate concentration and mixing on texture of the replaced phase. Comparing (a) and (c), we can see that at 2 M the replaced layer consists of small crystals as opposed to much larger crystals in (c) at 0.5 M. There is no significant change due to mixing at 2.2 M (part (b)) but at 0.5 M, mixing changes the texture and porosity of the replaced phase drastically. Even though the rim width did not change significantly with mixing, we did observe the presence of a powdery precipitate in solution that settled at the bottom of the reaction vessel and also joined the crystals into one mass (Figure 5-3c). This presence of precipitate was observed even with replacement at 200 C

(Table 5-4: Conditions for additional mixing experiments.) which was accompanied with a mm scale rim width. During additional mixing experiments we saw more peculiar effects. For example, for sample S95

(Table 5-4) we recovered none of the crystals that we had started out with. For S90 (Table 5-4) one of three

142 bags had no crystals left in it. The solution turned milky in both cases. While further analysis was not carried out, we found that for S90 the solution was white and gelatinous, and the particles did not settle out completely even in 2 weeks – showing that these were nanoparticles. It appears that this ‘runaway dissolution’ happens when the mixing is too good, which was the case for both these samples as the bags accidentally jammed into the magnetic stir bar. Our hypothesis for this behavior is that very good external transport causes nucleation to happen in the entire solution rather that on the crystal surface alone, since no concentration gradients are allowed to exist. The absence of surface precipitation allows dissolution to proceed unhindered as the solution simultaneously causes precipitation throughout the solution.

Table 5-4: Conditions for additional mixing experiments.

Sample Soak (NH4)2HPO4 pH Mixing Temperature Solution Crystal time conc (C) volume mass S95 6 days 2.16 M 5.11 500 rpm 50 60 mL 0.5 g S90 6 days 2.16 M 5.11 500 rpm 50 210 mL 15 g (5g x 3 bags) S23 6 days 2.0 M 8.11 300 rpm 200 50 mL 0.2 g

143

Figure 5-13: S90, Runaway reaction with mixing at pH 5. Left : Crystals in one of the 3 bags suspended in the reaction were transformed from separate crystals to one big white mass, taking shape of the bag. (b) The solution had a gelatinous precipitate of white nanoparticles which did not settle in 2 weeks.

5.5.6 Effect of Phosphate concentration

For use of the replacement reaction for porosity generation, it is desirable to use lower phosphate concentrations. Kinetically, a lower phosphate concentration would mean that nucleation would become slower as the supersaturation level would be smaller (assuming the Ca+2 dissolution rate remains unaffected by phosphate concentration). We tested samples S05 and S06 (Table 5-3) in 0.5 M phosphate at pH ~5, without and with mixing respectively. Figure 5-12 shows images of cleaved cross-sections for both the concentrations tested (2.2 M and 0.5 M). For S05 without mixing, the crystal surface looked translucent and shiny on the outside due to relatively large crystals precipitated on the surface. The existence of the rim and a jagged boundary confirms that dissolution and precipitation has taken place. A very zoomed out view of the outer surface shown in Figure 5-14 (b) reveals the coarseness of this porosity.

144 Mixing at 0.5 M phosphate drastically changed the replacement outcome. Figure 5-12 (d) shows a cleaved section of the replaced crystal, where a thick and fluffy rim is observed in sharp contrast to the coarse rim observed without mixing. Unlike S04, very little or no powdery precipitate was obtained in the solution.

Comparing the outside surface of the crystals replaced in 2.2 M vs. 0.5 M phosphates, we find that the precipitated calcium phosphate crystals become coarser with reduced phosphate concentration, both with and without mixing. This may be a result of changes in the kinetics of the process due to changing supersaturation changing the growth mechanism, as well as changing Ca:P ratio25,27.

Figure 5-14: FESEM images of outside surfaces of replaced crystals. At 2.2 M (image a), the outer surface has roughness on a much smaller length-scale than at 0.5 M (image b) without mixing. In the presence of mixing, the outer surface at 2.2 M does not change much (image not shown) but at 0.5 M, the outer surface has much finer roughness due to mixing (image c). These images correspond to the rim cross-sections shown in Figure 5-12.

5.5.7 Effect of Ca+2 addition

Pre-existing Ca+2 ions in solution would have the opposite effect of reducing phosphate concentration – the ions would increase supersaturation level in solution. We tested the replacement of 4 g fractured calcite

+2 crystals in 2.2 M (NH4)2HPO4 at pH = 5.02 (buffer B2) along with 1 mM Ca(NO3)2 as a source of Ca . On the outside, the crystals did not look different from samples S03. But 1 of 4 crystals imaged from this experiment shows that replacement had proceeded on the mm scale along a fracture in the calcite crystal

(Figure 5-15). Since this effect was observed in only one of the crystals, we hypothesize that it depends on pre-existing fractures within the crystal. It is unclear whether this mm scale replacement along the fracture

145 was really a result of the Ca+2 addition, or was just a rare feature that was discovered by chance in this particular sample and none of the others.

Figure 5-15: Replacement along a fracture. FESEM image and EDS maps (carbon and phosphorous respectively) of the cross- section of a crystal from the experiment with buffer P2 with 1 mM Ca+2 added.

5.5.8 Mercury porosimetry at 6 days.

Mercury porosimetry was used to measure the porosity and pore size distribution in replaced crystals. The porosity increased 10-fold after 10 days of replacement - from 0.78 % in unreplaced calcite crystals to 7.47 % for sample S03 (pH 5.11, 50 C) and to 7.73 % for sample S07 (pH 6.7, 80 C). The average pore size was

76 nm for S03 and about 245 nm for S07, tested using 3-4 g of replaced crystals. Figure 5-16 shows mercury intrusion vs. applied pressure for the unreplaced and replaced samples. The high intrusion at higher pressures points to a large number of small pores.

146

Figure 5-16: (Left)Cumulative mercury intrusion pore volume as a function of pressure for various partly replaced crystals (S03, S07, S10) and untreated calcite crystals.(Right) Differential intrusion volume against equivalent pore diameter on a logarithmic scale. The area under the curve corresponds to total intrusion.

5.5.9 Confocal at 6 days and 12 days. For nicely cut vs. hammered.

To visualize the open porosity in replaced crystals, we soaked them in a fluorescent dye solution after replacement, and then observed cleaved sections under a confocal microscope, as explained in the Methods section.The bright areas in maximum intensity z-projections shown in Figure 5-17 reveal the spots where the dye penetrated from the outside of the crystal. Note that the top and bottom surfaces were cleaved before observation, and we not exposed to the dye solution directly. The side surfaces were exposed to the dye directly.

A control unreplaced calcite crystal (not shown in image) shows only a thin bright boundary and no bright spots inside confirming that pre-existing internal channels, if at all present, were not accessible to the dye in an unreplaced calcite crystal. The thin bright boundary results from the surface directly exposed to the

147 dye solution. Note that due to the rhombohedral shape of calcite crystals, two of the edges appear thicker than the others in z-projections.

Crystals shown in Figure 5-17 a,b show dye penetration for fractured calcite crystals soaked in pH 5 solution for 6 and 12 days respectively. The much greater depth of dye penetration at 12 days shows that either the replacement proceeded much faster at pre-existing internal channels, or replacement simply opened up access to these channels. In any case, at 12 days we see mm scale ‘deep penetration’, much larger than the rim width observed for the same sample using FESEM (~ 60 m, Figure 5-11).

Figure 5-17: Z-projections of maximum fluorescence from confocal microscopy. Crystals (a)-(c) were soaked in dye without pre- cutting, while (d) was cut before soaking. Crystals in (e) were saw-cut crystals rather than fractured crystals, replaced in buffer B15 (acetate) instead of P2 (HCl) used for (a)-(d) for 12 days. Figure 5-17 c shows a crystal replaced in the same solution as S3, but with mixing. In this case, even after

12 days, we do not see the ‘deep penetration’ we saw above. However, further investigation shows that the

148 reason for this lack of ‘deep penetration’ is that the pores that were opened during replacement get closed off towards the end of reaction due to excessive solution precipitation or due to a much longer diffusion

2+ 2- 2- path for the Ca , CO3 and HPO4 ions. Evidence in favor of this argument is provided by the crystal shown in Figure 5-17 d, where the crystal was cut post-replacement, before soaking to provide access to the pores that may have been closed in the end. This crystal does show ‘deep penetration’, similar to Figure

5-17 b.

Unlike Figure 5-17 a-d, the crystal shown in Figure 5-17 e was saw-cut rather than fractured. It was replaced for 12 days without mixing, just like the crystal shown in Figure 5-17 b but no deep penetration was observed in this case. We attribute this stark difference between saw-cut (Figure 5-17 e) vs. fractured

(Figure 5-17 b) to the presence of pre-existing internal channels in the fractured crystal. The replacement serves to either open up access to these, or go along it faster. Whichever the reason for it, this finding has important implications for using this replacement reaction for enhanced recovery – even if the rim width is

~ 100 m, mm scale porosity can be generated along pre-existing internal fractures.

5.6 Conclusions

We have reported pseudomorphic mineral replacement of calcite in phosphate solutions at temperatures 50 and 80 C under pH conditions in the range 4-8. These are the first observations for replacement of calcite in phosphate solutions at temperatures lower than 120 C. The replaced crystals have been studied using electron microscopy and elemental mapping, confocal microscopy and mercury porosimetry. Barring at pH

8, we find measurable replacement in each case, with thicknesses in the range 10 -100 m. pH variations lead to non-monotonic variation in thickness of the rim, its waviness, texture and overgrowth owing to a complex inter-dependence of the rates of dissolution, precipitation and transport, each of which depends on

149 the pH in complex ways. The replacement width with time data suggests that at 50 C and pH 5, the replacement is internal transport controlled.

Our study finds that pMRR with phosphate solutions is a promising approach to increase oil recovery from mechanical fracking. Mercury porosimetry measurements indicate at least a 10-fold increase in porosity after replacement of calcite for 6 days. While the replacement rim width at 50 and 80 C is smaller in all cases than the mm scale replacement observed at 200 C at pH ~8, observations of dye penetration using confocal microscopy reveal that mm scale ‘deep penetration’ of dye is obtained when fractured calcite crystals are replaced without mixing at pH 5 for 12 days. Thus, a significant amount of porosity can be gained and it would have the additional advantage of being far-reaching due to deep penetration in fractures created by mechanical fracking.

5.7 Future work and improvements to experimental setup

Based on the significant amount of preliminary experimental work presented here, there are some aspects of the experiments that could be improved upon in further experiments.

1. Here, I have used either fracturing using a hammer or cutting with a precise motorized saw, but in

future calcite crystals could also be prepared by using a lapidary rock saw supplemented with a

jeweler’s loupe for magnified viewing.

2. To understand the effect of external fluid transport, a set-up with a well-defined shear rate should

be employed.

3. EBSD (electron backscattered diffraction) or XRD (X-ray diffraction) could be used to identify the

difference phases present in the crystal after replacement, though in my own attempts at EBSD, the

replaced layer charged too much upon turning on the electron beam even after careful sample

preparation with ion-milling.

150 4. Ideally, edge defects and crystal roughness should be controlled or measured to estimate their

contribution to variability in dissolution rates.

5. BET could be used to supplement the pore-size analysis presented here, to probe even finer pores.

Additionally, our observations also raise some important questions that need further probing, which were outside the scope of the present study:

6. How would the porosity generated and the replacement reaction in general be affected by higher

pressures such as those found in geological reservoirs?

7. How do the replaced crystals compare with untreated crystals, and crystals undergoing only

dissolution in terms of their mechanical strength?

8. How does the porosity change with time - do created pores close down with time and convert into

closed porosity?

9. And lastly, how does the channel depth created here compare with the channel depth in crystals

replaced under high T and P conditions at pH ~ 8 by previous researchers?

While many questions remain about this system, our study finds a broad range of conditions where replacement takes place, creating visible and measurable porosity. The variety of textures and porosities obtained suggest many levers for controlling the reaction for porosity generation for enhanced oil recovery or for synthesis of novel materials.

5.8 Acknowledgements

Funding for this work was provided in part by Halliburton Energy Services, by NSF under the Active Matter

Transport projects and by NSF MRSEC IRG2. I am thankful to the researchers at Halliburton – Sanja Natali,

Ron Dusterhoft, Aaron Russell, Denise Benoit and Enrique Reyes for their valuable questions, comments and ideas. I would like to thank the staff at Materials Characterization Laboratory at Penn State, specifically

151 Julie Anderson for mercury porosimetry measurements and help with EDS measurements and Eric Golden and Katherine Crespin for help with sample preparation. I am thankful to Travis Tasker in department of

Civil and Environmental Engineering at Penn State for help with carrying out high P/T experiments, as well as for helpful discussions. Last, but not the least, I would like to thank Prof. James H. Adair in Materials

Science and Engineering at Penn State for his insightful suggestions and comments.

5.9 References

(1) Merino, E.; Dewers, T. Implications of Replacement for Reaction-Transport Modeling. Science

(80-. ). 1998, 209, 137–146.

(2) Putnis, C. V.; Tsukamoto, K.; Nishimura, Y. Direct Observations of Pseudomorphism:

Compositional and Textural Evolution at a Fluid-Solid Interface. Am. Mineral. 2005, 90, 1909–1912.

(3) Putnis, A. Mineral Replacement Reactions. Rev. Mineral. Geochemistry 2009, 70, 87–124.

(4) Altree-Williams, A.; Pring, A.; Ngothai, Y.; Brugger, J. Textural and Compositional Complexities

Resulting from Coupled Dissolution-Reprecipitation Reactions in Geomaterials. Earth-Science Rev.

2015, 150, 628–651.

(5) Lund, K.; Fogler, H. S.; McCune, C. C.; Ault, J. W. Acidization-II. The Dissolution of Calcite in

Hydrochloric Acid. Chem. Eng. Sci. 1975, 30, 825–835.

(6) Putnis, a. Mineral Replacement Reactions. Rev. Mineral. Geochemistry 2009, 70, 87–124.

(7) Wang, Y.; Wang, Y.; Merino, E. Dynamic Weathering Model: Constraints Required by Coupled

Dissolution and Pseudomorphic Replacement. Geochim. Cosmochim. Acta 1995, 59, 1559–1570.

(8) Jonas, L.; John, T.; Putnis, A. Influence of Temperature and Cl on the Hydrothermal Replacement

of Calcite by Apatite and the Development of Porous Microstructures. Am. Mineral. 2013, 98, 1516–

152 1525.

(9) Kasioptas, a.; Perdikouri, C.; Putnis, C. V.; Putnis, a. Pseudomorphic Replacement of Single

Calcium Carbonate Crystals by Polycrystalline Apatite. Mineral. Mag. 2008, 72, 77–80.

(10) Kasioptas, A.; Geisler, T.; Perdikouri, C.; Trepmann, C.; Gussone, N.; Putnis, A. Polycrystalline

Apatite Synthesized by Hydrothermal Replacement of Calcium Carbonates. Geochim. Cosmochim.

Acta 2011, 75, 3486–3500.

(11) Yoshimura, M.; Sujaridworakun, P.; Koh, F.; Fujiwara, T.; Pongkao, D.; Ahniyaz, A. Hydrothermal

Conversion of Calcite Crystals to Hydroxyapatite. Mater. Sci. Eng. C 2004, 24, 521–525.

(12) Guo, Y. P.; Zhou, Y. Conversion of Nacre Powders to Apatite in Phosphate Buffer Solutions at Low

Temperatures. Mater. Chem. Phys. 2007, 106, 88–94.

(13) Ni, M.; Ratner, B. D. Nacre Surface Transformation to Hydroxyapatite in a Phosphate Buffer

Solution. Biomaterials 2003, 24, 4323–4331.

(14) Pedrosa, E. T.; Putnis, C. V.; Renard, F.; Burgos-Cara, A.; Laurich, B.; Putnis, A. Porosity

Generated during the Fluid-Mediated Replacement of Calcite by Fluorite. CrystEngComm 2016, 18,

6867–6874.

(15) Klobes, P.; Meyer, K.; Munro, R. G. Porosity and Specific Surface Area Measurements for Solid

Materials; 2006.

(16) Grenthe, I.; Plyasunov, A.; Spahiu, K. Estimations of Medium Effects on Thermodynamic Data. In

Modelling in aquatic chemistry; OECD Nuclear Energy Agency, 1997; pp. 325–426.

(17) Brantley, S. L.; White, A. F.; Kubicki, J. D. Kinetics of Water-Rock Interaction; 2008.

(18) Lasaga, A. C. Theory of Crystal Growth and Dissolution. In Kinetic Theory in the Earth Sciences;

Princeton University Press, 1998; pp. 581–712.

153 (19) Fernández, E.; Gil, F. J.; Ginebra, M. P.; Driessens, F. C.; Planell, J. a; Best, S. M. Calcium

Phosphate Bone Cements for Clinical Applications. Part I: Solution Chemistry. J. Mater. Sci. Mater.

Med. 1999, 10, 169–176.

(20) Xia, F.; Brugger, J.; Chen, G.; Ngothai, Y.; O’Neill, B.; Putnis, A.; Pring, A. Mechanism and

Kinetics of Pseudomorphic Mineral Replacement Reactions: A Case Study of the Replacement of

Pentlandite by Violarite. Geochim. Cosmochim. Acta 2009, 73, 1945–1969.

(21) Rickard, D. T.; Sjoberg, E. L. Mixed Kinetic Control of Calcite Dissolution Rates. American Journal

of Science, 1983, 283, 815–830.

(22) Sjöberg, E. L.; Rickard, D. T. Calcite Dissolution Kinetics: Surface Speciation and the Origin of the

Variable pH Dependence. Chem. Geol. 1984, 42, 119–136.

(23) Sjoberg, E. L.; Rickard, D. T. Temperature-Dependence of Calcite Dissolution Kinetics between 1-

Degree-C and 62-Degrees-C at Ph 2.7 to 8.4 in Aqueous-Solutions. Geochim. Cosmochim. Acta

1984, 48, 485–493.

(24) Dolgaleva, I.; Gorichev, I.; Izotov, A.; Stepanov, V. Modeling of the Effect of pH on the Calcite

Dissolution Kinetics. Theor. Found. Chem. Eng. 2005, 39, 614–621.

(25) Wang, L.; Nancollas, G. H. Calcium Orthophosphates: Crystallization and Dissolution. Chem. Rev.

2008, 108, 4628–4669.

(26) Xia, F.; Brugger, J.; Chen, G.; Ngothai, Y.; O’Neill, B.; Putnis, A.; Pring, A. Mechanism and

Kinetics of Pseudomorphic Mineral Replacement Reactions: A Case Study of the Replacement of

Pentlandite by Violarite. Geochim. Cosmochim. Acta 2009, 73, 1945–1969.

(27) Dorozhkin, S. V. Calcium Orthophosphates in Nature, Biology and Medicine. Materials (Basel).

2009, 2, 399–498.

154

Chapter 6

Future Work and Broader Impact

The objective of this dissertation was to gain a better understanding of aqueous saturated salt systems from a colloidal and surface chemistry perspective by answering questions about the surface charge and dissolution-precipitation reactions. To conclude, I would like to point out possible directions for further research and broader impact of the concepts and experimental techniques developed in this dissertation.

6.1 Zeta potential at high and saturated salt

6.1.1 Scientific advance

From my study, I concluded that the zeta potential of surfaces can be measurable and finite even under high salt conditions such as saturated NaCl (5.3 M). The magnitude and sign of the zeta potential at high salt depends on the surface-functionalization and can depend on the salt identity due to ion-specific interactions. Ion-specific interactions, especially with highly charged surfaces can lead to charge reversal even with monovalent salts. The work presented here is the first ever set of measurements of zeta potential on model polystyrene latex particles at such high salt concentrations. From a theoretical standpoint, this is a significant breakthrough in building a more accurate picture of the EDL at high salt.

155 6.1.2 Implications for saturated salt systems

Looking from a more practical stand-point, the presence of a finite zeta potential (~ 25 mV in our measurements) will lead to a finite electrokinetic particle or fluid transport in the presence of a high enough electric field (10 V/cm field can produce a velocity of ~ 5 m/s). We also find that the zeta potentials we measured were too small for electrostatic stabilization (~ 25 mV) at saturated salt.

The electric field may be generated spontaneously in presence of salt concentration gradients for very short times but will likely not be sustained for a long time. Even if such fields are applied externally to generate fluid or particle transport, a high electric current will be generally produced producing joule heating and potentially overshadowing the electrokinetic effect. However, the electrokinetic effect could be observed and used if a high electric field can be maintained and heat generation controlled by minimizing the current and maximizing the heat transfer. Additionally, systems at higher temperatures usually have a higher zeta potential1 and lower viscosity, which could translate into electrokinetic effects becoming more important at high temperatures for example in geological reservoirs.

6.1.3 Broader applicability of technique to biological systems

Apart from a scientific advance in understanding of zeta potential at high salt, the technique developed here can have a huge impact on our ability to understand the electrical behavior of biological colloidal systems where bacteria or human cells have a very low density contrast with the solution. Bodily fluids blood and urine have high salt concentrations for which the zeta potentials from commercial instruments are not reliable. The technique described here can provide these measurements reliably, and under constant microscopic observation that can allow us to study the charge on particles during transitions or interactions with other surfaces in the solution. This

156 can no doubt help unravel hitherto unsolved mysteries about the biological world. While some of the instrumentation required for it is not cheap at the moment, (namely a high speed camera) rapid advances in sensors, processing and memory technology will make these ubiquitous in research labs or even consumers within a few years. At the time of this writing, a 1000 fps camera is available on the market on Kickstarter under $1000. Such hardware combined with a simple software tool will enable microscopy based zeta potential measurements at high salt routine.

6.1.4 Future directions

First, the technique developed here is ripe for commercialization for application to biological systems. I think this is by far the most important future direction from this work. The technique developed here for measurement of zeta potential at high salt of particles can be extended to fixed surfaces in some cases. For example, the zeta of the glass capillary holding our particles can be easily measured if the depth in the cell can be measured accurately. This can broaden the scope of this technique.

Second, for application to geological reservoirs, we need to measure the zeta at high salt and at high temperatures. Measurements have been done previously by Lvov and coworkers2 at high temperatures, but at low salt concentration, but there is no method so far to carry out measurements at both high salt and T. Zeta potentials usually increase with temperatures and viscosity becomes lower, both of which will lead to greater electrokinetic transport at higher temperatures. Thus, we might find that indeed salt concentration gradient driven electrokinetic transport (diffusiophoresis) does have a role to play in enhanced oil recovery.

157 6.2 Pseudomorphic mineral replacement reactions

6.2.1 Scientific advance

There are two key scientific advances of broad interest to pseudomorphic mineral replacement reactions that came through the work presented in this dissertation. The first advance is that I have shown how kinetic and fluid transport effects can lead to replacement in a thermodynamically unfavorable system – KCl in KBr. The second advance was identification of the differences between internal and external fluid transport, showing that the progress of replacement is often controlled by internal fluid transport and yet, external transport through mixing does not make a difference to the rate of replacement.

My work also advances our understanding of pMRR in the calcite/phosphate system. This is the first study that discusses the replacement behavior in detail under moderate temperatures in the range 50 – 80 C while most previous studies discussed the behavior under hydrothermal conditions.

My work shows that we get small but finite replacement in this temperature range, and that pH is an important variable to control the replacement extent, rate and texture. The behavior of the system under any new conditions is not easy to predict based on theory due to the tangled inter-dependence of kinetics, transport and thermodynamics in pMRR, so experimental studies such as these are necessary to model the system well enough in order to eventually be able to predict and design behavior.

6.2.2 Application to enhanced oil recovery

The study of replacement in the calcite/phosphate system presented in this dissertation was aimed at using these reactions for enhanced oil recovery through fracking of geological reservoirs,

158 especially low temperature calcite reservoirs in the temperature range 50 – 90 C. In that context, my findings of up to ~100 m replacement width, a 10-fold increase in porosity (from 0.7% for untreated calcite to >7 % for replaced calcite) and the opening up of mm scale pre-existing internal fractures provide a proof-of-concept for this application. Increasing oil recovery from these massive fracking operations by simply changing the fracking fluid chemistry with relatively benign materials (ammonium phosphate is a common bulk fertilizer) and without additional water, can lead to a huge environmental saving in terms of the cost of drilling additional reservoirs and using more water for them.

6.2.3 Applicability to synthesis of templated or doped materials pMRRs have been used in a few studies for synthesis of novel templated materials. A specific application could be designer calcium phosphate based materials for drug delivery where the ability to template, create porosity and doping with drug particles would be of immense importance.

Additionally, the reaction could be carried out in multiple steps, to create multiple layers of materials for controlled drug release of multiple drugs. Lastly, pMRR can also be designed to create an end-product with desired mechanical properties – perhaps even stronger than the starting material, as is the case for calcite in phosphate solutions.

The experimental techniques developed here could be especially useful when designing a pMRR for a new application such as synthesis. For example, quantitative EDS of the replaced solid could be used to study the composition and confocal dye penetration could serve as a marker to observe the fluid pathways if porosity is of interest for the application. My studies with calcite/phosphate and KBr/KCl under many different conditions serve to illuminate some of the levers for controlling a reaction – time, temperature, solution to solid ratios, pH, reagent concentrations, mixing rate and exposed surface areas.

159 6.2.4 Future research directions

The most important future direction towards an industrial application would be to follow-up the calcite/phosphate study with an economic feasibility analysis and a second stage proof-of-concept by carrying out a lab replica of fracking using a phosphate solution as the fracking fluid.

The work on the KBr/KCl system should be followed by a similar analysis with the NaCl/KCl system, because in spite of some similarities, my preliminary work has shown very different behavior – this system does not seem to replace at all. Reconciliation of observations from similar simple systems would be an important step towards really understanding these reactions.

Lastly, it would be valuable to apply this replacement reaction for synthesis of calcium phosphate based biomaterials or for designer, multi-step controlled-release drugs.

6.3 References

(1) Kirby, B. J.; Hasselbrink, E. F. Zeta Potential of Microfluidic Substrates: 1. Theory,

Experimental Techniques, and Effects on Separations. Electrophoresis 2004, 25, 187–202.

(2) Zhou, X. Y.; Wei, X. J.; Fedkin, M. V.; Strass, K. H.; Lvov, S. N. Zetameter for

Microelectrophoresis Studies of the Oxide/water Interface at Temperatures up to 200 °C.

Rev. Sci. Instrum. 2003, 74, 2501.

160

Appendix A

Supplemental Information for Zeta at High Salt2

A.1 Lattice Model of the Electric Double Layer

We consider a symmetric monovalent electrolyte at thermal equilibrium within a rectangular cell of dimensions L1 × L2 × L3 in contact with a charged surface at x3 = 0. Within the cell, N ion pairs,

M surface charges, and M counterions are distributed on a regular cubic lattice with spacing ℓ

(Figure A-1). The lattice spacing is assumed to be comparable to the size of a solvated ion, such that each lattice site can accommodate a single ion. The M surface charges are distributed at random onto lattice sites within the x3=0 plane to create an average surface charge density of σ = eM=L1L2.

The 2N + M ions are distributed at random onto bulk lattice sites (x3  0) to give a nominal salt concentration of nc = N/L1L2(L3 - ℓ) within the cell.

2 This chapter has been adapted with permission from A. Garg, C. Cartier, K. Bishop, D. Velegol, Particle Zeta Potentials Remain Finite in Saturated Salt Solutions. Langmuir, 32, 11837. Copyright 2016 American Chemical Society. The work presented here on the theoretical model was done primarily by Prof. Kyle Bishop and I assisted him in testing out some of the simulations.

161

Figure A-1: Schematic illustration of the lattice model. The positions of the ions are equilibrated using the Metropolis Monte Carlo algorithm1; those of the surface charges are fixed throughout the simulation. During each Monte Carlo move, two sites are selected at random and their contents swapped. If the move lowers the electrostatic energy of the system, it is accepted unconditionally. If it raises the energy of the system, it is accepted with probability

(1) where kBT is the thermal energy, and ∆U is the energy increase accompanying the change in ion configuration. For lower salt concentrations, several such lattice swaps are conducted during each

Monte Carlo move to achieve an acceptance frequency of around 50%. During each simulation, the system is equilibrated for 5 × 105 attempted moves; the resulting distribution is then sampled over the course of an additional 5 × 105 moves. Each condition is simulated ten times using different realizations of the surface charge distribution; the final results are obtained by averaging over these realizations.

The electrostatic energy U is computed using the Ewald method1, which is summarized here for clarity. For a given distribution of point charges, the electrostatic energy is

162

(2)

th where xn and zn = ±1 are the position and valence of the n charge. The electric potential φ at each lattice site is governed by the Poisson equation

(3) where  is the uniform permittivity of the medium, and ρ(x) is the charge density

(4)

The boundary conditions are assumed to be periodic in all directions, such that the charge density and the potential can be expressed as Fourier series

(5)

(6)

with wavevectors kj = m/Lj for m = 0; ±1; ±2,…. Substituting these expansions into the Poisson equation and solving for the potential, we obtain the following expression for the potential in terms of the Fourier coefficients of the charge density

(7)

Note that the zero wavevector (k = 0) contributes an arbitrary constant and is therefore excluded.

This slowly converging summation (eq 7) can be accelerated using the classic Ewald approach,

163 whereby the point charges are replaced by Gaussian charge distributions in the wave-space summation. The error introduced by this approximation is then corrected in real-space to obtain

(8)

where r = x - x0, and α is the so-called splitting parameter, which characterizes the width of the

Gaussian distribution. Here, we choose α to be sufficiently small that only self-contributions (when r = 0) need be considered in the real-space calculation. With this assumption, the potential at xn due to all other charges is

(9)

The wave-space contribution is computed rapidly on a finer grid using the Fast Fourier Transform

(FFT) to obtain the electric potential at each of the lattice sites. We use an FFT grid spacing of δ =

ℓ/4 and splitting parameter of α = 2δℓ. With these parameter values, this approach reproduces the

Madelung constant of NaCl to three significant digits; higher precision can be achieved by decreasing δ but at greater computational expense. Finally, we define an average potential φ(x3) as

(10) which is useful for comparisons with the results of a 1D continuum model (see below).

Scaling lengths by ℓ, charges by e, potentials by e/4πℓ, the above model is characterized

3 2 by three dimensionless parameters: (i) ncℓ the electrolyte concentration, (ii) σℓ /e the surface

2 charge density, and (iii) β = e /4πℓkBT the inverse temperature. The parameter β can also be

2 interpreted as the ratio between the Bjerrum length λ = e /4πkBT and the ion size ℓ. For pure water

164 at 25 oC, the Bjerrum length is λ = 7.2 Å and the hydrodynamic diameter of NaCl is ℓ = 3.1 Å such that β = 2.3. Figure A-2 shows the average potential φ(x3) for β = 2.

Figure A-2: Average potential φ (scaled by e=4π퓵) as a function of position x3 (scaled by 퓵) as computed by the lattice model (black circles). Here, the simulation cell is L1 = L2 = 8퓵 by L3 = 32퓵 and contains M = 4 surface charges and N = 20 ion pairs; the inverse temperature is β = 1. The error bars represent 95% confidence intervals based on ten independent simulations. The open circles represent the average ζ- potential, which is assumed equal to the electric potential at x3 = 0.5퓵. The solid curve shows the solution to the nonlinear Poisson-Boltzmann equation for the same conditions.

A.2 Continuum Model of the Electric Double Layer

We now treat the same problem of the potential between two charged planes using the standard continuum theory based on the Poisson-Boltzmann equation. The results of this continuum description are compared to those of the lattice model described in the previous section. Integrating equation (S9) over the 1 and 2-directions and making use of the periodic boundary conditions, one can show that the average potential

165

(11) subject to the following boundary conditions at the charge surface

(12)

Additionally, we make the mean-field assumption that the ion concentrations within the electrolyte are Boltzmann-distributed as

(13)

0 where n ± are normalization constants for positive/negative ions. Assuming the surface is positively charge, these constants are determined by the following integral constraints on the total number of positive and negative ions within the cell

(14)

where nc is the electrolyte concentration in the cell, and the additional contribution to n- is due to the neutralizing counter-ions. The resulting charge density is given by

(15)

Note that there exists a potential φ0 for which the charge density is identically zero; this reference potential corresponds to some location within the bulk electrolyte far from the charged surfaces.

Substituting the above expression for the charge density into the Poisson equation, we obtain the

Poisson-Boltzmann equation for a symmetric monovalent electrolyte

(16)

166

Here, n0 is the salt concentration in the bulk electrolyte, which is somewhat larger than the nominal concentration nc within the cell

(17)

We use MATLAB’s bvp4c() function to solve equation 16 numerically subject the boundary conditions (equation 12) and the integral constraints (equation 14).

For small potentials (φ - φ0 < kBT/e), the Poisson-Boltzmann equation can be linearized and solved analytically. The resulting approximation provides a convenient initial guess for computing the numerical solution to the nonlinear problem. We start by linearizing equation 16 about the reference potential φ0 to obtain

(18)

2 2 where κ = 2e n0/kBT is the screening parameter. Solving this equation and applying the boundary conditions (equation 12), the potential in the cell becomes

(19)

To determine the bulk concentration n0, we first note that the average potential across the cell is related to the surface charge as

(20)

Using this relation, the integral constraints (equation 14) can be linearized to relate the bulk concentration to the cell concentration as

(21)

167 again assuming that σ > 0. Importantly, when the screening length is much smaller than the size of the cell (κL3 >> 1), the potential near the charge surface at x3 = 0 is well approximated as

(22)

A.3 Summary of zeta potential measurements

Table A-1: Summary of zeta potential measurements Zeta uncertainty (90% Relative Zeta confidence particle salt Molarity Conductivity Viscosity permittivity potential intervals) mM mS/cm Pa.s mV mV apsl KBr 1000 112.5 0.000853 67.5 -3.1 3.7 apsl KBr 2000 212 0.000835 55.8 -13.5 5.3 apsl KBr 3000 346 0.000834 47.6 -17.9 6.8 apsl KBr 4000 430 0.000846 42.9 -20 7.9 apsl KBr 4600 464 0.000859 40.4 -17.6 11.9 apsl KBr 500 57.3 0.000869 70.5 4 4 apsl KCl 1000 112.6 0.000892 67.1 7 4 apsl KCl 10 1.88 0.000892 77.9 38.7 5.2 apsl KCl 2000 217.5 0.000903 58.1 -0.8 5.2 apsl KCl 3000 305 0.000928 49.8 -0.2 6.7 apsl KCl 500 57.3 0.000891 72.1 17.2 3.5 apsl KCl 750 85 0.000891 69.2 7.2 3.9 apsl LiCl 1000 75.6 0.001018 65.3 -3.7 4.4 apsl LiCl 2000 124 0.001173 54.5 -6.6 5.6 apsl LiCl 4000 178.6 0.001587 37.1 -6.5 11.5 apsl LiCl 500 44.8 0.000953 70.8 4.9 4.2 apsl LiCl 750 61 0.000985 68 -3.1 4.3 apsl NaCl 1000 84.3 0.000981 66.6 12.8 4.6

168 apsl NaCl 100 10.8 0.000901 76.7 15.7 3 apsl NaCl 10 1.18 0.00089 78 68.6 4.1 apsl NaCl 1 0.128 0.00089 78.1 74.3 4.7 apsl NaCl 2000 151.6 0.001096 58 -1.6 5.4 apsl NaCl 4000 234 0.001443 46.5 -0.1 10.1 apsl NaCl 5400 254 0.001705 42.8 -6.6 12.5 cpsl LiCl 2000 124 0.001173 54.5 -22.1 5.6 cpsl LiCl 4000 178.6 0.001587 37.1 -17.1 11 cpsl NaCl 1000 84.3 0.000981 66.6 -26.7 4.5 cpsl NaCl 100 10.8 0.000901 76.7 -55.2 4.2 cpsl NaCl 10 1.18 0.00089 78 -72.2 4.9 cpsl NaCl 1 0.128 0.00089 78.1 -93.3 6 cpsl NaCl 2000 151.6 0.001096 58 -19.4 5.4 cpsl NaCl 4000 234 0.001443 46.5 -18.9 9.4 cpsl NaCl 5400 254 0.001705 42.8 -19.6 11.9 spsl CsCl 1000 111.5 0.000857 68.7 -36.5 4.1 spsl CsCl 2000 221 0.000846 63.6 -28.8 5.9 spsl CsCl 3000 323 0.000848 53.9 -31.7 5.8 spsl CsCl 4000 416 0.000874 48.1 -28 7.6 spsl KBr 1000 117.5 0.000853 67.5 -33 3.9 spsl KBr 2000 212 0.000835 55.8 -29.4 5.4 spsl KBr 3000 346 0.000834 47.6 -25.7 7.2 spsl KBr 4000 430 0.000846 42.9 -25.2 8.4 spsl KBr 4600 464 0.000859 40.4 -34.2 10.1 spsl KCl 1000 111.8 0.000892 67.1 -20.7 3.4 spsl KCl 2000 217.5 0.000903 58.1 -16.8 7 spsl KCl 250 29.4 0.000892 74.1 -57.8 3.3 spsl KCl 3500 360 0.000931 46.3 -24.1 7.7 spsl KCl 4200 405 0.000945 44 -13.7 8.8 spsl KCl 500 57.3 0.000891 72.1 -45.2 3.5 spsl KCl 750 85 0.000891 69.2 -39.3 4.4 spsl LiCl 1000 75.6 0.001018 65.3 -42.8 4.9

169 spsl LiCl 1500 103.2 0.001087 58.6 -46.2 5.7 spsl LiCl 2000 122.8 0.001173 54.5 -36.2 6.2 spsl LiCl 250 23.7 0.000923 74.3 -59.9 3.7 spsl LiCl 4000 178.6 0.001587 37.1 -35.3 11.1 spsl LiCl 500 39.8 0.000953 70.8 -35.6 3.8 spsl LiCl 750 61 0.000985 68 -43.5 4.3 spsl NaCl 1000 84.3 0.000981 66.6 -42.2 4.5 spsl NaCl 100 10.8 0.000901 76.7 -74.5 3.6 spsl NaCl 10 1.18 0.00089 78 -118.2 4.5 spsl NaCl 1 0.128 0.00089 78.1 -100.9 6.1 spsl NaCl 2000 151.6 0.001096 58 -29.2 5.1 spsl NaCl 3000 201 0.001246 51.5 -30.6 7.6 spsl NaCl 4000 234 0.001443 46.5 -28.8 9 spsl NaCl 5400 254 0.001705 42.8 -21.2 11.6

A.4 Contributions

The work on the theoretical model presented here was done by Prof. Kyle Bishop.

A.5 Copyright Notice

Adapted with permission from Garg, Astha, et al. "Particle Zeta Potentials Remain Finite in

Saturated Salt Solutions." Langmuir 32.45 (2016): 11837-11844. Copyright (2016) American

Chemical Society.

A.6 References

(1) Frenkel, D.; Smit, B. Understanding Molecular Simulation : From Algorithms to

Applications; 2nd ed.; Academic Press: San Diego, 2002.

170

Appendix B

Boundaries can steer active Janus spheres3

B.1 Abstract

The advent of autonomous self-propulsion has instigated research towards making colloidal machines that can deliver mechanical work in the form of transport, and other functions such as sensing and cleaning. While much progress has been made in the last ten years on various mechanisms to generate self-propulsion, the ability to steer self-propelled colloidal devices has so far been much more limited. A critical barrier in increasing the impact of such motors is in directing their motion against the Brownian rotation which randomizes particle orientations. In this context, here we report directed motion of a specific class of catalytic motors when moving in close proximity to solid surfaces. This is achieved through active quenching of their Brownian rotation by constraining it in a rotational well, caused not by equilibrium, but by hydrodynamic effects. We demonstrate how combining these geometric constraints can be utilized to steer these active colloids along arbitrary trajectories.

3 This chapter has been adapted from the scholarly article Boundaries can Steer Active Janus Spheres published in Nature Communications, 6:8999 (2015), licensed under a Creative Commons Attribution 4.0 International License, and is attributed to S. Das, A. Garg, A. I. Campbell, J. Howse, A. Sen, D. Velegol, R. Golestanian and S. J. Ebbens,

171 B.2 Introduction

The ability to accurately steer self-propelled particles without the application of an external force field should have far reaching consequences. Such precise navigational control is essential for cargo transport1–3, repair and drug delivery within the body4, sensing5, environmental remediation6 and even micro-surgery targeting individual cells.

Self-propelled colloidal motors7,8 which transduce chemical energy into mechanical motion are an important class of active matter9,10,11. A critical challenge is in directing the motion of such active colloids, which include bi-metallic nanorods,12 nanotubes13 and Janus spheres14. The major barrier is Brownian rotation which randomises particle orientations15, leading to long-time isotropic enhanced diffusion16,17. Current strategies for directing catalytic motors include external magnetic fields18, Earth’s gravitational field,19,20 and electrophoretic traps21. However, these strategies either lack autonomy22 (external fields and traps) or only constrain along 1-D (gravity).

Here we introduce a new method of steering individual micromotors, using geometric boundaries, in a way that does not lead to “global” steering, as happens when external fields are applied. We demonstrate directed motion of a special class of catalytic motors – the spherical Janus colloid with half-coating of platinum with variable thickness – when moving in close proximity to solid surfaces through active quenching of their Brownian rotation, which leads to constrained in- plane swimming along the wall. Such prolonged directed transport is not dependent on any external fields or potentials and continues for length scales much larger than previously reported23. We find that a very specific set of criteria needs to be satisfied for this steering mechanism to work, and that it is the dynamic flow field – not the equilibrium interaction of the particle with the wall – that constrains the motion from enhanced 3-D translational diffusion in the bulk to constrained 2-D enhanced diffusion at the walls. By combining geometric constraints, we demonstrate how it is

172 possible to reduce the number of degrees of freedom for these autonomously moving catalytic colloids.

B.3 Results and Discussion

B.3.1 Experimental Characterization of the Orientational Quenching

Our system consists of Janus24 particles composed of platinum (Pt) -capped polystyrene

(PS) colloids, which swim in solutions of hydrogen peroxide (H2O2) due to its catalytic decomposition into water and oxygen14. These Janus particles swim with the PS side forward, according to a self-electrophoretic mechanism, wherein ionic currents are set up at the Pt end, between the poles and the equator25,26. We aim to study these colloids stochastic trajectories when they swim near an interface, so that we can characterize its effect on their active motion.

The polar angle, , which probes how the polarity of a Janus particle is oriented with respect to the surface normal vector, and the corresponding in-plane orientation can be observed under fluorescence microscopy as ‘phases of the moon’27 (see Methods), Figure B-1. In the absence of H2O2, Janus particles sediment in a rectangular cuvette to the bottom wall and are observed to undergo Brownian translational and rotational diffusion about a cap-down equilibrium orientation dictated by gravity, due to the weight of the platinum ( Figure B-1 a, b). But in the presence of

H2O2, the Janus particles undergo enhanced diffusion in the bulk and accumulate at both the top and bottom walls, according to the competition between gravitaxis19 and sedimentation (see section

B.6.2) . Strikingly, once close to the wall, for sufficiently high H2O2 concentrations, the swimmers maintain an orientation such that the half-moon shape is persistently visible (Figure B-1 a, left and c). This suggests that the out-of-plane rotational diffusion is quenched so that the polar angle

(defined in Figure B-1c) is  90° (Figure B-1a, right, and Supplementary Movies 1-3).

173 The relationship between propulsion speed and quenching becomes clear from Mean

Square Angular Displacement (MSAD) curves for many Janus particles (n>25 at each condition) arranged by size and binned according to propulsion speeds (which was varied by changing fuel concentration, see Figure B-8), along with a comparison to un-fuelled, purely Brownian Janus colloids (Figure B-1 d). For particles undergoing un-quenched Brownian rotation, we have

2 MSAD= 〈Δ휃(푡) 〉 = 2퐷푟Δ푡. In water, a linear time dependence of the MSAD is observed at short time periods (MSAD trend towards an upper limit at longer times due to the periodic boundary conditions for the polar angle, 0° ≤ ϴ ≤ 180°). However, as the magnitude of v increases, the

MSAD becomes sub-diffusive after an initial period of linear time dependence at short timescales, and saturates at long times, to within our experimental resolution (Figure B-1 d). The strength of the orientational quenching is manifestly increased as the propulsion velocity is increased.

174

Figure B-1: Brownian rotational quenching and alignment near a planar surface (a) Left: Selection of frames from fluorescent microscopy videos (15 µm x 15 µm field of view) for fluorescent platinum-polystyrene (PS) Janus spheres of varying radii, the PS side of the colloid appears bright (a), near to a planar interface in de-ionized (DI) water, and in 10 % aqueous H2O2 solutions. Right: Polar angle, ϴ(t) for typical Janus particles determined from fluorescent microscopy videos (Note the a=2.4 µm particle in water shows strong gravitational alignment constraining ϴ close to 0° ). (b, c) Schematic 3D orientation and experimental trajectories (45 seconds duration, red line) for Pt-PS Janus particle with radius a = 1 µm in (b) DI water settled under gravity against a planar glass substrate and (c) 10 % H2O2 solution at either the top (+g) or bottom (-g) planar surface of a rectangular glass cuvette. (d) Polar Mean Square Angular Displacement (MSAD) as a function of time for three differently-sized Janus spheres. In each graph, the black “Water” line represents the MSAD for Janus particle settled at a planar interface under gravity in water (n>20). The additional curves represent the MSAD for Janus particles with speeds in the defined ranges, at both the top (n>20) and bottom (n>20) planar surfaces of a rectangular cuvette.

175 B.3.2 Steering the Active Colloid

To further probe the role of geometry on the orientational quenching, we examine swimming of Janus particles near two orthogonal surfaces (Figure B-2 a). We observe that a particle initially exhibiting 2-D enhanced diffusion at the surface (Figure B-2 c) undergoes persistent linear motion when it reaches the vertical edge of the cuvette. Figure B-2 d,e show examples of active colloids following curved and straight sections of the boundary for appreciable distances. The inset of Figure B-2 d also shows the transition from 2-D enhanced diffusion to boundary steering occurring at the moment the colloid reaches the wall. Additionally the left hand inset for Figure

B-2 e (and Supplementary Movie 4) verifies that the colloid equator is aligned at close to 90° as seen above for a single boundary. In fact, over many repeated experiments we observed that all the colloids investigated (which were of different sizes) were directed by the edge of the cuvette for length scales up to several centimetres. Colloidal motion continued around the entire macroscopic cuvette edge, only occasionally stopped by small blemishes (standard laboratory glassware that had not been precision engineered was used), or by encountering other stuck colloids, resulting in pile-ups of aligned colloids, Figure B-2 e right hand inset. Due to the build-up of colloids following the edge, sometimes in different directions, collisions between moving colloids were also observed (Supplementary Movie 5). Collisions result in stable agglomerates which continued to move along the boundary in a direction that qualitatively appeared to be determined by the relative propulsive velocities of the colliding components.

We also investigated an array of lithographically produced deep rectangular linear channels. Since the Janus particles now interact with three confining surfaces (two parallel walls and the bottom surface, as seen in Figure B-2 b), geometric constraining should lead to a strictly linear motion of the Janus particles in the channels, with few or no “Brownian escapes”. Grooves with a variety of widths were investigated, and the colloids with radius a = 2.4 µm were observed to be rotationally

176 quenched and surface-aligned within channels with widths 7-9 µm. For example, Figure B-2 f (and

Supplementary Movie 6) depicts an active Janus colloid exhibiting persistent linear motion when confined within a 9 µm wide channel, see Figure B-2 g.

177

Figure B-2: Particles moving along geometric boundaries, at speeds of up to 10 µm/s. (a-b) schematics of Janus particles encountering multi-planar geometries. Red axis indicates forbidden rotations due to proximity to a plane, green axis indicates unquenched axis of rotation: (a) Janus particle encountering a planar edge while moving along a 2D surface, expected to result in Brownian rotational quenching about two orthogonal axes. (b) Janus particle confined within a square groove; parallel vertical walls confine the rotational diffusion about one axis; however, if the particle descends

178 to the base of the groove, it is confined about two orthogonal axes. (c-f) Overlaid still frames from fluorescence microscopy videos with equal time gaps: yellow line shows complete trajectory, green line shows location of vertical cuvette walls, red arrows indication direction of motion: (c) a = 1.55 µm Janus particle (10 % H2O2) moving at the bottom of a rectangular glass cuvette a long way away from the edges. (d) a = 1.55 µm Janus colloid (10 % H2O2) moving along the curved edge of a glass cuvette – inset shows a colloid reaching the cuvette boundary. (e) a = 2.4 µm Janus colloid (10 % H2O2) moving along the straight edge of a glass cuvette- left hand inset shows a magnified region, right hand inset shows a “stuck” aligned agglomerate formed at the cuvette boundary. (f) a = 2.4 µm Janus colloid (10 % H2O2) moving within a rectangular section groove (width = 8.75 µm). (g) SEM image of a section of the rectangular grooves (widths 7-9 µm) used in (f).

B.3.3 Theoretical description of the catalytic colloid near a surface

Having determined the phenomenology of rotational quenching experimentally, we now set out to investigate this effect theoretically, and explore possible mechanisms that can account for these observations. The orientational dynamics of our active Janus particles could be affected by several mechanisms. These include equilibrium effects such as gravitational torque due to inhomogeneous weight distribution in the platinum cap and electrostatic interaction with the surface due to zeta potential difference between the two halves of the Janus particle. There are also non-equilibrium effects such as hydrodynamic coupling between the swimmer and the surface28–30, electroosmotic effects due to the ionic activity of the active Janus colloid31,25, and electrostatic contributions due to additional difference in zeta potential between the two halves as a result of the non-equilibrium catalytic activity on the platinum cap. We take a pedagogical approach and construct an approximate theoretical description of the phenomenon, so that we can highlight the key physical features and the relative significance of the above effects.

There are several observations pointing to the fact that we observe a non-equilibrium effect which is a result of the propulsion mechanism of the particles. Higher concentrations of the fuel, H2O2, result in a greater fraction of time spent by the particle in the half-moon orientation. As expected, higher H2O2 concentrations also result in higher particle speeds both at the surface and in the bulk

(Figure B-8). Moreover, when the electrokinetic component of the propulsion mechanism is

179 weakened by addition of 1 mM NaCl25, rotational quenching is also observed to decrease. We know that the gravitational torque is not responsible for the surface aligned Brownian rotation quenching since no surface alignment is observed in the absence of H2O2, and the surface aligned rotational quenching is apparently independent of the direction of gravity, as mentioned above.

B.3.4 Electrostatic colloidal forces.

If the surface potentials of the particles were to change significantly in the presence of the reaction, electrostatic colloidal forces could contribute to the rotational quenching. In order to estimate the contribution of the electrostatics with and without catalytic activity, we measured the zeta potential of each end (Pt and PS) of the Janus particle by combining both translational and rotational electrophoresis experiments in 1 mM NaCl, Figure B-3 a. In carrying out these measurements, the presence of bubbles due to the reaction at Pt end produces 2 challenges, which we addressed: 1) increased resistance to electric current, since the bubbles occupy cross section where the current could flow, and 2) complex convective flows around bubbles. We designed a disposable electrophoresis setup (B.6.5), applied a constant current across it, and worked in cross sections of the cell where we observed no bubbles, to get a constant electric field in the solution phase. The translational velocities of janus particles were measured with respect to velocity of tracers that had a known zeta potential, to account for the any convective flows, including electroosmotic flow near the cell wall. The transport of the janus particles and tracers were always found to be smoothly linear.

180

Figure B-3: Electrophoretic behaviour for Janus colloids. (a) Schematic representation of the rotational electrophoresis experiment. Left hand side shows the relevant physical quantities, a Janus sphere with hemispheres with two different zeta potentials (Pt and PS), giving a dipole vector 풆̂. The dipole vector rotates in an electric field, and the right hand side 3D schematics depict the effect of switching the direction of E. 1. Represents the initial misaligned dipole and applied field orientation immediately after the E field direction is switched, stages 2-5 show two possible rotations to re- align the dipole with the applied field: on the left hand side about an out of plane axis, with constant polar angle, ϴ, and on the right about an axis parallel to the plane where polar angle changes; at position 5 풆̂ reaches the steady state. The black arrows show the direction of translational motion, which is always aligned with the applied field (see b). (b) Typical position vs. time curves obtained by tracking a Janus particle (a = 2.4 µm) above a glass interface with E =

181

2.5 V/cm in 1 mM NaCl. E pointed in the negative x direction first and then switched every second. The red circles are times when E changed directions. (c) Typical  vs. time curve for rotation of a Janus colloid (2.5 V/cm, 5 % H2O2, 1 mM NaCl). We changed the direction of E after the particle aligned with the applied field. (d) f() vs. t for (b), from equation 21 (see Methods). (e, f) show still frames from a fluorescence microscopy video for a Pt-PS Janus particle rotating about an out of plane axis (e) and about an axis parallel to the plane (f) from the point at which the applied E- field polarity was reversed to the depicted direction (red arrow). (g) Measured zeta potentials for both Pt (blue markers) and PS (red markers) at two time points following sample preparation, each with and without hydrogen peroxide.

Next, we measured the difference between the Pt and PS sides of the particle. Owing to unequal zeta potentials at each end, the particle undergoes rotational electrophoresis to align the zeta

32 potential dipole with E (Figure B-3 c-f); see Methods . The contribution of complex convective flows to rotation is negligible in the presence of electric field as can be inferred from the final steady positions of the dipole (Figure B-3 c). We tracked particle rotations under fluorescence microscopy and used the measured angular velocity to obtain the zeta potentials for each case

(Figure B-3 d and g). The zeta potential of the PS end does not change appreciably upon the addition of fuel. We also find that the zeta potential of the Pt end stabilizes to a constant value, after about an hour in H2O2. While we observe some degree of sample-to-sample variability, based on the above measurements, we can largely conclude that the values of the zeta potentials of the two ends change by a small amount in relative terms due to the non-equilibrium catalytic activity. Hence, we can safely rule out electrostatic interactions as the main cause of rotational quenching.

As shown in Figure B-3 b, we see typical translational electrophoresis motion of a Janus particle (a

= 2.4 µm) in a 2.5 V/cm E-field in the direction of the E-field. The average zeta potential (J) was interpreted using the Smoluchowski equation. Averaged over many particles, we found an average of J = -87  5 mV (in 1 mM NaCl) with no fuel present, and J = -98  6 mV with 5 % H2O2 present (Figure B-3 g, and Table B-1).

182

Table B-1: Measured values of zeta potential of the Janus particle, platinum and polystyrene halves of the Janus particle in water and 5% H2O2 at time 10 mins and 60 mins. This table shows the actual values of the zeta potential measured for each particle and the averages for each trial, referenced in Figure 3g of the manuscript. The averages indicate very good consistency within a batch and the intra-batch variability is similar for trials with and without H2O2. Thus it is clear that the method of subtracting velocities of tracer particles in the presence of bubbles is effective in removing the effect of complex convective flows around a bubble.

J  Pt - PS  Pt  PS 

5 % H2O2, t = 10 mins -83.1 5.0 38.4 2.6 -63.9 5.2 -102.3 5.2 -79.5 5.5 38.5 2.2 -60.3 5.6 -98.8 5.6 -93.9 5.3 32.2 5.9 -77.8 6.0 -110.0 6.0 -90.1 5.4 36.3 6.4 -71.9 6.3 -108.3 6.3 -84.2 5.3 42.9 5.5 -62.7 6.0 -105.6 6.0 Averages -86.2 5.3 37.7 4.5 -67.3 5.8 -105.0 5.8

5 % H2O2, t = 60 mins -99.5 5.8 15.9 4.6 -91.5 6.3 -107.4 6.3 -98.4 4.9 22.8 2.8 -87.0 5.1 -109.8 5.1 -97.8 5.2 20.3 6.5 -87.7 6.1 -108.0 6.1 -98.3 6.1 27.2 5.7 -84.7 6.7 -111.9 6.7 -95.0 5.3 Averages -97.8 5.4 21.5 4.9 -87.7 6.0 -109.3 6.0 Water, t = 10 mins -89.1 4.6 14.2 2.1 -82.0 4.7 -96.2 4.7 -87.1 6.6 20.4 6.9 -76.9 7.4 -97.2 7.4 -89.1 5.6 12.5 1.2 -82.8 5.7 -95.3 5.7 -92.4 4.8 10.6 1.6 -87.2 4.9 -97.7 4.9 -90.1 6.3 Averages -89.5 5.6 14.4 2.9 -82.2 5.7 -96.6 5.7 Water, t = 60 mins -88.2 4.4 14.6 4.0 -80.9 4.8 -95.5 4.8 -87.7 5.2 13.7 2.5 -80.9 5.3 -94.6 5.3 -89.7 5.3 16.3 2.4 -81.5 5.4 -97.8 5.4 -86.3 5.3 14.4 2.7 -79.1 5.5 -93.5 5.5 -85.3 4.6 13.9 0.4 -78.3 4.6 -92.2 4.6 Averages -87.4 4.9 14.6 2.4 -80.1 5.1 -94.7 5.1

183 B.3.5 Phenomenology.

Driven motion of a sphere through viscous fluid near a solid substrate that imposes no-slip boundary condition introduces coupling between translation and rotation. We can get an intuitive idea of why this coupling comes about by thinking of a singular limit of the problem where a solid sphere in no-slip contact with a substrate is pulled by an external force; the sphere rolls along the surface. In the case of a sphere of radius a that is moving above a substrate at the (closest) distance of ℎ (See Figure B-4), the asymmetric shear stresses around the sphere will lead to rotation in addition to translation when the sphere is pulled through by an external force parallel to the wall.

While this is a comparatively small effect when h ≫ a, for small gap sizes ℎ ≤ 푎 the resulting translation and rotation are of the same order of magnitude and the coupling cannot be ignored.

For simplicity, we consider a reduced two dimensional version of the problem in which both translation and rotation of the sphere are restricted to the xz-plane, where the x-axis corresponds to the direction of the translational motion and the z-axis is perpendicular to the surface of the substrate

(See Figure B-4 a). The orientation of the sphere is described by the angle 휃 that the director of the

Janus particle makes with the z-axis.

Figure B-4: Schematics of the colloidal swimmer near a substrate, showing the geometry (a) and the surface slip velocity flow field (b) indicates the direction of propulsion.

184 The stochastic motion of the colloid, which is under the influence of surface interactions with a potential 푈(푧, 휃) as well as gravity, can be described via the following Langevin equations

푑푥 = 푉 + 푀 퐹 + 푀 [∓훾푚푔 푎 sin휃 − 휕 푈 + 푇 ] , (1) 푑푡 푥 푥푥 푥 푥휃 휃 푦

푑푧 = 푉 + 푀 [∓푚푔 − 휕 푈 + 퐹 ] , (2) 푑푡 푧 푧푧 푧 푧

푑휃 = Ω + 푀 퐹 + 푀 [∓훾푚푔 푎 sin휃 − 휕 푈 + 푇 ], (3) 푑푡 푦 휃푥 푥 휃휃 휃 푦

where 푉푥(푧, 휃) and 푉푧(푧, 휃) are the corresponding components of the spontaneous translational velocity of the particle, Ω푦(푧, 휃) is the corresponding spontaneous angular velocity, and 푀훼훽(푧, 휃)

33 are the relevant mobility coefficients. Note that 푀푥휃 = 푀휃푥]. The expression for gravitational torque contains 훾 that is a numerical constant of order unity that depends on the details of the weight distribution across the Janus particle (caused by the metallic cap). The choice of sign depends on the direction of gravity. Finally, 퐹푥, 퐹푧, and 푇푦 are noise terms.

The Langevin formulation can be used to construct an alternative description via the probability distribution 풫(푧, 휃, 푡) that satisfies the Fokker-Planck equation 휕푡풫 + 휕푥풥푥 + 휕푧풥푧 + 휕휃풥휃 =

0, where the fluxes are defined as

풥푥 = 푉푥풫 + 푀푥휃[∓훾푚푔 푎 sin휃 − 휕휃푈]풫 − 푀푥휃푘퐵푇휕휃풫, (4)

풥푧 = 푉푧풫 + 푀푧푧[∓푚푔 − 휕푧푈]풫 − 푀푧푧푘퐵푇휕푧풫, (5)

풥휃 = Ω푦풫 + 푀휃휃[∓훾푚푔 푎 sin휃 − 휕휃푈]풫 − 푀휃휃푘퐵푇휕휃풫. (6)

The x dependence is eliminated due to translational symmetry.

185 We take a pedagogical approach, and aim to construct an approximate solution to the above equation, so that we can highlight the key physical features. In stationary state, 푉푧 and 휕푧푈 will have relatively weak dependence on 휃 . We can obtain a stationary height 푧푠 = 푎 + ℎ by setting 풥푧 ≃ 0, from the balance between the interaction with the surface, gravity, and the average vertical component of the self-propulsion, possibly determined by the gravitational angular deflection due to the bottom-heaviness of the Janus particle. In the vicinity of the stationary height, we can find an approximate expression for the 휃 dependence of the stationary distribution by setting 풥휃 ≃ 0:

1 휃 훾푚푔푎 1 풫(푧푠, 휃) ≃ 풫(푧푠, 0) exp { ∫ 푑훼 Ω푦(푧푠, 훼) ∓ (1 − cos휃) − [푈(푧푠, 휃) − 푈(푧푠, 0)]}. (7) 푘퐵푇푀휃휃 0 푘퐵푇 푘퐵푇

1 휃 In the above equation, − ∫ 푑훼 Ω푦(푧푠, 훼) acts as an effective potential that represents the 푀휃휃 0 nonequilibrium activity of the colloid due to the catalysis.

We know from the experiments that the gravitational term has a small effect and can be ignored for most colloids (except for the largest radius). The equilibrium surface interaction potential 푈(푧푠, 휃) has contributions from different sources such as the electrostatic interactions due to the zeta potential difference across the Janus particle. The experiment tells us that the strong orientation quenching only exists when there is catalytic activity, which suggests that this term too can be neglected as compared with the nonequilibrium component. Therefore, we can write a simplified form for our approximate expression:

1 풫(푧푠, 휃) ≃ 풫(푧푠, 0) exp { ∫ 푑훼Ω푦(푧푠, 훼)}. (8) 푘퐵푇푀휃휃

The experimental observations are consistent with the phenomenological Hookean form of

Ω푦(휃) = −Γ(휃 − 휃푠), (9)

186

휋 for the angular velocity, where 휃 ≃ gives the stationary orientation and Γ > 0 acts as an 푠 2 effective (restoring) elastic constant. Based on the observation that the quenching effect increases with increasing fuel concentration, as well as dimensional grounds, the parameter Γ depends on the particle velocity and the radius of the sphere as

푣 Γ = 퐵 , (10) 푎 with the dimensionless prefactor 퐵 depending solely on the material and geometric parameters, such as zeta potential, height, and radius (see below).

The gravitational potential and the equilibrium surface interaction potential contributions in Eq. (7) will shift the equilibrium angle 휃푠 by a small amount. Putting Eq. (9) back in Eq. (8), we obtain

Γ 2 풫(푧푠, 휃) ≃ 풩 exp {− (휃 − 휃푠) }, (11) 2푘퐵푇푀휃휃 where 풩 is a normalization constant.

We now go back to the Langevin equation for angular fluctuations Eq. (3) and use it to calculate the mean-squared angular displacement (MSAD). The correlations of the noise terms defined via the generalized friction coefficients as follows

〈퐹푥(푡1)퐹푥(푡2)〉 = 2푘퐵푇푍푥푥훿(푡1 − 푡2), (12)

〈퐹푥(푡1)푇푦(푡2)〉 = 2푘퐵푇푍푥휃훿(푡1 − 푡2), (13)

〈푇푦(푡1)푇푦(푡2)〉 = 2푘퐵푇푍휃휃훿(푡1 − 푡2), (14) can be related to the mobility coefficients through the following identity

푀 푀 푍 푍 1 0 [ 푥푥 푥휃] [ 푥푥 푥휃] = [ ], (15) 푀휃푥 푀휃휃 푍휃푥 푍휃휃 0 1

187 due to the fluctuation-dissipation theorem. By solving the Langevin equation, we find the MSAD as

퐷 〈∆휃(푡)2〉 = 푟 [1 − 푒−2Γ푡], (16) Γ where the Einstein-Stokes relation 퐷푟 = 푘퐵푇푀휃휃 is used for the rotational diffusion coefficient to ensure the short time asymptotic behaviour of the above expression that gives a linear time dependence has the appropriate slope as expected for rotational Brownian motion.

Values of B can be obtained experimentally by plotting the averaged and re-scaled rotational

MSAD data as a function of time (Figure B-4). Data collapsed for a large number of individual trajectories at each particle size gives values of B of order unity.

1.5 a=1 m B=0.6 a=1.55 m B=1.05 1.0 a=2.4 m

B=1.65

r

MSD.v/aD 0.5

0.0 0.0 0.5 1.0 1.5 tv/a

Figure B-5: MSAD data re-scaled to allow comparison with theory, together with fits to equation 16 with estimated values for B (see equation 10). Mechanism. The parameter Γ that quantifies the degree of orientational quenching exhibits a dependence on the nonequilibrium catalytic activity that is controlled by the net propulsion velocity.

We now examine different mechanisms that could describe the observed rotational diffusion

188 quenching, and show that they lead to the behaviour described by Eqs. (10) & (16), and estimate their corresponding contributions to B.

When a sphere is pulled by a mechanical force along the horizontal axes, the presence of the surface induces a coupling that makes it roll (i.e. Ω푦 > 0 ). Phoretic transport of colloids with uniform

34 surface properties has been shown to exhibit a peculiar anti-rolling behaviour (i.e. Ω푦 < 0) . In self-phoretic propulsion9, a force-free and torque-free Janus colloid takes advantage of gradients that are generated due to the asymmetric activity across its surface to generate phoretic motion35 .

When a Janus particle self-propels in the vicinity of a surface, the coupling could lead to a mixture of rolling and anti-rolling tendencies, depending on the geometry and the type of activity36,37. This provides a promising starting point as it might be possible to find a specific combination that could lead to a fixed point as described by Eq. (9) and find the criterion for its stability, i.e. Γ > 0. The effect combines contributions from the phoretic and hydrodynamic interactions with the surface.36

The problem cannot be solved exactly in three dimensions, and a purely computational approach might not be particularly illuminating with regards to the exact mechanism that can explain the observed quenching. We use an alternative approach and estimate the contributions of various possible mechanisms using asymptotic approximations and scaling arguments.

The hydrodynamic flow field generated by colloidal particles that are transported via phoretic mechanisms are most commonly described by sources and sinks on their surfaces, in the form of a slip velocity v푠 rather than prescribed forces.38 For such a system, we can employ a far-field approximation, which has been shown to account for the hydrodynamic effect with high accuracy

28-30 even in the vicinity of the surface, to show that Ω푦 behaves as Eq. (9) with

9 Γ = ∫ 푑퐴 푛 푣푠, (17) ℎ 64휋(푎+ℎ)3 푥 푥

189 to the leading order (see Figure 4b). Equation (17) suggests that in order to have stable orientational quenching (i.e., Γ > 0) for spherical self-phoretic swimmers, two very specific criteria need to be met. The first criterion comes from symmetry: any surface slip velocity profile that is fore-aft symmetric will result in Γ = 0. This implies, for example, that the diffusiophoretic component of the surface slip velocity that is fore-aft symmetric cannot lead to the observed quenching of orientation. This type of surface slip velocity profile is known to lead to hydrodynamic flow field that decays as 1/푟2 , which leads to a stronger hydrodynamic coupling than the fore-aft symmetric contribution that decays as 1/푟3 representing a symmetric source-dipole.39 The second criterion involves the direction of swimming versus the catalytic coating; Γ > 0 only when the Janus particle swims away from the catalytic patch. The main motility mechanism of our platinum-coated Janus particles involves proton current loops that emanate from the vicinity of the equator and end near the pole25, hence satisfying both of these criteria by serendipity. For the sake of presentation, we

휋 use a representative velocity profile as v푠 = 푣 (1 + cos휗)(−cos휗) t̂ for < 휗 < 휋 (see 0 2

Figure 4b), which should provide a very good approximation to the exact velocity profile, and find

2 2 9 2휋푎 휋 푠 9 휋 2 푎 푣0 Γ = 휋 sin휗 푑휗 푛 (휗)푣 (휗) = ( − ) , (18) ℎ ( )3 ∫ 푥 푥 ( )3 64휋 푎+ℎ 2 32 16 15 푎+ℎ

which is positive. In Eq. (18), 푣0 is a characteristic velocity scale and t̂ is the unit tangent vector along the direction of increasing 휗. To relate 푣0 to the net propulsion velocity 푣 , we can use the following approximate expression

1 푠 1 휋 푠 1 1 휋 푣 ≃ − 2 ∫ 푑퐴 푣푥 = − ∫휋 sin휗 푑휗 푣푥 (휗) = ( − ) 푣0, (19) 4휋푎 2 2 2 3 16

푣 which should give us a reasonable estimate. Combining the above results, we find Γ = 퐵 , ℎ ℎ 푎

휋 2 ( − ) 3 3 9 16 15 푎 푎 where 퐵ℎ ≃ 1 휋 ( ) = 0.258765 ( ) . This result is in the expected form of Eq. 16 ( − ) 푎+ℎ 푎+ℎ 3 16

190

(10). We expect the prefactor 퐵ℎ to be stronger than the above estimate when the colloid is in close proximity of the surface. This is consistent with the experimental observations that give out fitted values for B that are of order unity. Recent observations on the motion of catalytic Janus spheres inside the matrix of a colloidal crystal reported by Poon et.al.23 are consistent with the existence of such a strong hydrodynamic coupling arising from a 1/푟2 decay due to the breakdown of the fore-aft symmetry in the surface slip velocity profile, as described above.

Figure B-6: Schematics of the colloidal swimmer near a substrate, showing the field lines created by the image

distribution and 퐄퐢퐦퐚퐠퐞 at the location of the Janus sphere (a), and the real electric field lines (b).

We know that the self-propulsion in our Janus particles is predominantly controlled by electrokinetic effects25. Therefore, it is natural to expect to have a significant electric field that results from the surface proximity of the proton currents caused by the catalytic activity of the Janus particle; the so-called “image field” (see Figure B-6). Due to the difference in the zeta potentials between the two halves of the Janus particle, the image field could contribute to the aligning tendency of the Janus particle via an electrophoretic contribution. In the bulk, we can relate the propulsion velocity to the self-generated electric field (caused by proton currents) using the

Smoluchowski equation, as 푣 ∼ 휖휁퐸̅ 푠푒푙푓/휂, where 휁 ̅ is the average zeta potential of the two halves

191 of the Janus particle. The presence of the surface will modify the proton currents and the resulting electric field, described by the image distribution. The electric field caused by the image at the location of the Janus particle, will create an angular velocity as given in Eq. (9) with the coupling Γ푒푙 ∼ 휖Δ휁퐸𝑖푚푎푔푒/(휂푎) , where Δ휁 is the difference between the zeta potentials of the two halves of the Janus particle. Considering the geometry as can be seen from Figure 6, we can relate

푎 3 the strengths of the electric fields as 퐸 ∼ 퐸 ( ) due to their dipolar nature. This can 𝑖푚푎푔푒 푠푒푙푓 푎+ℎ

푣 ∆휁 푎 3 be combined with the above results to give the estimate Γ = 퐵 , where 퐵 ~ ( ) . 푒푙 푒푙 푎 푒푙 휁 푎+ℎ

Therefore, we conclude that our estimate of the electrophoretic contribution Γ푒푙 is sub-dominant

(though not negligible), as it is smaller than our estimate of the hydrodynamic contribution Γℎ by a

∆휁 factor of < 1. Note that this statement uses the parametric form of the expressions we have 휁 obtained using scaling arguments, and thus the numerical prefactors in both estimates, which are likely to be comparable, are ignored for this comparison.

We can also estimate the contribution from electro-osmotic flows generated by the electric field in the vicinity of the wall. Unlike the case of electro-osmosis under externally applied electric field, the electric field generated by the proton fluxes is localized in the vicinity of the colloid only. The magnitude of this electric field near the wall will be given by 퐸𝑖푚푎푔푒 up to a numerical pre-factor

(see Figure B-6), and it will give rise to a localized surface-bound electro-osmotic flow of

푣푒표 ~ 휖휁푤퐸𝑖푚푎푔푒/휂 . This expression can be rewritten in terms of the propulsion velocity as 푣푒표 ∼

휁 푎 3 푣 푤 ( ) , which can then fed back to our calculation of Γ in Eq. (9) as an additional contribution 휁 푎+ℎ

푎2푣 휁 푎 6 to the surface slip velocity, which yields Γ ∼ 푒표 and 퐵 ∼ 푤 ( ) . Therefore, we 푒표 (푎+ℎ)3 푒표 휁 푎+ℎ

192 conclude that our estimate of the electrophoretic contribution Γ푒표 is smaller than our estimate of

휁 푎 3 the hydrodynamic contribution Γ by a factor of 푤 ( ) ≪ 1 within our approximation scheme. ℎ 휁 푎+ℎ

Finally, we estimate the contribution from self-diffusiophoresis. In the style of the above estimates,

푠 diffusiophoresis can be represented by a slip velocity of the form 푣 ~휇 ∇||퐶 where 휇 is a characteristic diffusiophoretic mobility and ∇||퐶 gives the tangential gradient of a relevant concentration field. In a realistic description of the reaction scheme we will need to use an algebraic sum of such terms representing each constituent of the reaction.38 For half-coated Janus configuration, an angular velocity is proportional to the difference between the the mobilities on the two halves ∆휇 and the relevant component of ∇||퐶 will be due to the image of the source/sink

푎 2 with respect to the impenetratable wall, which is a factor of ( ) smaller than the local ∇ 퐶 푎+ℎ || that is responsible to self-diffusiophoretic propulsion.37 Using this scale for the slip velocity in Eq.

∆휇 푎4푣 ∆휇 푎 5 푣 (17) gives the self-diffusiophoretic contribution as Γ ∼ 푠푑 and 퐵 ∼ ( ) 푠푑. 푠푑 휇 (푎+ℎ)5 푠푑 휇 푎+ℎ 푣

Therefore, we conclude that our estimate of the electrophoretic contribution Γ푠푑 is smaller than our

∆휇 푎 2 푣 estimate of the hydrodynamic contribution Γ by a factor of ( ) 푠푑 ≪ 1 within our scaling ℎ 휇 푎+ℎ 푣

푣 ∆휇 argument, since we know25 that 푠푑 ~ 0.1 and < 1. 푣 휇

The above perturbative analysis of the three effects shows that the unique quenching behaviour is a result of the hydrodynamic interaction of a force-free, self-electrophoretically propelled, particle with the wall and the broken symmetry of the surface flow field. It is the asymmetric slip velocity profile that differentiates these metal-insulator particles from other active particles propelled by self-electrophoresis such as bimetallic Janus particles as well as self-diffusiophoretic Janus spheres, where the slip velocity is symmetric and quenching by this mechanism is not expected. Moreover,

193 if the surface interactions are modified such that the Janus sphere swims with its platinum cap forward, then this mechanism cannot lead to stable orientational quenching. These spherical Janus swimmers are a special class, in their ability to be steered by boundaries (see discussion in section

B.5). It is hoped that in this context rotational confinement will be investigated in the future in a wider range of active colloid systems. For example a bi-metallic Janus particle41 has been reported, and fluorescent labelling could allow the above prediction to be tested.

Theoretically, we can estimate the mean residence time of the active colloids by roughly defining escape as jumping over the relevant barrier in the effective nonequilibrium potential. Although we have only calculated such a potential for small angular deviations, we can assume that the full potential will be a combination of trigonometric functions with the quadratic small angle limiting form. This suggests that the overall strength of the potential will still be controlled by Γ. Then we

Γ 8휋휂퐵푣푎2 −1 퐷 −1 푘 푇 can roughly estimate a mean residence time in the form of 휏res ≈ 퐷푟 푒 푟 = 퐷푟 푒 퐵 . Using

1 휇푚 the experimentally determined size dependence of the velocity 푣 ≈ ( ) ×10 휇푚/푠,40 we find 푎 that the residence time effectively depends on the exponential of the size (a weak dependency) and is significantly longer than the rotational diffusion time for 푎 > 50 푛푚.

B.4 Conclusion

We have shown that geometric constraints can help steer our Janus particles, by lowering the effective dimensionality of the space on their trajectories. One surface constrains the motion in 2D, and two perpendicular surfaces constrain it to 1D. This can be further exploited to create more elaborate constraints. For example, we expect a corner with three perpendicular surfaces to act as a trap and fully constrain the motion of our active colloids.

194 The ability to steer Janus motor particles uni-directionally along complicated trajectories by simply following an edge or groove opens the door for many transport and separation tasks such as directed cargo delivery, motility-based sorting, and flow-free microfluidics. Feature-directed steering combined with single-particle42 and collective many-body43 chemotactic response could provide an ideal toolkit for designing novel strategies to be employed in oil or mineral exploration tasks.44

Also as evident from Figure B-2 e this is a useful method for reversible bottom-up assembly of active colloids and gives us a unique advantage in controlling the orientation of the assembled active Janus particles. One further possibility, opened up by understanding the mechanism, is that in addition to “geometric railroad tracks” we could take advantage of chemical patterning of the surface to achieve a higher degree of control when we guide particles along pre-determined paths.

As a final comment, nature frequently produces motion at micron scales using moving parts to displace fluids, and in these systems hydrodynamic interactions between moving cellular structures and interfaces also leads to confinement at interfaces.28,45 Here we have shown that orientation confinement for phoretic synthetic active colloids with no moving parts can result in phenomenologically equivalent behaviour.

B.5 Discussion: exploring the limits for boundary steering phenomena

For boundary steering to be exploited for applications, it is important that active colloids retain the ability to be directed or confined at boundaries for sufficient durations to enable transport over useful distances. Encouragingly in this context, during our experiments, which involved repeated observations of hundreds of randomly selected particles for periods of several minutes, we observed very few cases where a particle would detach from a planar interface, or cease to be steered by a multi-planar geometry. Indeed we have shown that as a result micron sized active colloids can be steered by boundaries over macroscopic cm length scales. Because of this persistence of motion it

195 is experimentally difficult to investigate the frequency of “detachment” events where active colloids remove from a boundary or a plane. However, by inspection it appears that the residence time for a given active colloid will depend on the degree of rotational confinement, and the consequences of temporary misalignment. Our theoretical analysis and experimental data has indicated that at slower propulsion speeds rotational quenching is diminished, and so we can expect a higher frequency of events where the active colloid is no longer completely orientated parallel to the nearby interfaces. Therefore we may predict lower residence times for slower moving particles.

However, it is also clear that momentary misalignment may not necessarily lead to a particle separating from the interface. In some cases the misalignment may cause the propulsion vector to point towards the solid interface, which may not lead to detachment. Even if the propulsion vector is orientated away from the steering/confining interfaces, this may not generate sufficient force to result in the active colloid returning to “bulk”. This specifics of this case will depend on the geometry and size of the colloid, for example, for a colloid to leave the lower planar surface of a container will require it to produce sufficient upwards propulsion velocity to overcome its tendency to sediment, a factor that will actually encourage low velocity colloids to remain at the surface despite more frequent misalignments. Another consideration is stochastic variations in distance of the colloid from the constraining interface. Our theoretical analysis shows that the pre-factor determining the strength of orientational quenching has an inverse cubic relationship with separation from the wall. This leads to the possibility of Brownian “kicks” randomly translating the active colloid sufficiently far from the interface that orientational quenching is no longer experienced. Our experiments confirm that these events do not happen frequently at the sizes we have investigated. However, this effect could become limiting at smaller colloidal sizes, as

Brownian translational diffusion rate increases.

While this general discussion applies to all the geometries considered, we also present a more detailed consideration of the differences between the single plane active colloid arrangement which

196 we subjected to rigorous analysis, and the multi-plane arrangements. It was our hypothesis that adding an additional plane or planes beyond the single plane system would display the same hydrodynamic confinement phenomena, however now with additional axis of rotation being quenched, resulting in boundary steering. In comparison to the single planar system, the arguments we developed for hydrodynamic rotational confinement being a dominant factor in the system remain valid: the substrate material was unchanged for experiments conducted at an “edge”, and silicon substrates were used for the groove experiments which have also similar properties.

Consequently the electrostatic interactions are expected to stay the same. We observed that all sized particles moving within ten percent hydrogen peroxide solutions experienced prolonged boundary steering with long residence times at the edge formed between either the upper or lower planar cuvette surface and the vertical cuvette sidewall. This steering effect also tolerated the moderate curvature of the steering boundary found at the corners of the cuvette. As reported for the single plane case, the fluorescence intensity of the active colloid remained unchanged during confinement, and a constant “phase of the moon” was observed following the boundary direction and correlating with the expected propulsion direction away from the platinum cap. This shows that the boundary steering did correlate with the active colloid orientation confinement relative to both planes. For the case of the grooved lithographic example, the additional constraint that the colloid could physically fit within the groove was imposed, which in practise resulted in active colloids of a given size entering grooves that were approximately 2 µm wider than their diameters.

Due to optical reflections within the groove, it was not possible to determine the orientation of the active colloid as it followed a groove, however persistent propulsion motion along the groove was observed until the colloid reached the end at which point it became stuck. At this point microscopy did allow the orientation of the colloid to be viewed, and this was found to be aligned as expected relative to both boundaries. As mentioned in the main text, the groove system possesses the

197 potential to prevent Brownian kicks from allowing smaller colloids to escape orientational confinement, which could occur if only 2 planes are used to steer motion.

B.6 Methods

B.6.1 Janus Particle Preparation

Catalytic Janus spheres were prepared by spin coating a 0.1 wt % dispersion of fluorescent polystyrene microspheres (Themoscientific – Radius =1 µm, 1.55 µm and 2.4 µm) from ethanol onto freshly cleaned glass microscope slides. A 10 nm thick layer of platinum (>99.9 % Sigma

Aldrich) was then evaporated onto one side of the microspheres under vacuum in a Moorfield (U.K.)

Minilab 80 e-beam evaporator. Low volume fraction solutions of colloids in pure water

(Resistivity>15 MΩ) or the specified aqueous concentration of Hydrogen Peroxide (>99.9% Sigma

Aldrich) were prepared by removing the colloids from the glass slides by either ultra-sonication (as discussed in B.6.5) or physical transfer onto lens tissue and re-suspension.

B.6.2 Sample Preparation and Gravitaxis of Janus Colloids

Sample preparation: When Janus colloid solution are evenly dispersed into a low volume rectangular section glass cuvette (Suprasil glass cuvettes from Hellma Analytics) containing water, sedimentation leads to accumulation at the bottom planar surface. However, in the presence of hydrogen peroxide fuel, a propulsion velocity due to asymmetrical catalytic fuel decomposition is introduced. With fuel, radius=1 µm Janus colloids particles now rapidly arrive at both the top and bottom interior planar surfaces of the cuvette; whereas for radius=1.55/2.4 µm colloids,

198 accumulation preferentially occurs at the top interface. This biased accumulation for the larger colloids is due to the asymmetrical mass distribution caused by the dense platinum hemisphere resulting in gravitaxis.19 For all the sizes of colloids measured, microscopic examination of the

Janus particles at the interior cuvette surfaces, revealed persistent 2D enhanced in plane motion at the interfaces. Furthermore, if the cuvettes are inverted, the enhanced in plane motion continued, despite the reversal in gravitational field direction. Consequently observations for many catalytic

PS-Pt Janus colloids moving in 2D along planar interfaces could be made in two gravitational orientations (+g at the top surface, and –g for the bottom, Figure 1c of the manuscript). It is interesting to note that inactive colloids sediment away from the top surface, whereas the active fuelled colloids investigated here do not, and remain at the top interface. No significant changes in the surface-colloid interactions have been found on the addition of fuel to explain this effect.

Instead a small upwards component of propulsion sufficient for the colloid to maintain contact with the wall and prevent sedimentation back into the bulk solution may be present. A force balance approach can be applied to this scenario, and shows that e.g. for a=1 µm colloid, a modest cant angle towards the top surface (ϴ=86°) is sufficient to balance sedimentation. These small incline angles will be difficult to detect within our experimental ability to resolve ϴ in this geometry, however cant angles deviating from 90 ° were observed for confinement in multi-planar geometries

(See Supplementary Movie 4). This analysis suggests that reducing propulsion velocity and increasing particle size will require a greater cant angle to overcome sedimentation, and that also for sufficiently large colloids, sedimentation will become inevitable, as due to the (1/a) dependency on propulsion velocity there will not be sufficient force to overcome sedimentation irrespective of

ϴ. To illustrate this experimentally, we investigated some a=5 µm colloids, and found that for low fuel concentrations (approximately 1.5 %), sedimentation from the top surface was observed.

199 B.6.3 Microscopy and Image Analysis

The active colloids were observed using the fluorescence mode of a Nikon Eclipse LV100 microscope with illumination through blue excitation band of a Nikon B2A filter cube. Videos were captured using an Andor Neo camera at a frame rate of 33 Hz. Custom software developed using LabView vision identified the x,y coordinate for the centre of each bright particle throughout each video, thus generating trajectory data. Such trajectories can be quantified using Mean-Square

Displacement (MSD) versus time analysis to determine propulsion velocity as previously described.16

The software also determined the average pixel intensity for the central region of each tracked particle as a function of time. For a given particle size, illumination and camera settings were fixed, allowing quantitative comparison between these intensity values. To convert these relative fluorescence emission values to polar angles, it was necessary to establish the intensities that corresponded to cap down (θ = 0° maximum brightness) and cap up (θ = 180° minimum brightness) orientations. For a = 1 µm and a = 1.55 µm particles a self-calibration approach was used, where long duration intensity versus time plots were obtained for a number of particles in water, and the average maximum and minimum intensities were assigned to θ = 0° and θ = 180° polar angles. The intensity between these limits was assumed to vary sinusoidally to complete the conversion.27

However for a = 2.4 µm Pt-PS Janus particles the gravitational torque makes the probability of rotation to a cap up configuration too low to observe within a reasonable time frame. To overcome this, a = 2.4 µm colloids were “frozen” within a transparent gellan gum, allowing the now solid sample to be inverted. The minimum intensity found from observing many of these inverted particles was used as the θ = 180 ° limit, subsequently allowing conversion to polar angles as before.

All presented polar angle data was obtained in this way, however the data collapse shown in Figure

200 4 also constrained the short term behaviour to match the theoretical value for the rotational diffusion coefficient to further correct for particle to particle variations in fluorescence intensity.

B.6.4 Colloid settling and diffusion

Table B-2: Diffusion coefficients parallel to a wall (푫||) and the corresponding gap heights (풉) of a suspension of Janus colloids in both water and KNO3 solution at 21oC. Values in brackets are the standard deviation of the mean values.

2 -1 2 -1 a (m) Dm s ) D|| (m s ) ℎ (m) Water 퐼 = 8×10−5 M Water 퐼 = 8×10−5 M 0.95 0.232 0.163 (0.039) 0.082 (0.026) 0.860 0.030 1.55 0.142 0.095 (0.018) 0.046 (0.011) 1.068 0.040 2.40 0.092 0.042 (0.013) 0.039 (0.014) 0.379 0.240

A suspension of colloidal particles diffusing well away from a wall has a Brownian diffusion coefficient 퐷∞ described by the Stokes-Einstein relation (퐷∞ = 푘퐵푇⁄6휋휂푎). As the colloids settle under gravity and reach an equilibrium height above the bottom surface of their container, their diffusion is hindered by an additional hydrodynamic drag. We can use these deviations from the bulk diffusion coefficient for colloids moving near a wall to estimate the colloidal suspension height above a wall, as described by Kazoe.49 The hindrance factor for diffusion parallel to a surface is given by Eq. (20)

3 4 5 퐷|| 9 푎 1 푎 45 푎 1 푎 = 1 − ( ) + ( ) − ( ) − ( ) (20) 퐷∞ 16 푧 8 푧 256 푧 16 푧 where 퐷|| is the diffusion coefficient parallel to the surface and the gap height ℎ is contained in the

푧 = 푎 + ℎ term.50,51

To estimate the height of the particles above a fused quartz silica wall we suspended a mixture of green fluorescent Janus colloids (a = 0.95, 1.55 and 2.40 m) in water (Elga Purelab Option, 15

201

Mcm) and allowed them to settle out. With the Janus colloids at their equilibrium height above the bottom wall of the cell, videos of 25 individual particles of each size were recorded at a frame rate of 100 Hz (Andor Neo camera). Image analysis algorithms were used to extract 2D trajectories of the particles diffusing in the (x, y) plane and the mean squared displacements ∆퐿2 computed. A

2 fit to ∆퐿 yielded the hindered diffusion coefficient 퐷||.

As it is problematic to extract accurate values for Brownian diffusion coefficients for Janus colloids also undergoing significant propulsion, we use measurements on unfuelled colloids as a proxy for the active colloid parameters. The strongest H2O2 solution (Sigma Aldrich, 30.0 wt%, pKa = 11.75) that we add to a suspension of Janus sphere swimmer particles to observe swimming behaviour contains the stabiliser dipicolinic acid at a concentration of 40 mg L-1 (pKa = 2.2) giving an estimated ionic strength, 퐼 = 8×10−5 M, at 10 % wt. To allow for the expected screening effect in this solution we used a KNO3 solution with the same ionic strength. These results, together with those in pure water are shown in Table B-2.

B.6.5 Electrophoresis/Rotational Electrophoresis

Standard Dynamic Light Scattering (DLS) equipment is not designed to handle either the material heterogeneity of metal-insulator Janus particles or bubbling at the electrodes in the presence of

H2O2. Measurement of Pt and PS using separate homogeneously coated Pt and PS particles is possible, but a constant thickness Pt surface does not undergo the same reaction as a Janus particle with an inhomogeneous coating on the Pt half.25 Consequently, in order to measure the average colloid zeta potential, J, we carried out translational electrophoresis in a homemade closed capillary setup, as described below.

202 The platinum coated glass slide covered in Janus colloids (prepared as described above in B.6.1) was submerged in 25 mL DI water in a petri dish and sonicated for 30 minutes. The resulting dispersion was concentrated to 3 mL by centrifugation at 1000 g in a Sorvall Biofuge Primo centrifuge. 50 L of the concentrated Janus solution was diluted to 600 L using 1 mM NaCl solution containing 0.003 % by volume of sPSL tracers. For samples with H2O2, equal volumes of concentrated Janus solution and 30 % H2O2 were sonicated for 5 minutes and left for another 20 minutes to activate the Pt surface. Then they were diluted to 5 % H2O2.

The solution for analysis was fed into a cleaned glass capillary (0.9 mm square cross section, 50 mm long and 0.18 mm wall thickness from Vitrocom, RCA-I cleaned) and placed on a glass slide making sure there are no bubbles or air gaps. About 1.5 cm long piece of gold wire (0.5 mm, 99.99% purity Alfa Aesar, cleaned using water and ethanol) was inserted at each end and the capillary was sealed off using wax and a UV-curing adhesive. The capillary set up was mounted on the motorized stage of a Nikon TE 300 inverted microscope, so that the capillary surface faced the 20x or 40 x objective. The gold wires were connected to electrodes from a Keithley 2410 source meter which was operated in the constant current mode at 3 A resulting in an electric field of 2.5 V/cm.

A small number of sPSL particles were added as tracers for translational electrophoresis. Measuring against tracers with known zeta potential allows us to account for pressure driven flows in closed capillary and convective flows due to bubbles that form in-situ in the presence of H2O2.

To validate our technique, we measured the zeta potential of sPSL tracer particles using our setup and compared it to the measurements from the DLS based Malvern Zetasizer. To do so, we measured the translational velocity of particles in 8 different planes across the square capillary,

46 centred on the width and used Bowen’s equation to fit for spsl and wall using Mathematica (Figure

B-7). Zeta potential of sPSL particle (spsl),was measured using DLS (Malvern Nano ZS Zeta Sizer) as well as using our setup, which served to verify that our set-up did not introduce impurities to the

203 system. We measured the velocity of sPSL tracers in at least 7 different planes across the depth of the capillary. We fitted these velocities for w and spsl to the known parabolic flow profile due to electroosmotic flow in a closed capillary10. Limited available volume fraction and fast settling did not permit a similar experiment on Janus particles to measure zeta potential of the Janus particle

(J) directly.

Two problems arise due to the formation of bubbles. First, there is increased resistance to current flow but we managed to keep the electric field constant by applying a constant current through the capillary. Secondly, bubbles lead to convective flows in the capillary which we account for by measuring translational speeds relative to tracer particles with known zeta potentials. We have addressed the issue of bubbles on multiple fronts :

1) We measured the zeta potential () of sPSL (sulfated polystyrene latex), say in 1 mM

NaCl, both with and without H2O2. We wanted to test whether the H2O2 made a

difference on the zeta of ONLY the sPSL, with no Janus particles present. The data show

that the zeta potential of sPSL does not change in the presence of H2O2, to within the

statistical uncertainty of about 5%.

2) We then followed our Janus particles, using the sPSL as tracers. We did this away from

any visible bubbles. The particles had a smoothly linear motion, with no visible effect of

any complex flow fields that might have resulted from bubbles.

3) Knowing sPSL, we were able to assess Janus, including uncertainties. Importantly, our

zeta measurements in the presence of bubbles (5 % H2O2) give the same standard

deviations as in the absence of bubbles (no H2O2), up to 6 mV (Table B-1), indicating the

effectiveness of the subtraction of flows through tracer particles. This is a typical

deviation, similar to that obtained by standard methods.

204

(a)

Figure B-7: Calibration of cell for electrophoretic measurements. (a) Markers represent experimentally measured velocities of sPSL particles as a function of height from the centre. The solid line represents the best fit parabola using

spsl and wall as parameters in Bowen’s equation. (b) Shows the spsl and wall calculated from the fit in fig a for 2 independent trials, C1 and C2. The last row shows the spsl measured using Malvern Zetasizer. To determine the zeta potential for each side of the colloid we measured particle orientation changes during field direction reversals in the same set up. Using image analysis software it was possible to directly measure the orientation of the Janus colloid as a function of time for re-alignments that proceeded via an in plane rotation, Figure 4e. For a spherical Janus particle with potentials Pt and

푑 PS on each half, the predicted angular velocity, during a rotational event is a function of the 푑푡 angle  between the normal to the dipole moment (Pt - PS) 푒̂ and the electric field E∞32:

푑 9 휀 = ( −  ) 퐸 cos . (21) 푑푡 16 휂푎 Pt PS ∞

Integrating this equation with respect to  and time (푡) yields a function 푓() that varies linearly with 푡:

9 휀 푓() = ln|sec  + tan | = ( −  ) 퐸 푡 + 푐. (22) 16 휂푎 Pt PS ∞

205 Equations (21) and (22) predict that if the particles are aligned anti-parallel to the E-field, the rotation rate will be zero and time will be . This is an unstable equilibrium position. The slightest kick from the ever-present Brownian rotational motion will bring the particle to its stable equilibrium position where the dipole is parallel to the E-field. Thus flipping the polarity of E-field each time the particle reaches its equilibrium allows us to observe the rotation of the particle as it aligns with the E-field

We do a frame by frame analysis of at least 4 rotations induced by field-reversal within the same field of view and obtain  − 푡 curves (Figure B-3 c). Using a least-squares fit of the linear portions of corresponding) vs 푡 curves (Figure B-3 d), we calculate the value of Pt - PS for each particle observed. Together with the knowledge of J = (Pt + PS)/2, we are able to determine Pt and PS separately. Averages and standard deviations were obtained from the distribution of average ZP for each particle within a trial (Table B-1).

B.6.6 Lithographic Manufacture of Rectangular Channels

The channels were made using standard lithography techniques47. The rectangular wells were fabricated over silicon wafers in the Nanofabrication Laboratory of Materials Research Institute,

Penn State University. Silicon Wafers (4" wafer, 100 prime, 0.1 ohm cm conductivity and 500 µm thick) were cleaned with acetone and air-dried. The wafers were then spin-coated with 1 mL of

SPR 955 photoresist (Microposit) at 900 rpm for 10 sec and then at 3000 rpm for 60 sec. This was followed by soft-baking the coated wafers over a hot plate at 95 0C, for 60 sec. The well geometry was modelled in CAD and printed over a chrome-on-glass mask (Nanofabrication Laboratory,

Materials Research Institute, Penn State). For photolithography, the mask was placed in direct contact with the photoresist over the wafers. The resist was then exposed to UV radiation (Intensity

206 8 mW/cm2) for 8 sec in a Karl Suss MA/BA6 Contact Aligner. The exposed wafers were post- baked for 1 min over a hot plate at 95 0C to cross link the exposed film. MF CD26 developer was used to remove the unexposed resist SPR 955 from the wafers. The mould was developed for 90 sec while being agitated, followed by washing it thoroughly with deionized water. After the wafers were dried with a nitrogen blower, a 30 µm deep master pattern was created on them using Deep

Reactive Ion Etching to yield the rectangular wells on the silicon surface. The ion etching was done using Alcatel Silicon DRIE in Nanofabrication Laboratory of Materials Research Institute, Penn

State and the process used was Low ARDE. After etching the remaining resist was removed from the wafers by agitating the wafers in a solution of NanoRemover PG at 60 0C for 2 hours. As a final cleaning step, the wafers were cleaned with isopropanol, hexane and acetone for 30 min each and air dried. To homogenize the surface chemistry, the cleaned wafers were then treated with oxygen plasma (200 sccm, 400 W) for 30 min.

207

4.0

3.5 2um top 3.0 2um bottom 3um top 2.5 -1 3um bottom

5 um top ms

 2.0

5um bottom

1.5

1.0 Velocity / / Velocity 0.5

0.0

-0.5 0 2 4 6 8 10 [H2O2] %

Figure B-8: Relationship between particle size, fuel concentration and velocity. Average propulsion velocity versus fuel concentration. Supplementary Figure 1 shows the mean velocity as a function of hydrogen peroxide (H2O2) concentration for all of the Pt-PS (Platinum-Polystyrene) Janus particles investigated in section 1. “Top” refers to particles swimming at the top surface of the cuvette, and “Bottom” to particles sediment at the bottom of the cuvette.

B.7 Acknowledgements

The authors thank Prof. P. Cremer and Mathew Poyton for use and advice on their fluorescence microscopes. This work was funded by Penn State MRSEC under NSF grant DMR-1420620. This publication was supported by the Pennsylvania State University Materials Research Institute

Nanofabrication Lab and the National Science Foundation Cooperative Agreement No. ECS-

0335765. Dr Ebbens and Dr Campbell would like to thank support from Dr Ebbens’ EPSRC fellowship (EP/J002402/1) and Dr Ebbens further acknowledges a World Universities Network travel grant. Prof. Ramin Golestanian would like to acknowledge support from the EPSRC.

208 B.8 Contributions

Dr. Steve Ebbens and Dr. Jonathan Howse established the orientation quenching phenomena reported here, with Dr. Andrew Campdell assisting with later experiments. Dr. Ramin Golestanian developed the theoretical understanding and treatment of the data to rationalise these observations.

Dr. Ayusman Sen and Dr. Darrell Velegol, and associated researchers (Dr. Sambeeta Das and Astha

Garg) enabled experiments to further demonstrate and elucidate the reported behaviour. Dr.

Sambeeta Das performed the experiments regarding quenching behavior in lithographically produced structures. Astha Garg performed the experimental measurements of zeta potential. All authors assisted in manuscript preparation and data interpretation discussions.

B.9 References

(1) Baraban, L.; Makarov, D.; Streubel, R.; Mönch, I.; Grimm, D.; Sanchez, S.; Schmidt, O. G.

Catalytic Janus Motors on Microfluidic Chip: Deterministic Motion for Targeted Cargo

Delivery. ACS Nano 2012, 6, 3383–3389.

(2) Simmchen, J.; Baeza, A.; Ruiz, D.; Esplandiu, M. J.; Vallet-Regí, M. Asymmetric Hybrid

Silica Nanomotors for Capture and Cargo Transport: Towards a Novel Motion-Based DNA

Sensor. Small 2012, 8, 2053–2059.

(3) Sundararajan, S.; Lammert, P. E.; Zudans, A. W.; Crespi, V. H.; Sen, A. Catalytic Motors

for Transport of Colloidal Cargo. Nano Lett. 2008, 8, 1271–1276.

(4) Sen, A.; Patra, D.; Sengupta, S.; Duan, W.; Zhang, H.; Pavlick, R. Intelligent, Self-Powered,

Drug Delivery Systems. Nanoscale, 2012, 5, 1273–1283.

(5) Kagan, D.; Calvo-Marzal, P.; Balasubramanian, S.; Sattayasamitsathit, S.; Manesh, K. M.;

209 Flechsig, G.-U.; Wang, J. Chemical Sensing Based on Catalytic Nanomotors: Motion-Based

Detection of Trace Silver. J. Am. Chem. Soc. 2009, 131, 12082–12083.

(6) Soler, L.; Sánchez, S. Catalytic Nanomotors for Environmental Monitoring and Water

Remediation. Nanoscale 2014, 6, 7175–7182.

(7) Ebbens, S. J.; Howse, J. R. In Pursuit of Propulsion at the Nanoscale. Soft Matter 2010, 6,

726.

(8) Wang, W.; Duan, W.; Ahmed, S.; Mallouk, T. E.; Sen, A. Small Power: Autonomous Nano-

and Micromotors Propelled by Self-Generated Gradients. Nano Today, 2013, 8, 531–534.

(9) Golestanian, R.; Liverpool, T. B.; Ajdari, A. Propulsion of a Molecular Machine by

Asymmetric Distribution of Reaction Products. Phys. Rev. Lett. 2005, 94, 1–4.

(10) Wang, Y.; Hernandez, R. M.; Bartlett, D. J.; Bingham, J. M.; Kline, T. R.; Sen, A.; Mallouk,

T. E. Bipolar Electrochemical Mechanism for the Propulsion of Catalytic Nanomotors in

Hydrogen Peroxide Solutions. Langmuir 2006, 22, 10451–10456.

(11) Manjare, M.; Yang, B.; Zhao, Y. P. Bubble Driven Quasioscillatory Translational Motion

of Catalytic Micromotors. Phys. Rev. Lett. 2012, 109, 1–5.

(12) Paxton, W. F.; Baker, P. T.; Kline, T. R.; Wang, Y.; Mallouk, T. E.; Sen, A. Catalytically

Induced Electrokinetics for Motors and Micropumps. J. Am. Chem. Soc. 2006, 128, 14881–

14888.

(13) Solovev, A. a; Xi, W.; Gracias, D. H.; Harazim, S. M.; Deneke, C.; Sanchez, S.; Schmidt,

O. G. Self-Propelled Nanotools. ACS Nano 2012, 6, 1751–1756.

(14) Howse, J. R.; Jones, R. A. L.; Ryan, A. J.; Gough, T.; Vafabakhsh, R.; Golestanian, R. Self-

Motile Colloidal Particles: From Directed Propulsion to Random Walk. Phys. Rev. Lett.

2007, 99, 8–11.

210 (15) Ebbens, S. J.; Howse, J. R. Direct Observation of the Direction of Motion for Spherical

Catalytic Swimmers. Langmuir 2011, 27, 12293–12296.

(16) Dunderdale, G.; Ebbens, S.; Fairclough, P.; Howse, J. Importance of Particle Tracking and

Calculating the Mean-Squared Displacement in Distinguishing Nanopropulsion from Other

Processes. Langmuir 2012, 28, 10997–11006.

(17) Golestanian, R. Anomalous Diffusion of Symmetric and Asymmetric Active Colloids. Phys.

Rev. Lett. 2009, 102, 188305.

(18) Wang, J. Cargo-Towing Synthetic Nanomachines: Towards Active Transport in Microchip

Devices. Lab on a Chip, 2012, 12, 1944.

(19) Campbell, A. I.; Ebbens, S. J. Gravitaxis in Spherical Janus Swimming Devices. Langmuir

2013, 29, 14066–14073.

(20) ten Hagen, B.; Kümmel, F.; Wittkowski, R.; Takagi, D.; Löwen, H.; Bechinger, C.

Gravitaxis of Asymmetric Self-Propelled Colloidal Particles. Nat. Commun. 2014, 5, 4829.

(21) Gangwal, S.; Cayre, O. J.; Velev, O. D. Dielectrophoretic Assembly of Metallodielectric

Janus Particles in AC Electric Fields. Langmuir 2008, 24, 13312–13320.

(22) Ebbens, S. A Fantastic Voyage? Mater. Today 2012, 15, 294.

(23) Brown, A. T.; Vladescu, I. D.; Dawson, A.; Vissers, T.; Schwarz-Linek, J.; Lintuvuori, J.

S.; Poon, W. C. K. Swimming in a Crystal: Hydrodynamic, Phoretic and Steric Interactions.

arXiv:1411.6847 2015, 7.

(24) Choi, S. Q.; Jang, S. G.; Pascall, A. J.; Dimitriou, M. D.; Kang, T.; Hawker, C. J.; Squires,

T. M. Synthesis of Multifunctional Micrometer-Sized Particles with Magnetic, Amphiphilic,

and Anisotropic Properties. Adv. Mater. 2011, 23, 2348–2352.

211 (25) Ebbens, S.; Gregory, D. a.; Dunderdale, G.; Howse, J. R.; Ibrahim, Y.; Liverpool, T. B.;

Golestanian, R. Electrokinetic Effects in Catalytic Platinum-Insulator Janus Swimmers.

EPL (Europhysics Lett. 2014, 106, 58003.

(26) Brown, A.; Poon, W. Ionic Effects in Self-Propelled Pt-Coated Janus Swimmers. Soft

Matter 2014, 4016–4027.

(27) Behrend, C. J.; Anker, J. N.; McNaughton, B. H.; Brasuel, M.; Philbert, M. a.; Kopelman,

R. Metal-Capped Brownian and Magnetically Modulated Optical Nanoprobes

(MOONs): Micromechanics in Chemical and Biological Microenvironments †. J. Phys.

Chem. B 2004, 108, 10408–10414.

(28) Elgeti, J.; Kaupp, U. B.; Gompper, G. Hydrodynamics of Sperm Cells near Surfaces.

Biophysj 2010, 99, 1018–1026.

(29) Lauga, E.; DiLuzio, W. R.; Whitesides, G. M.; Stone, H. a. Swimming in Circles: Motion

of Bacteria near Solid Boundaries. Biophys. J. 2006, 90, 400–412.

(30) Spagnolie, S. E.; Lauga, E. Hydrodynamics of Self-Propulsion near a Boundary: Predictions

and Accuracy of Far-Field Approximations. J. Fluid Mech. 2012, 700, 105–147.

(31) Chiang, T.-Y.; Velegol, D. Localized Electroosmosis (LEO) Induced by Spherical Colloidal

Motors. Langmuir 2014, 30, 2600–2607.

(32) Anderson, J. L. Effect of Nonuniform Zeta Potential on Particle Movement in Electric Fields.

J. Colloid Interface Sci. 1985, 105, 45–54.

(33) Happel, J.; Brenner, H. Low Reynolds Number Hydrodyanamics; 2nd ed.; Kluwer Academic:

AH Dordrecht, 1973.

(34) Yariv, E.; Brenner, H. Near-Contact Electrophoretic Motion of a Sphere Parallel to a Planar

Wall. J. Fluid Mech. 2003, 484, 85–111.

212 (35) Golestanian, R.; Liverpool, T. B.; Ajdari, A. Designing Phoretic Micro- and Nano-

Swimmers. New J. Phys. 2007, 9, 126–126.

(36) Crowdy, D. G. Wall Effects on Self-Diffusiophoretic Janus Particles: A Theoretical Study.

J. Fluid Mech. 2013, 735, 473–498.

(37) Uspal, W. E.; Popescu, M. N.; Dietrich, S.; Tasinkevych, M. Self-Propulsion of a

Catalytically Active Particle near a Planar Wall: From Reflection to Sliding and Hovering.

Soft Matter 2014, 1, 1–15.

(38) Anderson, J. L. Colloid Transport by Interfacial Forces. Annu. Rev. Fluid Mech. 1989, 21,

61–99.

(39) Jülicher, F.; Prost, J. Generic Theory of Colloidal Transport. Eur. Phys. J. E 2009, 29, 27–

36.

(40) Ebbens, S.; Tu, M.-H. M.; Howse, J. R. J. J. R.; Golestanian, R.; Mei-Hsien Tu. Size

Dependence of the Propulsion Veloicty for Catalytic Janus-Sphere Swimmers. Phys. Rev.

E. 2012, 85, 20401.

(41) Wheat, P. M.; Marine, N. a.; Moran, J. L.; Posner, J. D. Rapid Fabrication of Bimetallic

Spherical Motors. Langmuir 2010, 26, 13052–13055.

(42) Baraban, L.; Harazim, S. M.; Sanchez, S.; Schmidt, O. G. Chemotactic Behavior of

Catalytic Motors in Microfluidic Channels. Angew. Chemie - Int. Ed. 2013, 52, 5552–5556.

(43) Saha, S.; Golestanian, R.; Ramaswamy, S. Clusters, Asters, and Collective Oscillations in

Chemotactic Colloids. Phys. Rev. E - Stat. Nonlinear, Soft Matter Phys. 2014, 89, 1–25.

(44) Kar, A.; Chiang, T.; Rivera, I. O.; Sen, A.; Velegol, D. Enhanced Transport into and out of

Dead-End Pores. 2015, 746–753.

213 (45) Sipos, O.; Nagy, K.; Di Leonardo, R.; Galajda, P. Hydrodynamic Trapping of Swimming

Bacteria by Convex Walls. Phys. Rev. Lett. 2015, 114, 258104.

(46) Bowen, B. D. Effect of a Finite Half-Width on Combined Electroosmosis - Electrophoresis

Measurements in a Rectangular Cell. J. Colloid Interface Sci. 1981, 82, 574–576.

(47) Dey, K. K.; Das, S.; Poyton, M. F.; Sengupta, S.; Butler, P. J.; Cremer, P. S.; Sen, A.

Chemotactic Separation of Enzymes. ACS Nano 2014, 8, 11941–11949.

(48) Crowdy, D. Wall Effects on Self-Diffusiophoretic Janus Particles: A Theoretical Study. J.

Fluid Mech. 2013.

(49) Kazoe, Y.; Yoda, M. Measurements of the near-Wall Hindered Diffusion of Colloidal

Particles in the Presence of an Electric Field. Appl. Phys. Lett. 2011, 99, 2009–2012.

(50) Goldman, a. J.; Cox, R. G.; Brenner, H. Slow Viscous Motion of a Sphere Parallel to a

Plane wall—II Couette Flow. Chem. Eng. Sci. 1967, 22, 653–660.

(51) Banerjee, A.; Kihm, K. Experimental Verification of near-Wall Hindered Diffusion for the

Brownian Motion of Nanoparticles Using Evanescent Wave Microscopy. Phys. Rev. E 2005,

72, 42101.

214

Appendix C

Origins of concentration gradients for diffusiophoresis4

C.1 Abstract

Fluid transport that is driven by gradients of pressure, gravity, or electro-magnetic potential is well-

known and studied in many fields. A subtler type of transport, called diffusiophoresis, occurs in a

gradient of chemical concentration, either electrolyte or non-electrolyte. Diffusiophoresis works

by driving a slip velocity at the fluid-solid interface. Although the mechanism is well-known, the

diffusiophoresis mechanism is often considered to be an esoteric laboratory phenomenon. However,

in this article we show that concentration gradients can develop in a surprisingly wide variety of

physical phenomena – imposed gradients, asymmetric reactions, dissolution, crystallization,

evaporation, mixing, sedimentation, and others – so that diffusiophoresis is in fact a very common

transport mechanism, in both natural and artificial systems. We anticipate that in georeservoir

extractions, physiological systems, drying operations, laboratory and industrial separations,

crystallization operations, membrane processes, and many other situations, diffusiophoresis is

already occurring – often without being recognized – and that opportunities exist for designing this

transport to great advantage.

4 Adapted from the article D. Velegol, A. Garg, R. Guha, A. Kar, M. Kumar, Origins of

Concentration Gradients for Diffusiophoresis. Soft Matter, 12, 4686 (2016), with permission from The Royal Society of Chemistry.

215

Keywords: diffusiophoresis, concentration gradients, natural systems, artificial systems, chemical processes

C.2 Introduction

Fluid flow occurs in the presence of various externally-applied potential gradients, including pressure, gravity, electric, and magnetic. Each of these gradients produces a direct force on a fluid or particle, which drives transport (Figure 1). For a Newtonian fluid, the resulting velocity field is described by the Navier-Stokes equations, giving for instance a parabolic velocity field in a tube due to a pressure gradient1, or a potential velocity field around a particle in electrophoresis2,3. There is, however, a more subtle and indirect transport mechanism, due to a chemical potential gradient. For a fluid this mechanism is called diffusioosmosis, and for a particle it is called diffusiophoresis; in this article we will use the word diffusiophoresis to refer to both, unless we need to specify.

Diffusiophoresis is the chemically-driven transport of colloidal particles driven by a concentration gradient of solute, without the application of any outside force.3 The mechanism has been recognized since 1947 when Derjaguin discovered it4. Diffusiophoresis converts the chemical energy of concentration gradients into the mechanical energy of colloids moving through fluids

(Figure 2), and it does so by producing an electric field or pressure gradient. Transport rates of the order of 1-10 m/s are typical. The physics of diffusiophoretic transport has been well-established both theoretically and experimentally5,6, and for several decades diffusiophoresis has often been considered to be an esoteric laboratory phenomena. The purpose of this article is to enumerate a

216 surprising variety of situations in which concentration gradients arise (Table 1), and which therefore can cause unanticipated diffusiophoretic transport. Whether we are examining a chemistry wet lab, a reverse osmosis membrane, the human body, or a geological reservoir, concentration gradients frequently exist or arise, and the resulting chemically-driven transport can potentially cause unsuspected transport. Recognizing diffusiophoresis, therefore, can also offer novel opportunities. The primary purpose of this paper is to identify and classify carefully the variety of physico-chemical and biochemical situations where diffusiophoresis can occur, so that researchers can identify situations where diffusiophoresis might cause flows, for better or worse.

Diffusioosmosis

Molecular Architecture

Figure C-1: Causes of fluid transport and interconversions of driving forces. These include well-known causes like pressure gradients, electrical or magnetic gradients, and gravitational gradients. However, chemically-driven flows, caused by chemical concentration gradients, are not widely-appreciated. The prevalence of chemical gradients, and thus chemical potential gradients, makes diffusiophoretic transport more widespread and pervasive than has been realized. Table C-1: Categories of concentration gradient origins. Some of the systems “known” to involve diffusiophoresis (DP) are recent discoveries (e.g., diffusive mixing). Also, for some of the categories, the authors are not aware of any known examples in the literature.

Origin of concentration Systems known to involve Systems predicted or Category gradient (c) diffusiophoresis observed situations

217

Dialysis/Diafiltration7 and reverse Particle exclusion zones, osmosis8 membranes, electrical Molecular One or more ions cannot access fouling of membranes, double layers (EDLs), DNA exclusion the same volume as the fluid. Ion exchange chromatography diffusiophoresis across biological for protein separation. membranes.

Reactive surface or particle Tracer particles form different Synthetic motors such as Janus produces a concentration gradient patterns according to 9 10 11 12 Chemical reaction particles , Ag , AgCl , Ag3PO4 due to asymmetry of reactivity, light substrate materials- emergent 13 and TiO2 . stimulus or shape. behavior.

Estuaries, flooding of fresh Solutions with different salt water in geological reservoirs, concentrations are allowed to mix Idealized pore experiments showing earthquakes, fracking, Diffusive mixing due to diffusion, including boundary flows in dead-end pores14. capillaries of microfluidic layer diffusiophoresis (BLDP). networks that are changed transiently.

Spreading and focusing of colloidal Manipulation of colloidal Ion diffusion in microchannels particles in microchannels15, entities in microchannels or in (transverse)/ nanopores, substrate concentrating DNA, non-biological controlled environments to Externally-imposed diffusion, temperature/ thermally chemotaxis, colloidal pattern induce accelerated transport induced. formation with light, protein- or separation between electrolyte interactions. different species.

Coating of latex particles due to dissolution of stainless steel in simple salt dissolution in 16 Solutions saturated at point of solid acid , CaCO3 and BaSO4 Salt or crystal water, diffusioosmotic dissolution, lower concentration in micropumps17, bone crack dissolution transport near wall of system bulk. detection18, repair of cracks in undergoing mixing. polymers19, mineral replacement reactions20.

During sedimentation or screening Coagulant addition in waste processes, particle concentrations Sedimentation, centrifugation, water treatment, non- generate ion gradients, which Sedimentation membranes, screening barometric particle number produce a bulk electric field. More operations22. distribution upon than just a sedimentation centrifugation. potential21.

Geological or biological Saturated at point of crystal mineralization, scale Molecular formation, super-saturated in bulk. formation, kidney stones, sea- crystallization Freezing operations have ice, mineral replacement similarities. reactions, silicon purification.

218

Higher or lower solute Hygroscopic salts, drying of a Evaporation and concentration at the surface of liquid droplet, dehydration of

condensation evaporation or condensation skin or food materials, cloud respectively. condensation on aerosol.

Heated regions of a Temperature gradients give a microfluidic network, even if Solubility or activity variation in chemical activity, which starting at a uniform variation drives diffusion even with an initially concentration, liquid-liquid uniform concentration gradient. extraction system.

C.3 Essential Theory of Diffusiophoresis

There are two general types of diffusiophoresis: electrolyte and non-electrolyte.

Electrolyte diffusiophoresis causes particle transport by generating spontaneous electric fields

(electrophoresis)23 and pressure fields in the Debye layer (chemiphoresis)24,25. The fluid physics

(Figure 2) is well-described in a review article by Anderson3. Non-electrolyte diffusiophoresis

(NEDP) also works by creating a pressure gradient across the particle due to particle-solute interactions, such as repulsive steric exclusion26,27 or attractive van der Waals interactions28,29, and this also is a type of chemiphoresis. Diffusiophoresis has been observed in a variety of natural and artificial systems and applied to cause a broad range of outcomes: aggregation16, pumping17, sensing18,19, separation14,15, active matter and autonomous motion9,10,30 and pattern formation11,13,31,32. In this paper we concentrate mainly on electrolyte diffusiophoresis, which typically produces faster diffusiophoretic transport than NEDP, but most of the mechanisms recognized in this article for generating concentration gradients apply to both.

Part of the mechanism of diffusiophoresis occurs in a manner similar to electrophoresis, except that a concentration gradient produces the electric field (E) instead of direct application of

219 the E field. In an applied E field, the electrophoretic velocity (vep) for a uniformly charged particle with a thin double layer (a >> 1)33 , of any shape2 is given by the Smoluchowski equation2,34 as

ϵ  퐄 퐯 = p (1) ep η

Here ϵ is the dielectric permittivity of the medium, p is the particle zeta potential, and η is the solution viscosity. Eq 1 is a good approximation even for a non-uniformly charged particle with a thin double layer, using the surface average zeta potential.35,36, In chemically driven flows, due to the existence of a diffusion potential, the electric field is generated spontaneously, almost like a

“micro battery” within the solution. This occurs when the ions have different diffusion coefficients.

In order to maintain electroneutrality in the solution, a spontaneous E field arises to slow down the faster-diffusing ion, and speed up the slower ion. This E field, however, acts not only on the ions, but on any surface having a finite zeta potential or charge.

This internal electric field for a Z:Z electrolyte is proportional to the ion diffusivity difference factor (β) and concentration gradient of solute (∇C), given as37

k T 훁C 퐄 = β (2) Z e C where, k is the Boltzmann constant, T is the system temperature, Z is the valence of the constituent ions of the solute, e is the proton charge, and C is the ionic concentration. The diffusivity difference factor () is solely dependent on the nature of the salt and can be expressed in terms of constituent ion valences (zi) and diffusivities (Di). For Z:Z electrolytes,

훽 = (퐷+ − 퐷−)/(퐷+ + 퐷−) (3) where D+ is the cation diffusivity and D- is the anion diffusivity. Note that although each of the individual diffusivities depends upon temperature, this dependence gets cancelled out for β. β values of some common and important electrolytes are listed in Table 2. In chemically-driven

220 electrokinetic flows such as diffusiophoresis, the electrical driving force is generated in-situ and E is replaced in equation (1) with terms comprising concentration gradients, as given by equation (2).

Diffusiophoresis also arises due to a chemiphoretic mechanism. Inside the electrical double layer (EDL) near a charged surface, the pressure is higher than in the bulk, due to the attraction of ions in the EDL. At higher ionic strengths, this pressure becomes higher. If we have a concentration gradient tangential to a surface, we therefore have a gradient of pressure, which causes a tangential flow near the surface in the EDL. This is the essential physics of chemiphoresis.

A similar physics applies for non-electrolyte systems, although the forces are due to steric repulsion

(e.g., hard sphere repulsion) or van der Waals attraction.

Figure C-2: Essential mechanism of electrolyte diffusiophoresis. The mechanism consists of two parallel additive phenomena: electrophoresis caused by an in-situ electric field (E) generated by a concentration gradient (or more precisely, a chemical potential μ gradient) of NaCl, and chemiphoresis caused by a gradient of NaCl concentration, and therefore a gradient of pressure in the EDL, tangential to the particle surface. This pressure gradient drives fluid

221

flow along the particle surface inside the EDL. For a β negative salt like NaCl, both mechanisms transport the negatively-charged particle towards higher ionic concentration.

We note that diffusiophoretic transport has a different mechanism from Marangoni flows.

Marangoni flows are generated by gradients in surface tension, which can arise from gradients in temperature or concentration. On the other hand, diffusiophoretic flows are generated primarily by concentration gradients (i.e., gradient of chemical potential), and no gradient of surface tension is needed. Indeed, diffusiophoresis occurs usually for immersed colloids, and the surface tension mechanism is inoperative for solid particles. In some cases temperature gradients can actually generate concentration gradients, since the temperature induces a gradient in chemical potential, so that diffusion occurs. The Marangoni effect is not dependent on the particle surface charge nor

EDL like electrolyte diffusiophoresis. Therefore, the fundamental difference between Marangoni and diffusiophoresis, is that the former is driven by surface tension gradients derived from a stress discontinuity at the fluid-fluid interface, whereas the latter is driven by a slip velocity at the solid- fluid interface.

Table C-2: Ionic diffusivities at 25˚C and temperature independent β values of common electrolyte species38. Gradients of some salts (e.g., KCl) will result primarily in chemiphoresis. For the reaction CaCO3 + H2O = Ca+2 + OH- + HCO3-, apparent β of CaCO3 is based on equation (3). Cation Cation Anion Anion β values of sample diffusivity diffusivity electrolytes -9 2 -9 2 Salts β (D+, 10 m /s) (D-, 10 m /s) H+ 9.311 OH- 5.273 HCl +0.642

K+ 1.957 Br- 2.080 K-acetate +0.285

+ - NH4 1.957 Cl 2.032 KNO3 +0.014

+ - Na 1.334 NO3 1.902 KCl -0.019

222

+ - Li 1.029 HCO3 1.185 BaSO4 -0.057

2+ 2- Ba 0.847 SO4 1.065 NaCl -0.207

+2 Ca 0.792 1.089 CaCO3 -0.506 Acetate-

+2 - Mg 0.706 H2PO4 0.846 NaOH -0.596

As a result of both the electrophoresis and chemiphoresis components, the overall diffusiophoretic velocity (Udp) of a particle with zeta potential ζp in a symmetric Z:Z electrolyte is39

ϵ k T 2 k T Z e ζ ∇C U = {β ζ − ln [1 − tanh2 ( p)]} (4) dp η Z e p Z e 4 k T C

In a salt gradient of NaCl ( = -0.207), electrophoresis dominates for || < 42 mV, while chemiphoresis dominates for || > 42 mV. This cutoff value changes from 42 mV for NaCl, to 3.7 mV for KCl ( = -0.019), to 184 mV for KHP (potassium hydrogen phthalate,  = +0.65). The key factors leading to diffusiophoretic flows are thus 1) a concentration gradient, 2) a finite zeta potential on a surface or particle, and 3) a difference in ionic diffusivities of the corresponding electrolytes (퐷+ − 퐷− ≠ 0). The equation uses the local concentration field, which for a sphere is taken at the center of the sphere34. The length scale of the salt gradients leading to diffusiophoresis

(L) are almost always larger than particle size (R), i.e. L ≫ R. For multiple ions with arbitrary valence, a generalization of Eq 4 for low zeta potentials is:37

2  z i n i kT z i Din i U  i  2  i  dp 8 z 2 n p e z 2 D n p i i i i i i i (5)

The diffusiophoretic motion of neutral or charged particles, arising out of neutral polymer, polyelectrolyte or other active molecular gradients, can also be fundamentally treated from

223 a colloidal perspective rather than a fluid continuum approach. Brady40 showed that such an approach gives the continuum approach in limiting cases, and also captures the relevant thermodynamic forces at the statistical average level,41 including potential gradients associated with particle and fluid motion, Brownian motion of particles, and interparticle forces. In the hydrodynamic perspective, such forces are balanced by Stokes drag to generate particle motion. However, ionic interaction-driven diffusiophoresis is yet to be validated by such a micromechanical colloidal approach. One of the challenges would be to model the charge interactions within a thin double layer in order to generate sufficient slip velocity.

In addition to the diffusiophoresis theory given above, recent theoretical advances relax a number of assumptions (e.g., soft particles42,43, porous particles44,45 and arbitrary double layer thicknesses46, non-uniform charge47, polyelectrolytes48,49, pH changes50). However, the three essential factors remain the same, of needing a concentration gradient, a finite zeta potential on the particle, and for the electrophoresis part of diffusiophoresis, having a difference in diffusion coefficients of the ions.

At high ionic strength, solution properties such as viscosity and dielectric constant deviate significantly from that of pure water, and must be taken into account. Moreover, the assumptions in the Poisson Boltzmann theory used to derive Eq 4 are pushed to their limits, owing to finite ion size51, dielectric decrement52 and inter-ion correlations53–55. While a theory for diffusiophoresis specifically at high ionic strength has not been developed yet, recent theoretical advances on the treatment of these effects using a modified Poisson Boltzmann equation52,53,56 have progressed.

We note that other transport mechanisms can still be operative in the presence of diffusiophoresis. One of these is Brownian motion, in which the particle (radius a) has a diffusion

224 coefficient D  kT / 6a given by the Stokes-Einstein equation. In order to compare diffusiophoresis with Brownian motion, we can use a Peclet number, Pe = Udpa / D. Keeping just the electrophoresis part of Eq 4 assuming particles with a a low magnitude of zeta potential, we find that

6a 2 c  c Pe ~ 2 1 (6) ZeL c

where c2 is the concentration of the diffusing electrolyte species at the particle surface, c1 is the concentration of diffusing electrolyte species in the bulk, L is the length scale of the concentration gradient, and c is the bulk concentration of electrolyte species, including non-diffusing species.

Note that even DI water at room temperature typically has dissolved CO2 that gives an ionic strength of roughly 2.5 µM. For a dissolving sphere, the radius (a) would be the length scale. Thus, if a species has only a slight concentration difference ~ (c2 – c1) in a large background concentration,

Pe is small. However, even in sea water or saturated conditions, diffusiophoresis can be important if (c2 – c1) is large compared with c.

Concentration gradients can also lead to density-driven flows. However, there are several tests one can use to distinguish between density-driven flows and diffusiophoretic flows: 1)

Particles with different zeta potentials should move at different velocities if moving due to diffusiophoresis. Similarly, when diffusioosmosis is present, changing the zeta potential of the stationary surface should change the velocity of particles near the surface. In both these cases, the buoyancy driven flow does not change. 2) When feasible, one can try to reverse the direction of diffusiophoretic transport by switching from beta negative to beta positive salts (Table 2), even while maintaining the same density gradient. 3) Density gradients can be reduced by the addition of a non-ionic solute, such as sucrose, to the low concentration solution, in order to increase the

225 fluid density. Changing the salt or adding solutes might also lead to a change in viscosity and zeta potential, which must be accounted for.

C.4 Origins of Concentration Gradients

Here we discuss a range of molecular mechanisms which can give rise to concentration gradients and hence diffusiophoresis. These mechanisms follow from physicochemical and biological processes that have been analyzed previously, but not always in the context of producing transport. The first six of the mechanisms we discuss, namely - molecular exclusion, chemical reaction, diffusive mixing, externally imposed, salt or crystal dissolution, and sedimentation – each have existing examples of diffusiophoresis in the literature. We add potential situations where these mechanisms might also be acting. The final three mechanisms, namely - molecular crystallization, evaporation and condensation and solubility or activity variation are not usually considered as causes for diffusiophoresis.

C.4.1 Molecular Exclusion

Membrane separation involves a myriad of electrokinetic effects both at microscale and nanoscale. Capillary osmosis was one the first electrokinetic effects which was demonstrated across a microporous membrane57. Derjaguin et al. also theorized existence of reverse capillary osmosis originating from capillary osmotic fluid flow when fluid of higher concentration is pressurized across membrane pores58. Denisov et al. performed theoretical analysis to describe protein transport through a porous membrane59. They predicted that in presence of pH gradients across membrane

226 pores, diffusiophoresis significantly enhances protein transport. Such electrokinetic effects were termed as Non-Equilibrium Electric Surface Phenomena (NESP) by S. S. Dukhin who highlighted their applicability to several membrane processes60. Membrane processes are increasingly being proposed for use in sustainable energy generation by harvesting salinity gradients61,62.

Electrokinetic flows, such as diffusioosmotic fluid flow through single boron nitride nanotube, were demonstrated to convert such salinity gradient energy to electric current with high efficiency63.

Molecular exclusion is an important mechanism in reverse-osmosis membrane systems. In these systems, a solution is pushed through a membrane by a pressure-driven flow, and the electrolytes are excluded from passing through the membrane. The result is a buildup of ions at the membrane, which is called a “concentration polarization layer”. This concentration of ions is higher than that in the bulk, and as described previously, if the ions have different diffusivities – as occurs with NaCl – this concentration gradient will lead to an E field. The solution will usually have some finite concentration of particles or bacteria that have a finite zeta potential, and now we have all three pieces required for diffusiophoresis: a concentration gradient, different ion diffusivities, and finite particle zeta potentials. The outcome of this convergence is that not only do particles build up at the surface due to membrane exclusion, but also, the concentration gradient of ions actually pulls the particles there faster.

One of the basic questions in membrane diffusiophoresis is to define a characteristic length of the phenomenon. The length scale of diffusiophoresis can be conveniently deduced from the definition of diffusiophoretic co-efficient. For example, a particle with zeta potential of -25 mV at

298K in NaCl has 훾 ≈ 10−10 m2/s. Here 훾 represents an equivalent diffusivity term composed of

ϵ k T β 휁 , inverse of which signifies resistance to electrophoretic part of diffusiophoretic flow ηZ e 푝

(supplementary information). If 10 mM NaCl feed is purified using a salt rejecting membrane in a

227 system volume (V) of 1 liter, then ratio of overall change in volumetric Gibbs free energy of mixing to salt osmotic pressure (supplementary information) over a characteristic length scale is as follows-

x1 logx1+ x2 logx2 ∆G̅ (N1+N2) R T mix = V ≈ 10 (7) ∆π N1 RT V

Where, 푁1 and 푥1 are solute moles and mole fraction, respectively and 푁2 and 푥2 are solvent moles and mole fraction, respectively. For a measured particle diffusiophoretic velocity of 1 m/s, the characteristic length scale,푙~ γ/ Udp ∆G/∆π, stays in the order of millimeters (mm) which is reasonable since diffusiophoresis is a microscale phenomenon active within concentration boundary layers on membranes. Additionally, the Peclet number from Eq 6 can be used to deduce importance of convective contribution in membrane systems filtering dilute concentrations of NaCl with particle length scale (L) ~ 100 nm and -30 mV zeta potential, which was estimated to be

~O(10). This directly implies a significant electrokinetic contribution in salt rejecting membrane systems.

The established theory for particle deposition on dense membranes, such as nanofiltration and reverse osmosis processes, is cake enhanced concentration polarization64. This is simply driven by enhanced salt osmotic pressure due to hindered back diffusion through deposited particulate layers on such salt rejecting membranes. Water permeation through dense membranes is pressure driven solution diffusion. Therefore, particle deposition on such membranes was theoretically accepted as purely convective, devoid of any electrokinetic transport mechanisms. However, diffusiophoresis in salt rejecting membrane surfaces, particularly within the concentration boundary layer has been demonstrated, both experimentally and theoretically, by Guha et al. by probing the permeate flux and cake layer growth rate with different feed salts under similar conditions8.

228

Figure C-3: Diffusiophoretic velocity increases and dominates over convective velocity with reduction in porosity. 20 mM NaCl is rejected by an ideal semipermeable membrane with water flux of 23 LMH. Porosity change is time dependent, and also depends on the type of salt (훃) and particle zeta potential 퐩. With higher values of 훃 and 퐩 the diffusiophoretic velocity increases further.

The relative contribution of diffusiophoresis to convection (Udp/vw) determines the magnitude of particle deposition and fouling. For just the diffusiophoresis part, the system at hand dictates the contributions of the electrophoretic and chemiphoretic parts of the diffusiophoresis. For example, in a typical low salinity water processing scenario such as that in wastewater recycling with a high salt rejecting reverse osmosis membrane, around 56% diffusiophoretic particle flux is contributed by electrophoresis and the rest (~ 44%) is contributed by chemiphoresis, both inside and outside of deposited cake layer. This is important in the context of reversing the diffusiophoresis which only reverses the former component, the later (chemiphoresis) is always oriented with convection in salt rejecting membrane systems. Therefore, in common NaCl rejection process, diffusiophoresis aligns with convection to enhance particle flux significantly and causes accelerated growth of the cake layer. In 20 mM NaCl feed and at 23 liters per meter squared per hour (LMH) permeate flux, diffusiophoresis contributes to more than half of the particle flux transported through pure convection (Figure 3). Increased particle transport rate causes decreased

229 cake layer porosity which could be correlated to β values of the salt8. Particle fouling on dense membranes, therefore, has been proposed to run in a vicious cycle through enhancement of particle transport rates towards membrane due to decreased porosity over time mediated by diffusiophoresis.

A particle boundary layer can be defined where cake porosity reaches 0.36 and no further movement of particles towards the membrane is possible owing to limiting dense random packing achieved. Within this particle boundary layer, diffusiophoretic velocity competes with convective velocity and no particle could escape. This effect would be even higher at theoretical minimum porosity of 0.26. However, with non-aggregating different size spheres, porosity can become smaller than 0.2665. In such cases, diffusiophoresis would take over as the dominant mechanism of cake compaction and membrane fouling would be severe. Over and above this, membrane fouling further propagates, if 푝 of particles or β of salt increases (Figure 3).

Diffusiophoresis across a porous membrane is interesting from the standpoint of both facilitated colloidal transport as well as control of colloidal deposition. Derjaguin et. al. first showed diffusiophoresis can deposit particles transiently on cellophane membranes separating concentrated

60 CaCl2 solution and dilute latex particle suspension . Ebel et al. demonstrated transport of latex particles across a track-etched mica membrane due to diffusiophoresis generated by concentration

5 gradients of different salts . Kar et al. reported salt (LiCl, NaCl, KCl and CaCO3) dependent transient deposition of latex particles on microporous hollow fiber membrane driven by diffusiophoresis7. Florea et al. demonstrated transient exclusion zone formation in the vicinity of ion-exchanging Nafion membrane surface due to diffusiophoresis66. Manipulation of the electric field in the membrane vicinity, could therefore, be key in controlling particle deposition and fouling mitigation.

Since a decrease in hindered diffusion co-efficient (Dh) orchestrated by decrease in cake porosity, is the leading factor in colloidal fouling of dense membranes, porosity increase is one of

230 the ways to mitigate fouling. Concentration gradients of β +ve salt (eg. KIO3) would be advantageous in this context. The reverse electric field generated by such salts would cause diffusiophoretic velocity opposite to pressure induced convection leading to loosening of the cake and decrease in diffusiophoretic fouling.

Chromatography, used to separate charged species such as proteins is another system where ionic or polymer gradients may be generated. For example, both pore diffusion and homogeneous diffusion are important mechanisms for protein uptake by porous ion exchange beads67. Protein exclusion mechanism, equivalent to non-electrolyte diffusiophoresis, was also hypothesized to explain enhanced dynamic binding of proteins to ion exchange resin surfaces68. Such volume exclusion mechanism is important in propelling particles (charged and uncharged) away from higher concentrations of solute.

Diffusiophoresis of proteins in presence of amino acid, salts and pH gradients may also have significant relevance in chromatographic separation of proteins. Studies have shown that PEG diffusiophoresis increases with increasing molecular weight in steady state salt concentration

69 70 gradients as well as in salting out salts such as Na2SO4 . Thus concentration gradients and associated electric fields could significantly contribute to protein transport. Similarly, elution of captured protein could be modified depending on the salts or amino acids (for eg. arginine hydrochloride vs NaCl) used71. McAfee and Annunziata demonstrated that with salting in solutes, anion binding to PEG effect dominates the interactions72. Therefore, proper solute incorporation may modify protein-polymer interaction and thereby, improve elution performance.

231 C.4.2 Chemical Reaction

Reactions on surfaces can generate either ionic or non-ionic gradients. When the reacting surface is fixed, diffusioosmotic fluid flows result at a charged surface, which can in turn drive macroscopic fluid movement. At the same time, charged particles in the surrounding fluid can undergo diffusiophoretic transport, in addition to being entrained by macroscopic fluid movement.

For example, TiO2 particles immobilized on a glass surface pump silica tracers away from the

13 particle when placed in an environment of H2O2 in DI water, and irradiated with UV light .

For particles suspended in a fluid, a chemical reaction involving electrolyte species will cause the particle to produce a self-generated electric (E) field, resulting in autonomous motion due to self-diffusiophoresis. A previous review from our group on mechanisms of biomimetic colloidal behavior covers examples and the historical development of this phenomenon32. Such motion requires a broken symmetry, which can be for instance in the mobility or surface reactivity.73 The asymmetry might be naturally present10,11 or can be obtained by introducing dimers74,75, Janus particles76, or shape asymmetry77. An example of such diffusiophoresis powered motion are the light activated Ag-dynabead heterodoublets which swim in a self-generated gradient of Ag+ and

- 10 OOH ions obtained from the decomposition of H2O2 . Another similar example is the case of AgCl particles in water which swim autonomously upon shining a UV light due to the release of H+ and

Cl- ions, at remarkable speeds of up to 100 body lengths per second at low ionic strength11. In both these cases, the reaction begins only upon activation by UV light and thus the autonomous motion can be switched on or off externally. Autonomous swimming of platinum-polystyrene Janus particles in H2O2 has also been considered for a diffusiophoretic mechanism, through the asymmetric generation of ionic intermediates on the platinum surface78,79. Table 3 summarizes various self-diffusiophoretic particles reported in the literature, along with brief descriptions of where the reaction has produced transport.

232 In analyzing the relative impact of a chemical reaction in causing transport, we use the

Peclet number from Eq 6 to compare it to Brownian motion. Reaction rates can vary widely, but a typical reaction rate on metals has been reported as 0.071018 / m2-s to 0.81018 / m2-s80. Using Eq

6, and entering typical values such as a = 1 m,  = 808.85410-10 C2/N-m2,  = 0.5 (approximated from Table 2, taken from the range),  = kT/e = 0.0257 V, Z = 1, e = 1.610-19 C, and L = 1 m, one finds that Pe = 1040. That is, diffusiophoretic convection dominates Brownian motion in the presence of this chemical reaction.

A self-generated electric field can cause rotational motion in addition to translation when the mobility, or zeta potential is non-uniform on the particle surface, resulting in a net dipole moment which tries to align itself with the E field. The dependence of orientation and swimming velocity on substrate concentrations gives these reactive particles the ability to sense gradients – a phenomenon called chemotaxis similar to that demonstrated by live bacteria81.

Table C-3: Chemical reactions given in the literature that produce diffusiophoretic transport. Particle Reaction Fuel Phenomena

10 + - Ag Ag + HOOH  Ag + HOO + H2O H2O2 + UV in Schooling, partly DI water irreversible AgCl11 AgCl  Ag+ + Cl- UV in DI water Schooling, predator prey, reversible

11 +2 - MgO MgO + H2O  Mg + 2 OH DI water exclusion, partly irreversible

13 TiO2 h   O UV in DI water Surface pumps, TiO  TiO (h  e ) 2O 2 2 2 motors, reversible H O  2OH  H

12 Ag3PO4 2UV and 2mM Exclusion and Ag 3PO4 + 6 NH3 + H2O  3Ag(NH3 ) 2- -NH 3 in water schooling – + HPO4 + OH oscillatory, reversible

82 + + - Gold N2H5  N2 + 5H + 4e N2H2 + H2O2 Reversible, schooling

+ - H2O2 + 2H + 2e  2H2O

233

PHSA Light induced redox reaction or refractive index- 3 D colloidal crystals, coated ionization of dissolved species and density- reversible PMMA 83 matched low- dielectric solvent Titania or Light induced decomposition of H2O2 + light Live colloidal crystals hematite H2O2 (blue or UVA- due to collision, self- composite violet) propulsion and with TPM pumping of tracers polymer77 towards, reversible

When particle density is high enough, particles – even non-reactive particles – can also start to respond to concentration gradients set up by other particles. A type of communication can therefore result, giving schooling, exclusion regions and phase separation behavior, which have been demonstrated experimentally in the case of light triggered Ag-dynabead doublets10, AgCl

11 77 particles , the light triggered colloidal crystal formation of TiO2-polymer autonomous particles , autonomous gold particles82 and the silver phosphate oscillatory system12. Communication between particles through diffusiophoresis has been the subject of a number of recent theoretical studies, such as those by Golestanian et al84. Swimming of Janus particles close to a wall has also been considered theoretically29,85,86 and experimentally87 and diffusiophoresis has been suggested to be the reason for restricted rotational diffusion of these particles close to a wall. Uspal et al discuss modes of interaction due to hydrodynamics and self-generated concentration gradients, breaking of symmetry by the wall, sliding and hovering88. Brownian dynamics simulations can be applied to reveal the dynamics of these systems89,90.

C.4.3 Diffusive mixing

When two phases out of electrochemical equilibrium are placed in contact with each other, diffusion tries to establish equilibrium. In the absence of convective flows, diffusive mixing is a

234 slow process with a timescale of L2/D for diffusion to dissipate the gradient. When the gradient in question is that of a salt, diffusiophoresis can significantly speed up or slow down the mixing process. Diffusive mixing can be used to create gradients in the laboratory. In a study on autonomous particles, we created gradients of hydrogen peroxide by putting a concentrated solution inside a capillary closed at one end, and then introducing it into a petridish which was filled with

DI water91 (Figure 4a).

In another study, we placed one capillary partly inside another, each of them containing different salt concentrations14 (Figure 4b). One of the capillaries contained oil droplets and was closed at the outside end to simulate a dead-end pore. Thus we demonstrated that oil droplets can be extracted from a dead-end pore using diffusiophoresis. In fact, our experiment is a laboratory scale simulation of a geological reservoir which is flooded with freshwater in the LoSal process of enhanced oil recovery92. While the situation in question is that of high ionic strength as opposed to the low ionic strength used in our study, oil droplets trapped in the dead-end pores will nevertheless be affected by salt gradients, and diffusiophoresis may be able to explain the increased yield.

Figure C-4: Generating gradients by diffusive mixing in the laboratory. (a) 2 cm long and 0.15 m thick glass capillaries were filled with different concentrations of a solute and placed in a petri dish that contained no solute91. At the mouth of the capillary, two solutions of different concentrations meet, creating a concentration gradient. (b) A capillary was filled with salt solution and one end was sealed off. The other end was dipped into a bigger capillary that served as a reservoir containing only DI water14. Again, two salt concentrations were placed adjacent, giving a gradient.

235 Natural gradients due to mixing are also found in the highly-stratified regions of water bodies such as estuaries where layers of different salt concentrations exist on top of each other. Salt concentrations in the two layers may be different due to phenomena such as high surface evaporation, mixing of seawater and freshwater, currents or melting of sea ice. High values of gradients are encountered in thin intermediate layers (~m or tens of m) called haloclines over which salt concentrations change drastically. Stable haloclines, where the density of the top layer is lower than that of the bottom layer, are encountered in estuaries and polar seas. Steep haloclines resulting in a maximum C/C of 0.18-0.25 are found in estuarine haloclines such as Rio De La Plata93 estuary in South America, Kattegat94,95 in the Baltic Sea region and San Pablo bay96 in the San

Francisco estuary. These would result in a very small diffusiophoretic velocity of ~ 0.1 nm/s, which might at first thought seem neglible. However, the residence time of some of these water-bodies is several years (for example 22 years for the Baltic sea) so that Pe ~ 10 may be observed over a 10 year period for a 1 m particle. Moreover, much steeper local, transient salt gradients may be obtained96 during which diffusiophoresis would result in additional particle transport over and above the fluid mixing.

Freshwater and seawater mixing must result in gradients of other salts as well, apart from

NaCl. If such gradients of higher beta value salts are present when the overall salt concentration is low (as will be on the fresh water side), much higher Pe numbers may result. However, this hypotheses is not easy to verify since most existing measurements of salinity in the literature are in practical salinity units97,98, a conductivity based measurement, which accounts for changes in concentration of all salts.

Apart from direct competition with convection under low flow rate or static conditions, diffusiophoresis may impact turbulent flow through a diffusion-like contribution across streamlines, resulting in orders of magnitude lower Schmidt numbers for the flow15. Thus, in spite of the low

236 Pe numbers obtained in our analysis, diffusiophoresis may still impact the transport characteristics of pollutants and other colloids in estuaries and coastal aquifers99. Diffusiophoretic transport has never been accounted for in these situations and mixing in the presence of vertical or horizontal stratification is assumed to be purely diffusive if not forced by convection due to tides, currents or wind. The theoretical studies of Anderson et al100, quantifying the effect of fluid convection perpendicular to concentration gradient as well as that of Chiang & Velegol in considering the effect of multiple ions101 will be important in understanding these complex situations.

Concentration boundary layers form when two fluids of different compositions are transported over a surface, one after the other provide an interesting example of concentration gradients generated due to the interplay of diffusion and the boundary no-slip conditions100,102. Boundary layer diffusiophoresis (BLDP) can cause particles to deflect perpendicular to the flow, allowing crosswise mixing as observed in the above mentioned study14 of an idealized dead-end pore (Figure

5).

237

Figure C-5: Boundary Layer Diffusiophoresis (BLDP). (a) A transverse salt gradient is created due to the concentration boundary layer as the DI water from outer reservoir flows diffusioosmotically on the inner capillary that contains 10 mM NaCl or KCl. This causes tracer particles to deflect towards the walls due to BLDP as they enter the capillary102. (b) & (c) Tracks of tracer particles as they enter from the outer capillary containing DI water into the inner capillary containing 10 mM NaCl in (b) and 10 mM KCl in (c). The deflection towards the wall is attributed to BLDP. The deflection is small in this case since KCl has a much lower  value14. It must be mentioned that two layers in contact may be at electrochemical equilibrium in spite of being at different salinities, if the two layers have different solubilities of the salt. In such a situation, no diffusiophoretic transport will occur. The E field arises only when the gradient leads to net diffusion.

C.4.4 Externally-imposed

Microfluidic set-ups are being used extensively to induce concentration gradient driven diffusiophoresis at colloidal length scales. Membrane imposed gradients were initially used to validate the diffusiophoretic hypothesis and develop theoretical frameworks5,23. However, since the advent of microfluidics, gradient generation has become easier, and different microchannels have been used to impart the diffusiophoretic driving force. The sandwiched channel configuration by

238 Diao et al. enabled them to study bacterial chemotaxis under concentration gradients of specific molecules circumventing any other hydrodynamic effects81. Abecassis et al. devised a three arm microchannel to generate transverse concentration gradients to focus and spread colloidal particles along the length of microchannel15. They found that the extent of colloidal focusing or spreading depends on the type of salts used, validating the diffusiophoretic hypothesis. In a similar three channel setup81 carved out in an agarose gel, Abecassis et al. demonstrated diffusiophoretic movement of fluorescently labeled λ-DNA in different solute contrasts similar to fluorescent colloidal particles103. Liu et al. hypothesized that polyelectrolytes, such as DNA, can have larger diffusiophoretic mobility than corresponding rigid particles. This is due to the fact that fluid flow might take place inside the polyelectrolyte structure itself (i.e., free-draining), apart from surrounding flow due to in-situ electric field. For example, in case of negatively charged DNA in

NaCl gradients, the co-ions inside DNA would move away from approaching Cl- ions generating a drag force in the same direction as NaCl diffusion induced outer electric force. Therefore, polyelectrolyte diffusiophoresis is different due to dependency on drag coefficients and internal counter ion condensation at higher charge density, unlike rigid particles48. Wanunu et al. showed that with imposed salt gradients across nanopores of SiN membrane, double stranded DNA translocation rate increased by more than 30 fold104 and theoretical analyses have been developed for this situation105. Therefore, diffusiophoretic mechanism can be further developed to concentrate

DNA and possibly separate different DNA fragments based on charge.

Palacci et al. demonstrated osmotic trapping and pattern formations of colloidal particles by utilizing logarithmic sensing of diffusiophoresis at low solute concentrations106. They have used a sandwiched configuration having two outer hydrogel channels to flow solutes with a syringe pump. They observed spatio-temporal colloidal patterns in the center channel. Maeda et al. demonstrated pattern formations with temperature-gradient-induced diffusiophoresis of T4 DNA

239 and small RNAs in entropic gradients of PEG107. Such optothermal diffusiophoresis was effective in generating geometric patterns of colloidal particles, DNA and bacteria with simple spatial manipulation of laser beams108. With higher volume fraction PEG, diffusiophoresis was able to overcome the Soret effect (due to temperature gradient transport) and therefore, depletion patterns changed to accumulation with different colloids.

It is worth noting that the diffusiophoretic hypothesis of DNA transport may have far reaching implications in answering some fundamental questions such as how viral DNA can locate and travel towards the nucleus after entering into the cell (Figure 6)? Such cellular diffusiophoresis due to imposed metabolic effects can be manifested in both intracellular as well as extracellular transport. Intracellular pH gradients may lead viral DNA to migrate towards the nuclear pores after accumulating in the microtubule organizing center (MTOC)109. It is known that pH microenvironment of human cell nuclei prevails at ~ 0.3 – 0.6 unit higher than cytosolic pH110.

Therefore, a viral DNA of zeta potential ~ -30 mV can sense a diffusiophoretic driving force of

+ + + ∇c/c ~ {(퐻푛푢푐푙푒푢푠 − 퐻푀푇푂퐶)⁄푙}/(퐶푀푇푂퐶 + 퐻푀푇푂퐶) , over a characteristic length of l ~ 500 nm

+ + between MTOC and nuclear pore. 퐻푛푢푐푙푒푢푠 (pH ~ 7.8) and 퐻푀푇푂퐶 (pH ~ 7.2) are proton concentrations in nuclear region and MTOC, respectively and 퐶푀푇푂퐶 is the cytosolic ionic concentration (~ 150 mM). Therefore, the characteristic transport time of viral DNA from MTOC to nucleus is 푡 ~푙/푈푑푝 ~ O (1 h). This is of the same order of transport time when majority of the

HSV-1 virus capsids are transported to fibroblast nuclei111.

240

Figure C-6: Schematic of hypothesized viral transport from MTOC compartment to nuclear pore in eukaryotic cell. One of the transport mechanisms could be diffusiophoresis driven by pH gradients (훁퐜/퐜) near the vicinity of nucleus from cytosol. A novel microfluidic setup with a hydrogel window has been used to generate a concentration gradient generation with nonionic solutes such as ethanol112 producing transport according to the laws of diffusiophoresis113. In all the examples above, colloid movement was achieved by direct or indirect salt concentration gradient generation and primarily through diffusiophoresis, at colloidal length scales.

Table C-4: Summary of the effects of externally induced diffusiophoresis. Mechanisms Relevant Effects/ References Applications

Ionic diffusiophoresis a. Colloid focusing/ spreading 15 in a microchannel

b. dsDNA translocation through 104 nanopores

241

Optothermal Pattern formation with diffusiophoresis particles, DNA/ RNA and bacterial cells with changing 107,108 macromolecular concentrations by laser heating

Polymer diffusiophoresis and Enhanced diffusivity of PEG in ion-polymer interactions salt gradients following 69,70,72 Hofmeister type interactions.

Solvophoresis Particle movement away from diffusing polar solvent 113 gradients.

Cellular diffusiophoresis Exclusion zone (halos) around red blood cells, migration of 114,115 Dictyostelium cells in PEG conc. gradients

C.4.5 Salt or crystal dissolution

A fractured bone, a scratched glass slide submerged in water, or a rupture in the earth caused due to an earthquake or fracking, are all examples of cases where a new surface is exposed to a fluid, resulting in the transport of ions that equilibrate with the surrounding liquid. Such situations create a transient ionic gradient in the fluid due to dissolution or leaching, where the ion concentration goes from saturated at the surface to a lower concentration in the bulk.

Calcium carbonate microparticles acting as pumps as they dissolve in water is a demonstrative example of diffusio-osmotic and -phoretic transport due to dissolution17. As the

- - 2+ CaCO3 particles dissolve, HCO3 , OH and Ca ions are released in the solution and an electric

242 field is generated due to the different speeds of these ions, resulting in the pumping of charged tracer particles towards or away from the pump, depending upon their polarity (Figure 7). When particles are settled close to the bottom wall, the net motion of tracers is a sum of the diffusiophoretic transport and diffusioosmotic flow. A direct application of calcium carbonate pumps is for detection and healing of bone cracks18, based on the principle that the leaching of calcium ions from a freshly ruptured bone diffusioosmotically drives particles towards it. While the direct application of this method in vitro is challenging due to the high physiological salt concentration (~100 mM), its application for general crack detection in polymeric media is more viable19.

Figure C-7: Left17: Diffusioosmotic pumping of 1.4 m sulfated polystyrene latex tracer particles by two interacting 7 m CaCO3 particles settled close to a glass slide. Right17: A clear exclusion region of tracers develops around a barium sulfate micropump on a glass surface. Despite the very low solubility of BaSO4, the diffusiophoretic pumping still occurs. These tracers, after being ejected, exhibit mostly Brownian motion. The scale bars are 10 m. The generation of gradients due to dissolution is actually quite general – all it requires is a surface that can leach ions, in contact with a liquid phase not in equilibrium with these ions. There are three important parameters to be concerned about when designing a system to make use of diffusioosmotic and phoretic phenomena – the magnitude of the electric field, its profile with time owing to its transient nature and lastly, its potential to cause flows via the  potential of affected

243 surfaces. The magnitude of the E-field, as pointed out before, depends on the  value that quantifies differences in ion diffusion coefficients. The evolution of the E-field with time depends on the rate of dissolution and the salt solubility. Lastly, the  potential depends on the native surface properties as well as the type of salt116 and salt concentration117, which in general will vary as a function of position and time in the system.

Sometimes dissolution may be coupled with reprecipitation of another species present as an electrolyte in the solution, leading to a pseudomorphic mineral replacement reaction (pMRR).

For example, a KBr crystal placed in a saturated KCl solution will undergo a pMRR resulting in a porous KCl crystal in a KBr/KCl aqueous solution118. While the presence and role of fluid flows in carrying out mineral replacement was acknowledged previously119, the cause of this fluid flow in the pores has not been elucidated. Recent work in our group has shown that diffusiophoresis and diffusioosmosis are instrumental in the replacement of KBr by KCl, as these mechanisms generate fluid flows in pores generated during replacement. The developed pores are dead end, so pressure driven mechanisms are ruled out, but diffusioosmotic and phoretic flows are known to occur in dead end pores14. Salt gradients generated during the replacement are able to predict the fluid flow within the correct order of magnitude unlike solid state or liquid diffusion alone. These fluid flows have been shown to be able to eject particles from inside the crystal suggesting the possibility of extracting oil without fracturing rocks102.

A significant surprise has been the role of diffusiophoresis and diffusioosmosis in mineral replacement reactions, since systems are at very high ionic strength. Can electrokinetic effects exist at such high ionic strength, say saturated systems at ~ 5M? The classical DLVO theory predicts

Debye length of ~0.1 nm at such high ionic strength117. Eq 4 suggests that we need both a non-zero

 potential as well as a finite Debye length (-1) for the generated E field to cause any motion. To that end, the  potential of silica surfaces up to saturation has been measured and found to asymptote

244 to roughly -20 mV120. Non-zero  has been recorded on mineral particles at high ionic strength in the molar range121. The structure of Debye lengths at such high concentrations remains an open problem, subject to speculation and further research122.

C.4.6 Molecular Crystallization

Crystallization is a useful phenomenon for separating extremely pure material from solution and is broadly used in the synthesis of proteins, nano- or micro-particles and of course, crystals of minerals. Even though thermodynamically favorable, crystallization does not begin the moment a solution is saturated, and supersaturation is well known to exist in varying degrees in nature123, especially in geological processes where thermal gradients are commonplace and can easily result in supersaturated solutions124. Recent research on crystallization suggests that it proceeds by a particle attachment mechanism125,126, as opposed to a monomer addition mechanism as previously believed from classical theory127. Thus, the onset of crystallization and its growth is greatly affected by the dynamics of the process which is held accountable for a number of unexplained phenomena such as branched and irregular morphologies of synthetic nanoparticles as well as bio-minerals found in organisms125. We believe that accounting for diffusiophoresis can play an important role in understanding the driving force behind some of these unexplained kinetic effects. The presence of supersaturation raises the possibility of a spatially non-uniform salt concentration in which charged surfaces or nuclei could translate convectively towards or away from a growing crystal.

Let us estimate the possible speed resulting from diffusiophoresis for the crystallization of a sparingly soluble salt, barite (BaSO4) which is a nuisance as it forms scales in industrial pipes.

245

0 The saturation concentration (Cs) of barite in water at 25 C is 0.01 mM. Since nucleation is a kinetically controlled event, sparingly soluble salts require a higher degree of supersaturation in order for nucleation to occur123, so let us assume that crystallization begins when the bulk concentration (Cb) is 0.02 mM. In practice much higher concentrations can be achieved before the

123 onset of crystallization (>1 mM) . We assume that when a nucleus of radius (r0) 10 nm is formed, salt concentration at its surface is Cs and it exponentially increases to Cb, as we move away from the crystal. The formation of the crystal depletes salt from the solution in its vicinity and a mass balance gives the length scale (R) of this exponential decay to be 68 nm. Now assuming a simple

5 4 gradient of 0.02 mM to 0.01 mM over 680 nm gives C = 1.5x10 mol/m . For a 0.02 mM bivalent

-1 salt, the Debye length ( ) is 40 nm and BaSO4 = -0.06. For such a salt gradient, a particle with  =

25 mV will move with a velocity of 16 m/s due to diffusiophoresis towards the higher salt concentration which is away from the nucleus formation site. If instead the nucleus is negatively charged, it will move towards the decreasing salt concentration and hence towards the site of crystallization. For an unmixed, purely diffusive system, this can be important for transport over say 10 m for which Pe = uL/D = 13. Thus, diffusiophoresis provides a clear pathway for kinetic effects in crystallization (Figure 8). The direction of velocity depends on the beta value of the salt and the sign of charge on particle, and the magnitude of velocity depends on the ionic strength and degree of supersaturation. Diffusiophoresis can also explain some effects imparted by nucleation inhibitors which are often organic molecules that could carry charge and hence affect the magnitude and sign of particle zeta potential by getting absorbed to the surface. Some of the effects of nucleation inhibitors are increased supersaturation and changed crystal morphology128,129. If the kinetics of crystallization are indeed affected by diffusiophoresis, we might be able to better design methods to impregnate certain additives in a growing crystal.

246 While the contribution of diffusiophoresis in crystallization rate in diffusion limited conditions or due to the presence of crystallization inhibitors is speculative at this point, some hypotheses can be tested by looking at the effect of positive vs. negative crystallization inhibitors.

Another possibility is looking at the crystallization kinetics of beta positive vs beta negative salts.

Diffusiophoresis should only be important for low solubility salts since high supersaturations and hence salt gradients can result in only those cases.

The presence of pores can significantly increase the salt gradient present in the system as it slows down transport. For example, significant salt gradients have been documented for BaSO4 crystallization through porous silica gel123. The influence of porosity on transport and hence diffusiophoresis may also be pertinent to the salt-crystallization observed in building materials where the walls provide a porous framework for salt transport130.

Figure C-8: A schematic of diffusiophoresis in crystallization: A small seed crystal of size r0 is spontaneously formed, with density,, radius r0 and >0 (say). Its formation depletes salt of -value, over a length scale R so that the salt concentration goes from saturation, Cs to that of the supersaturated bulk, Cb. The d\Debye length in bulk is -1. The salt gradient gives rise to a spontaneous electric field and the particle moves towards higher salt concentration as indicated by the direction of udp. An important situation pertinent to salt gradients and crystallization is the formation of kidney stones, as well as the method to treat it131. Stones are usually caused by the precipitation of

-9 2 calcium oxalate in the kidneys. It is treated by the addition of citrate (D- = 0.6 x 10 m /s) which

247 binds calcium and reduces calcium supersaturation in the urine. It has been found that potassium citrate is much more effective at reducing the crystallization than sodium citrate132. While both of these salts are  positive and will both tend to decrease the crystallization rate of positively charged particles, potassium citrate ( = 0.53) will increase crystallization more than sodium citrate ( =

0.38). Since the zeta potential of particles in the kidney is a function of the particle type, additives

(such as inhibitor), pH (regulated by citrate and bicarbonate) and specific adsorption (of citrates or oxalates) in the renal stream, diffusiophoresis may have a contribution in explaining the observed differences, meaning that the dissolution has some transport limitations with the sodium citrate.

Another interesting situation where salt gradients are formed due to precipitation is the formation of sea ice in polar regions. As sea water freezes at -1.81 ⁰C, crystallized water rejects salts which results in a porous ice which contains pockets of brine133. The presence of salinity gradients134, especially while melting must result in diffusioosmotic and phoretic flows that may be the primary cause of transport in the porous sea-ice network, mediating the kinetics of ice- melting as well as transfer of nutrients and wastes between sea water and sea-ice. Such diffusiophoresis aided transport processes may be important to the survival of biosphere-regulating algae and bacteria133,134 in the harsh sea ice conditions.

C.4.7 Sedimentation and screening

Sedimentation creates an electric field in two ways. Local electric field exists due to deformation and polarization of electrical double layer of charged colloids moving under sedimentation or centrifugation21. This is called a “”. Since the colloids have a finite zeta potential, the local electric field slows the particle sedimentation

248 However, an anomaly in such sedimentation-diffusion processes was first observed by

Piazza et al. who attributed an observed non-barometric particle number density distribution to an electric field existing in the bulk solution at equilibrium135. This equilibrium electric field was first experimentally observed with very low salt concentrations, using charged silica nanoparticles in ethanol by Rasa and Philipse22. Under conditions of ultracentrifugation, the particles settle, carrying ions in their electrical double layers (EDLs). These ions want to diffuse back into the solution, but are held back by the particle EDLs. The result is an “entropic lift” (i.e., diffusive lift) of the charged colloids, and therefore, the particle number density deviates from ideal barometric profiles. This is analogous to generation of a pseudo double layer consisting of entropically-dispersed ions and settling particles. Such an electric field is also analogous to a Donnan potential across a membrane; however, instead of an ionic charge imbalance, the potential arises from a discrepancy in the colloidal particle distribution136.

Centrifugation can thus create salt concentration gradients, and diffusiophoresis can play an important role in the resulting equilibrium distribution of charged colloids. The effect of ionic gradients in concentrating Rubella virus antigens in CsCl density gradient centrifugation was reported to yield higher buoyant densities than sucrose or Ficoll imposed density gradient centrifugations137. The ionic diffusivities of Cs+ and Cl- are almost similar and therefore, the diffusiophoresis of antigen particles was expected to take place primarily due to chemiphoresis in this case. Anytime particles are screened, and form a layer, a concentration gradient will form, giving a bulk electric field.

C.4.8 Evaporation and condensation

Evaporation, such as that encountered during drying of a salt solution without stirring, will lead to a concentration gradient normal to the surface, where concentration at the surface is higher

249 than in the bulk. Furthermore, evaporation from a droplet, in which the different regions of the droplet have different thicknesses, will produce lateral concentration gradients. Such effects may be important in paint drying, spray drying, paper coloring, ink-jet printing, food processing, cosmetics drying, Raman spectroscopy of proteins, DNA orientation and detection, coffee rings, and more138–140. In an evaporating environment, ionic systems tend to produce electrokinetic transport and rearrangement of colloids on the surface which may result into different dispersion or patterns of colloids. Therefore, evaporation can be manipulated to orchestrate charge and size based separation of colloids.

In fact, one might expect a salt gradient even in the absence of a free liquid surface, such as the dehydration of a biological organism when placed in a dry atmosphere or dehydration of food materials for preservation. The presence of diffusiophoresis may alter the distribution of nutrients or other colloidal species in each case.

Similarly, condensation over a salt crystal will lead to a transient salt gradient. Gradients due to condensation would arise when water condenses on a hygroscopic salt such as LiCl, CaCl2 and NaOH or when a cloud droplet condenses on sulfate aerosol which acts as a cloud condensation nucleus (CCN).

Additionally, if there are temperature gradients laterally across a surface, then the evaporation or condensation process will have gradients laterally across the surface. Thus, these processes are capable of creating concentration gradients both lateral and tangential to the surface.

250 C.5 Conclusions

Our survey of situations where significant diffusiophoretic transport may be encountered underscores the need to carefully evaluate where gradients may be generated spontaneously, as a result of the dynamics of that specific system. When a system has 1) a concentration gradient due to any number of physical processes (e.g., reactions, dissolution, precipitation, evaporation, condensation), 2) ions with different diffusion coefficients, and 3) particles or surfaces with finite zeta potentials, the system can have diffusiophoretic transport. Varying salt types or surface charge is often a useful test in concluding whether an observed effect is caused due to diffusiophoresis.

Diffusiophoresis is a non-equilibrium phenomenon, but sometimes a system is in a steady state out of equilibrium, as in the case of haloclines which maintain a steady gradient between different layers of the sea and on the surface of salt rejecting reverse osmosis membranes, where energy is being input into the system. Thus, the transient nature does not mean that diffusiophoresis is always short-lived. Moreover, often short-lived transient gradients can lead to significant flows and thus be instrumental in deciding the fate of the system, as we expect for flows in mineral replacement reactions.

Diffusiophoresis was first predicted in 1947, yet its application has been limited. Looking into the future, diffusiophoresis may become an important tool for several industries such as mining, oil exploration, nuclear waste disposal, membrane purification, biological processes, medical treatments, and many more natural and artificial processes. The prime advantage of using diffusiophoresis for colloidal manipulation over other external fields is that it has a broad range of accessibility, be it membranes, geological reservoirs or the deep seas. Moreover, a concentration gradient is often a feature of the system, which can be tuned in a number of ways through changes in composition, temperature or flow rate. In the future we can envision situations where specially-

251 designed colloidal particles sense and are propelled by naturally generated gradients, which leads them to their target such as oil or toxic wastes.

A charged micron or nanoscale particle in water may go from casual Brownian diffusion into a directed motion regime due to diffusiophoresis if it is simply present in a concentration gradient of the right salt. At colloidal length scales, the regime of Brownian diffusion and creeping flows, diffusioosmosis can create convective flows and mixing, speeding up heat and mass transport processes and significantly changing the dynamics of the system. Advances in measurement of zeta potentials at high ionic strength and in diffusiophoretic theory calls for a fresh reconsideration of its applicability or inadvertent presence in a large number of commonly encountered systems at high salinity, remarkably those encountered in marine and geological environments. Finally, we note that while this article has focused on electrolyte systems, non- electrolyte diffusiophoretic transport also exists, and can operate whenever a chemical gradient occurs.

C.6 Acknowledgments

DV thanks the National Science Foundation for funding this work through MSREC, DMR-08-

20404. We thank our collaborators in the MRSEC, especially Ayusman Sen and Tom Mallouk, for many helpful discussions. DV and MK also acknowledge funding from the Penn State College of Engineering’s ENGINE program.

C.7 Contributions

The origins of concentration gradients through molecular exclusion, externally-imposed, sedimentation and screening and evaporation and condensation were written primarily by Rajarshi

252 Guha. The parts on chemical reaction, diffusive mixing, salt or crystal dissolution and molecular crystallization were written primarily by me. All the authors contributed to editing these parts and writing the rest of the paper.

C.8 Copyright Notice

Adapted from the article D. Velegol, A. Garg, R. Guha, A. Kar, M. Kumar, Origins of

Concentration Gradients for Diffusiophoresis. Soft Matter, 12, 4686 (2016), with permission from The Royal Society of Chemistry.

C.9 References

(1) Bird, R. B.; Stewart, W. E.; Lightfoot, E. N. Transport Phenomena; John Wiley & Sons,

2007.

(2) Morrison, F. . Electrophoresis of a Particle of Arbitrary Shape. J. Colloid Interface Sci. 1970,

34, 210–214.

(3) Anderson, J. Colloid Transport By Interfacial Forces. Annu. Rev. Fluid Mech. 1989, 21, 61–

99.

(4) Derjaguin, B.V.; Sidorenkov, G.P.; Zubashchenkov, E.A.; Kiseleva, E. V. Kinetic

Phenomena in Boundary Films of Liquids. Kolloidn. Zh. 1947, 9, 335–347.

(5) Ebel, J. P.; Anderson, J. L.; Prieve, D. C. Diffusiophoresis of Latex Particles in Electrolyte

Gradients. Langmuir 1988, 4, 396–406.

(6) Prieve, D. C.; Anderson, J. L.; Ebel, J. P.; Lowell, M. E. Motion of a Particle Generated by

253 Chemical Gradients. Part 2. Electrolytes. J. Fluid Mech. 1984, 148, 247.

(7) Kar, A.; Guha, R.; Dani, N.; Velegol, D.; Kumar, M. Particle Deposition on Microporous

Membranes Can Be Enhanced or Reduced by Salt Gradients. Langmuir 2014, 30, 793–799.

(8) Guha, R.; Shang, X.; Zydney, A. L.; Velegol, D.; Kumar, M. Diffusiophoresis Contributes

Significantly to Colloidal Fouling in Low Salinity Reverse Osmosis Systems. J. Memb. Sci.

2015, 479, 67–76.

(9) Howse, J.; Jones, R.; Ryan, A.; Gough, T.; Vafabakhsh, R.; Golestanian, R. Self-Motile

Colloidal Particles: From Directed Propulsion to Random Walk. Phys. Rev. Lett. 2007, 99,

48102.

(10) Chaturvedi, N.; Hong, Y.; Sen, A.; Velegol, D. Magnetic Enhancement of Phototaxing

Catalytic Motors. Langmuir 2010, 26, 6308–6313.

(11) Ibele, M.; Mallouk, T. E.; Sen, A. Schooling Behavior of Light-Powered Autonomous

Micromotors in Water. Angew. Chem. Int. Ed. Engl. 2009, 48, 3308–3312.

(12) Duan, W.; Liu, R.; Sen, A. Transition between Collective Behaviors of Micromotors in

Response to Different Stimuli. J. Am. Chem. Soc. 2013, 135, 1280–1283.

(13) Hong, Y.; Diaz, M.; Córdova-Fteueroa, U. M.; Sen, A. Light-Driven Titanium-Dioxide-

Based Reversible Microfireworks and Micromotor/micropump Systems. Adv. Funct. Mater.

2010, 20, 1568–1576.

(14) Kar, A.; Chiang, T.; Rivera, I. O.; Sen, A.; Velegol, D. Enhanced Transport into and out of

Dead-End Pores. 2015, 746–753.

(15) Abécassis, B.; Cottin-Bizonne, C.; Ybert, C.; Ajdari, A.; Bocquet, L. Boosting Migration of

Large Particles by Solute Contrasts. Nat. Mater. 2008, 7, 785–789.

254 (16) Smith, R. E.; Prieve, D. C. Accelerated Deposition of Latex Particles onto a Rapidly

Dissolving Steel Surface. Chem. Eng. Sci. 1982, 37, 1213–1223.

(17) McDermott, J. J.; Kar, A.; Daher, M.; Klara, S.; Wang, G.; Sen, A.; Velegol, D. Self-

Generated Diffusioosmotic Flows from Calcium Carbonate Micropumps. Langmuir 2012,

28, 15491–15497.

(18) Yadav, V.; Freedman, J. D.; Grinstaff, M.; Sen, A. Bone-Crack Detection, Targeting, and

Repair Using Ion Gradients. Angew. Chemie - Int. Ed. 2013, 52, 10997–11001.

(19) Yadav, V.; Pavlick, R. A.; Meckler, S. M.; Sen, A. Triggered Detection and Deposition:

Toward the Repair of Microcracks. Chem. Mater. 2014, 26, 4647–4652.

(20) Kar, A. Diffusiophoresis in Dead-End Pores; Ph.D. Thesis, Pennsylvania State University,

2015.

(21) Marlow, B. J.; Rowell, R. L. Sedimentation Potential in Aqueous Electrolytes. Langmuir

1985, 1, 83–90.

(22) Raşa, M.; Philipse, A. P. Evidence for a Macroscopic Electric Field in the Sedimentation

Profiles of Charged Colloids. Nature 2004, 429, 857–860.

(23) Lin, M. M.-J.; Prieve, D. C. Electromigration of Latex Induced by a Salt Gradient. J. Colloid

Interface Sci. 1983, 95, 327–339.

(24) Lechnick, W. J.; Shaeiwitz, J. A. Measurement of Diffusiophoresis in Liquids. J. Colloid

Interface Sci. 1984, 102, 71–87.

(25) Lechnick, W. J.; Shaeiwitz, J. A. Electrolyte Concentration Dependence of Diffusiophoresis

in Liquids. J. Colloid Interface Sci. 1985, 104, 456–470.

(26) Staffeld, P. O.; Quinn, J. A. Diffusion-Induced Banding of Colloid Particles via

255 Diffusiophoresis: 1. Electrolytes. J. Colloid Interface Sci. 1989, 130, 69–87.

(27) Staffeld, P. O.; Quinn, J. A. Diffusion-Induced Banding of Colloid Particles via

Diffusiophoresis 1. Electrolytes. J. Colloid Interface Sci. 1989, 130, 69–87.

(28) Sharifi-Mood, N.; Koplik, J.; Maldarelli, C. Molecular Dynamics Simulation of the Motion

of Colloidal Nanoparticles in a Solute Concentration Gradient and a Comparison to the

Continuum Limit. Phys. Rev. Lett. 2013, 111, 184501.

(29) Sharifi-Mood, N.; Koplik, J.; Maldarelli, C. Diffusiophoretic Self-Propulsion of Colloids

Driven by a Surface Reaction: The Sub-Micron Particle Regime for Exponential and van

Der Waals Interactions. Phys. Fluids 2013, 25, 12001.

(30) Golestanian, R.; Liverpool, T. B.; Ajdari, A. Propulsion of a Molecular Machine by

Asymmetric Distribution of Reaction Products. Phys. Rev. Lett. 2005, 94, 1–4.

(31) Zhang, H.; Yeung, K.; Robbins, J. S.; Pavlick, R. a.; Wu, M.; Liu, R.; Sen, A.; Phillips, S.

T. Self-Powered Microscale Pumps Based on Analyte-Initiated Depolymerization

Reactions. Angew. Chemie - Int. Ed. 2012, 51, 2400–2404.

(32) Hong, Y.; Velegol, D.; Chaturvedi, N.; Sen, A. Biomimetic Behavior of Synthetic Particles:

From Microscopic Randomness to Macroscopic Control. Phys. Chem. Chem. Phys. 2010,

12, 1423–1435.

(33) Hunter, R. Zeta Potential in Colloid Science: Principles and Applications; Academic Press

Limited: London, 1981.

(34) Anderson, J. L. Effect of Nonuniform Zeta Potential on Particle Movement in Electric Fields.

J. Colloid Interface Sci. 1985, 105, 45–54.

(35) Anderson, J. L. Effect of Nonuniform Zeta Potential on Particle Movement in Electric Fields.

J. Colloid Interface Sci. 1985, 105, 45–54.

256 (36) Fair, M. .; Anderson, J. . Electrophoresis of Nonuniformly Charged Ellipsoidal Particles. J.

Colloid Interface Sci. 1989, 127, 388–400.

(37) Chiang, T.-Y.; Velegol, D. Multi-Ion Diffusiophoresis. J. Colloid Interface Sci. 2014, 424,

120–123.

(38) Flury, M.; Gimmi, T. F. Solute Diffusion. Methods of Soil Analysis, 2002, 1323–1351.

(39) Prieve, D. C.; Anderson, J. L.; Ebel, J. P.; Lowell, M. E. Motion of a Particle Generated by

Chemical Gradients. Part 2. Electrolytes. J. Fluid Mech. 1984, 148, 247–269.

(40) Brady, J. F. Particle Motion Driven by Solute Gradients with Application to Autonomous

Motion: Continuum and Colloidal Perspectives. J. Fluid Mech. 2011, 667, 216–259.

(41) Batchelor, G. K. Diffusion in a Dilute Polydisperse System of Interacting Spheres. J. Fluid

Mech. 1983, 131, 155.

(42) Huang, P. Y.; Keh, H. J. Diffusiophoresis of a Spherical Soft Particle in Electrolyte

Gradients. J. Phys. Chem. B 2012, 116, 7575–7589.

(43) Tseng, S.; Chung, Y.-C.; Hsu, J.-P. Diffusiophoresis of a Soft, pH-Regulated Particle in a

Solution Containing Multiple Ionic Species. J. Colloid Interface Sci. 2015, 438, 196–203.

(44) Huang, H. Y.; Keh, H. J. Diffusiophoresis in Suspensions of Charged Porous Particles. J.

Phys. Chem. B 2015, 119, 2040–2050.

(45) Wei, Y. K.; Keh, H. J. Diffusiophoretic Mobility of Charged Porous Spheres in Electrolyte

Gradients. J. Colloid Interface Sci. 2004, 269, 240–250.

(46) Keh, H. J.; Wei, Y. K. Diffusiophoretic Mobility of Spherical Particles at Low Potential and

Arbitrary Double-Layer Thickness. Langmuir 2000, 16, 5289–5294.

(47) Luu, X.-C.; Hsu, J.-P.; Tseng, S. Diffusiophoresis of a Nonuniformly Charged Sphere in an

257 Electrolyte Solution. J. Chem. Phys. 2011, 134, 64708.

(48) Liu, K.-L.; Hsu, J.-P.; Hsu, W.-L.; Yeh, L.-H.; Tseng, S. Diffusiophoresis of a

Polyelectrolyte in a Salt Concentration Gradient. Electrophoresis 2012, 33, 1068–1078.

(49) Tseng, S.; Lin, D.-H.; Hsu, J.-P. Influence of Double-Layer Polarization and Chemiosmosis

on the Diffusiophoresis of a Non-Spherical Polyelectrolyte. J. Colloid Interface Sci. 2015,

446, 272–281.

(50) Tseng, S.; Su, C.-Y.; Hsu, J.-P. Diffusiophoresis of a pH-Regulated, Zwitterionic

Polyelectrolyte in a Solution Containing Multiple Ionic Species. Chem. Eng. Sci. 2014, 118,

164–172.

(51) Bikerman, J. J. Structure and Capacity of Electrical Double Layer. Philos. Mag. 1942, 33,

384–397.

(52) Nakayama, Y.; Andelman, D. Differential Capacitance of the Electric Double Layer: The

Interplay between Ion Finite Size and Dielectric Decrement. J. Chem. Phys. 2015, 142.

(53) Quesada-Pérez, M.; González-Tovar, E.; Martín-Molina, A.; Lozada-Cassou, M.; Hidalgo-

Álvarez, R. Overcharging in Colloids: Beyond the Poisson-Boltzmann Approach.

ChemPhysChem 2003, 4, 234–248.

(54) van der Heyden, F.; Stein, D.; Besteman, K.; Lemay, S.; Dekker, C. Charge Inversion at

High Ionic Strength Studied by Streaming Currents. Phys. Rev. Lett. 2006, 96, 224502.

(55) Parsons, D. F.; Boström, M.; Maceina, T. J.; Salis, A.; Ninham, B. W. Why Direct or

Reversed Hofmeister Series? Interplay of Hydration, Non-Electrostatic Potentials, and Ion

Size. Langmuir 2010, 26, 3323–3328.

(56) Borukhov, I.; Andelman, D.; Aviv, T. Steric Effects in Electrolytes: A Modi Ed Poisson-

Boltzmann Equation. 1997, 435–438.

258 (57) Derjaguin, B. V; Dukhin, S. S.; Koptelova, M. M. Capillary Osmosis through Porous

Partitions and Properties of Boundary Layers of Solutions. J. Colloid Interface Sci. 1972,

38, 584–595.

(58) Derjaguin, B. V; Churaev, N. V; Muller, V. M. Surface Forces in Transport Phenomena. In

Surface Forces SE - 11; Springer US, 1987; pp. 369–431.

(59) Denisov, G. A.; Kaluta, V. K.; Nikolaev, E. V; Tishchenko, G. A.; Shataeva, L. K. Modeling

of Coupled Trasport of Ions and Zwitterions across Porous Ion Exchange Membranes. J.

Memb. Sci. 1993, 79, 211–226.

(60) Dukhin, S. S. Non-Equilibrium Electric Surface Phenomena. Adv. Colloid Interface Sci.

1993, 44, 1–134.

(61) Ramon, G. Z.; Feinberg, B. J.; Hoek, E. M. V. Membrane-Based Production of Salinity-

Gradient Power. Energy Environ. Sci. 2011, 4, 4423–4434.

(62) Logan, B. E.; Elimelech, M. Membrane-Based Processes for Sustainable Power Generation

Using Water. Nature 2012, 488, 313–319.

(63) Siria, A.; Poncharal, P.; Biance, A.-L.; Fulcrand, R.; Blase, X.; Purcell, S. T.; Bocquet, L.

Giant Osmotic Energy Conversion Measured in a Single Transmembrane Boron Nitride

Nanotube. Nature 2013, 494, 455–458.

(64) Hoek, E. M. V; Elimelech, M. Cake-Enhanced Concentration Polarization: A New Fouling

Mechanism for Salt-Rejecting Membranes. Environ. Sci. Technol. 2003, 37, 5581–5588.

(65) Serrano, C. G.; McDermott, J. J.; Velegol, D. Sediments of Soft Spheres Arranged by

Effective Density. Nat Mater 2011, 10, 716–721.

(66) Florea, D.; Musa, S.; Huyghe, J. M. R.; Wyss, H. M. Long-Range Repulsion of Colloids

Driven by Ion Exchange and Diffusiophoresis. Proc. Natl. Acad. Sci. U. S. A. 2014, 111,

259 6554–6559.

(67) Langford Jr., J. F.; Xu, X.; Yao, Y.; Maloney, S. F.; Lenhoff, A. M. Chromatography of

Proteins on Charge-Variant Ion Exchangers and Implications for Optimizing Protein Uptake

Rates. J. Chromatogr. A 2007, 1163, 190–202.

(68) Gundersen, S. I.; Palmer, A. F. Conjugation of Methoxypolyethylene Glycol to the Surface

of Bovine Red Blood Cells. Biotechnol. Bioeng. 2007, 96, 1199–1210.

(69) McAfee, M. S.; Annunziata, O. Effect of Particle Size on Salt-Induced Diffusiophoresis

Compared to Brownian Mobility. Langmuir 2014, 30, 4916–4923.

(70) Mcafee, M. S.; Zhang, H.; Annunziata, O. Amplification of Salt-Induced Polymer

Diffusiophoresis by Increasing Salting-Out Strength. Langmuir 2014, 30, 12210–12219.

(71) Vagenende, V.; Han, a X.; Mueller, M.; Trout, B. L. Protein-Associated Cation Clusters in

Aqueous Arginine Solutions and Their Effects on Protein Stability and Size. ACS Chem.

Biol. 2013, 8, 416–422.

(72) McAfee, M. S.; Annunziata, O. Effects of Salting-In Interactions on Macromolecule

Diffusiophoresis and Salt Osmotic Diffusion. Langmuir 2015, 31, 1353–1361.

(73) Golestanian, R.; Liverpool, T. B.; Ajdari, A. Designing Phoretic Micro- and Nano-

Swimmers. New J. Phys. 2007, 9, 126–126.

(74) Yake, A. M.; Panella, R. a; Snyder, C. E.; Velegol, D. Fabrication of Colloidal Doublets by

a Salting out-Quenching-Fusing Technique. Langmuir 2006, 22, 9135–9141.

(75) Ibisate, M.; Zou, Z.; Xia, Y. Arresting, Fixing, and Separating Dimers Composed of

Uniform Silica Colloidal Spheres. Adv. Funct. Mater. 2006, 16, 1627–1632.

(76) Jiang, S.; Chen, Q.; Tripathy, M.; Luijten, E.; Schweizer, K. S.; Granick, S. Janus Particle

260 Synthesis and Assembly. Adv. Mater. 2010, 22, 1060–1071.

(77) Palacci, J.; Sacanna, S.; Steinberg, A. P.; Pine, D. J.; Chaikin, P. M. Living Crystals of

Light-Activated Colloidal Surfers. Science 2013, 339, 936–940.

(78) Ebbens, S.; Gregory, D. a.; Dunderdale, G.; Howse, J. R.; Ibrahim, Y.; Liverpool, T. B.;

Golestanian, R. Electrokinetic Effects in Catalytic Platinum-Insulator Janus Swimmers.

EPL (Europhysics Lett. 2014, 106, 58003.

(79) Brown, A.; Poon, W. Ionic Effects in Self-Propelled Pt-Coated Janus Swimmers. Soft

Matter 2014, 4016–4027.

(80) Kline, T. R.; Iwata, J.; Lammert, P. E.; Mallouk, T. E.; Sen, A.; Velegol, D. Catalytically

Driven Colloidal Patterning and Transport. J. Phys. Chem. B 2006, 110, 24513–24521.

(81) Diao, J.; Young, L.; Kim, S.; Fogarty, E. a; Heilman, S. M.; Zhou, P.; Shuler, M. L.; Wu,

M.; DeLisa, M. P. A Three-Channel Microfluidic Device for Generating Static Linear

Gradients and Its Application to the Quantitative Analysis of Bacterial Chemotaxis. Lab

Chip 2006, 6, 381–388.

(82) Kagan, D.; Balasubramanian, S.; Wang, J. Chemically Triggered Swarming of Gold

Microparticles. Angew. Chem. Int. Ed. Engl. 2011, 50, 503–506.

(83) Kim, Y.; Shah, A. A.; Solomon, M. J. Spatially and Temporally Reconfigurable Assembly

of Colloidal Crystals. Nat Commun 2014, 5.

(84) Saha, S.; Golestanian, R.; Ramaswamy, S. Clusters, Asters, and Collective Oscillations in

Chemotactic Colloids. Phys. Rev. E - Stat. Nonlinear, Soft Matter Phys. 2014, 89, 1–25.

(85) Mozaffari, A.; Sharifi-Mood, N.; Koplik, J.; Maldarelli, C. Self-Diffusiophoretic Colloidal

Propulsion Near a Solid Boundary. 2015.

261 (86) Ibrahim, Y.; Liverpool, T. B. The Dynamics of a Self-Phoretic Janus Swimmer near a Wall.

EPL (Europhysics Lett. 2015, 111, 48008.

(87) Das, S.; Garg, A.; Campbell, A. I.; Howse, J. R.; Sen, A.; Velegol, D.; Golestanian, R.;

Ebbens, S. J. Boundaries Can Steer Active Janus Spheres. Nat. Commun. 2015, 6, 1–10.

(88) Uspal, W. E.; Popescu, M. N.; Dietrich, S.; Tasinkevych, M. Self-Propulsion of a

Catalytically Active Particle near a Planar Wall: From Reflection to Sliding and Hovering.

Soft Matter 2014, 1, 1–15.

(89) Sen, A.; Ibele, M.; Hong, Y.; Velegol, D. Chemo and Phototactic Nano/microbots. Faraday

Discuss. 2009, 143, 15.

(90) Buttinoni, I.; Bialké, J.; Kümmel, F.; Löwen, H.; Bechinger, C.; Speck, T. Dynamical

Clustering and Phase Separation in Suspensions of Self-Propelled Colloidal Particles. Phys.

Rev. Lett. 2013, 110, 238301.

(91) Hong, Y.; Blackman, N. M. K.; Kopp, N. D.; Sen, A.; Velegol, D. Chemotaxis of

Nonbiological Colloidal Rods. Phys. Rev. Lett. 2007, 99, 1–4.

(92) Lager, A.; Webb, K. J.; Collins, I. R.; Richmond, D. M. LoSal Enhanced Oil Recovery:

Evidence of Enhanced Oil Recovery at the Reservoir Scale. In SPE Symposium on Improved

Oil Recovery; Society of Petroleum Engineers, 2013.

(93) Madirolas, a.; Acha, E. M.; Guerrero, R. a.; Lasta, C. Sources of Acoustic Scattering near

a Halocline in an Estuarine Frontal System. Sci. Mar. 1997, 61, 431–438.

(94) Tiselius, P.; Nielsen, G.; Nielsen, T. G. Microscale Patchiness of Plankton within a Sharp

Pycnocline. J. Plankton Res. 1994, 16, 543–554.

(95) Samuelsson, M. Interannual Salinity Variations in the Baltic Sea during the Period 1954-

1990. Cont. Shelf Res. 1996, 16, 1463–1477.

262 (96) Lougee, L. A.; Bollens, S. M.; Avent, S. R. The Effects of Haloclines on the Vertical

Distribution and Migration of Zooplankton. J. Exp. Mar. Bio. Ecol. 2002, 278, 111–134.

(97) Bjoerk, G.; Liungman, O.; Rydberg, L. Net Circulation and Salinity Variations in an Open-

Ended Swedish Fjord System . Estuaries 2000, 23, 367–380.

(98) Stigebrandt, A.; Aure, J. Vertical Mixing in Basin Waters of Fjords. Journal of Physical

Oceanography, 1989, 19, 917–926.

(99) Dror, I.; Amitay, T.; Yaron, B.; Berkowitz, B. Salt-Pump Mechanism for Contaminant

Intrusion into Coastal Aquifers. Science 2003, 300, 950.

(100) Anderson, J. L.; Prieve, D. C.; Ebel, J. P. Chemically Induced Migration of Particles Across

Fluid Streamlines. Chem. Eng. Commun. 1987, 55, 211–224.

(101) Chiang, T. Y.; Velegol, D. Multi-Ion Diffusiophoresis. J. Colloid Interface Sci. 2014, 424,

120–123.

(102) Kar, A. Diffusiophoresis in Dead-End Pores, 2015.

(103) Palacci, J.; Abécassis, B.; Cottin-Bizonne, C.; Ybert, C.; Bocquet, L. Colloidal Motility and

Pattern Formation under Rectified Diffusiophoresis. Phys. Rev. Lett. 2010, 104, 1–4.

(104) Wanunu, M.; Morrison, W.; Rabin, Y.; Grosberg, A. Y.; Meller, A. Electrostatic Focusing

of Unlabelled DNA into Nanoscale Pores Using a Salt Gradient. Nat Nano 2010, 5, 160–

165.

(105) Joo, S. W.; Lee, S. Y.; Liu, J.; Qian, S. Diffusiophoresis of an Elongated Cylindrical

Nanoparticle along the Axis of a Nanopore. Chemphyschem 2010, 11, 3281–3290.

(106) Palacci, J.; Cottin-Bizonne, C.; Ybert, C.; Bocquet, L. Osmotic Traps for Colloids and

Macromolecules Based on Logarithmic Sensing in Salt Taxis. Soft Matter 2012, 8, 980.

263 (107) Maeda, Y. T.; Tlusty, T.; Libchaber, a. Effects of Long DNA Folding and Small RNA

Stem-Loop in Thermophoresis. Proc. Natl. Acad. Sci. 2012, 109, 17972–17977.

(108) Maeda, Y. T. (2+1)-Dimensional Manipulation of Soft Biological Materials by Opto-

Thermal Diffusiophoresis. Appl. Phys. Lett. 2013, 103.

(109) Brandenburg, B.; Zhuang, X. Virus Trafficking - Learning from Single-Virus Tracking. Nat.

Rev. Microbiol. 2007, 5, 197–208.

(110) Seksek, O.; Bolard, J. Nuclear pH Gradient in Mammalian Cells Revealed by Laser

Microspectrofluorimetry. J. Cell Sci. 1996, 109 ( Pt 1, 257–262.

(111) Sodeik, B.; Ebersold, M. W.; Helenius, a. Microtubule-Mediated Transport of Incoming

Herpes Simplex Virus 1 Capsids to the Nucleus. J. Cell Biol. 1997, 136, 1007–1021.

(112) Kosmulski, M.; Matuevi, E. Solvophoresis of Latex. J. Colloid Interface Sci. 1992, 150,

291–294.

(113) Paustian, J. S.; Angulo, C. D.; Nery-Azevedo, R.; Shi, N.; Abdel-Fattah, A. I.; Squires, T.

M. Direct Measurements of Colloidal Solvophoresis under Imposed Solvent and Solute

Gradients. Langmuir 2015, 31, 4402–4410.

(114) Derjaguin, B. V; Golovanov, M. V. On Long-Range Forces of Repulsion between

Biological Cells. Prog. Surf. Sci. 1992, 40, 210–217.

(115) Fukuyama, T.; Fuke, A.; Mochizuki, M.; Kamei, K.; Maeda, Y. T. Directing and Boosting

of Cell Migration by the Entropic Force Gradient in Polymer Solution. Langmuir 2015,

acs.langmuir.5b02559.

(116) Elimelech, M.; O’Melia, C. R. Effect of Electrolyte Type on the Electrophoretic Mobility

of Polystyrene Latex Colloids. Colloids and Surfaces 1990, 44, 165–178.

264 (117) Russel, W. B.; Saville, D. A.; Schowalter, W. R. Colloidal Dispersions; Cambridge

University Press, 1992.

(118) Putnis, A. Mineral Replacement Reactions: From Macroscopic Observations to

Microscopic Mechanisms. Mineral. Mag. 2002, 66, 689–708.

(119) Putnis, A. Mineral Replacement Reactions: From Macroscopic Observations to

Microscopic Mechanisms. Mineral. Mag. 2002, 66, 689–708.

(120) Vinogradov, J.; Jaafar, M. Z.; Jackson, M. D. Measurement of Streaming Potential Coupling

Coefficient in Sandstones Saturated with Natural and Artificial Brines at High Salinity. J.

Geophys. Res. 2010, 115, B12204.

(121) Kosmulski, M.; Rosenholm, J. B. High Ionic Strength Electrokinetics. Adv. Colloid

Interface Sci. 2004, 112, 93–107.

(122) Dukhin, a; Dukhin, S.; Goetz, P. Electrokinetics at High Ionic Strength and Hypothesis of

the Double Layer with Zero Surface Charge. Langmuir 2005, 21, 9990–9997.

(123) Prieto, M.; Putnis, A.; Fernandez-Diaz, L. Factors Controlling the Kinetics of

Crystallization: Supersaturation Evolution in a Porous Medium. Application to Barite

Crystallization. Geol. Mag. 1990, 127, 485.

(124) Putnis, A.; Prieto, M.; Fernandez-Diaz, L. Fluid Supersaturation and Crystallization in

Porous Media. Geol. Mag. 1995, 132, 1.

(125) De Yoreo, J. J.; Gilbert, P. U. P. A.; Sommerdijk, N. A. J. M.; Penn, R. L.; Whitelam, S.;

Joester, D.; Zhang, H.; Rimer, J. D.; Navrotsky, A.; Banfield, J. F.; et al. Crystallization by

Particle Attachment in Synthetic, Biogenic, and Geologic Environments. Science (80-. ).

2015, 349, aaa6760-aaa6760.

(126) Nielsen, M. H.; Aloni, S.; De Yoreo, J. J. In Situ TEM Imaging of CaCO₃ Nucleation

265 Reveals Coexistence of Direct and Indirect Pathways. Science 2014, 345, 1158–1162.

(127) Kashchiev, D. Thermodynamically Consistent Description of the Work to Form a Nucleus

of Any Size. J. Chem. Phys. 2003, 118, 1837.

(128) Bromley, L. A.; Cottier, D.; Davey, R. J.; Dobbs, B.; Smith, S.; Heywood, B. R. Interactions

at the Organic/inorganic Interface: Molecular Design of Crystallization Inhibitors for Barite.

Langmuir 1993, 9, 3594–3599.

(129) Kowacz, M.; Putnis, A. The Effect of Specific Background Electrolytes on Water Structure

and Solute Hydration: Consequences for Crystal Dissolution and Growth. Geochim.

Cosmochim. Acta 2008, 72, 4476–4487.

(130) Sawdy, A.; Heritage, A.; Pel, L. A Review of Salt Transport in Porous Media: Assessment

Methods and Salt Reduction Treatments. In Salt Weathering on Buildings and Stone

Sculptures, 22–24 October 2008, The National Museum Copenhagen, Denmark; 2008; pp.

1–27.

(131) De Yoreo, J. J.; Qiu, S. R.; Hoyer, J. R. Molecular Modulation of Calcium Oxalate

Crystallization. Am. J. Physiol. Renal Physiol. 2006, 291, F1123–F1131.

(132) Sakhaee, K.; Nicar, M.; Hill, K.; Pak, C. Y. Contrasting Effects of Potassium Citrate and

Sodium Citrate Therapies on Urinary Chemistries and Crystallization of Stone-Forming

Salts. Kidney Int. 1983, 24, 348–352.

(133) Eicken, H. The Role of Sea Ice in Structuring Antarctic Ecosystems. Polar Biol. 1992, 12,

3–13.

(134) Thomas, D. N.; Dieckmann, G. S. Antarctic Sea Ice--a Habitat for Extremophiles. Science

2002, 295, 641–644.

(135) Piazza, R.; Bellini, T.; Degiorgio, V. Equilibrium Sedimentation Profiles of Screened

266 Charged Colloids: A Test of the Hard-Sphere Equation of State. Phys. Rev. Lett. 1993, 71,

4267–4270.

(136) Warren, P. Metrology: Electrifying Effects in Colloids. Nature 2004, 429, 822.

(137) Schmidt, N. J.; Lennette, E. H.; Dennis, J. Density Gradient Centrifugation Studies on

Rubella Complement-Fixing Antigens. J. Immunol. 1967, 99, 399–405.

(138) Smalyukh, I. I.; Zribi, O. V.; Butler, J. C.; Lavrentovich, O. D.; Wong, G. C. L. Structure

and Dynamics of Liquid Crystalline Pattern Formation in Drying Droplets of DNA. Phys.

Rev. Lett. 2006, 96, 177801.

(139) Zhang, D.; Xie, Y.; Mrozek, M. F.; Ortiz, C.; Davisson, V. J.; Ben-Amotz, D. Raman

Detection of Proteomic Analytes. Anal. Chem. 2003, 75, 5703–5709.

(140) Wong, T. S.; Chen, T. H.; Shen, X.; Ho, C. M. Nanochromatography Driven by the Coffee

Ring Effect. Anal. Chem. 2011, 83, 1871–1873.

267

VITA

Astha Garg Education Doctor of Philosophy (PhD), Chemical Engineering May 2017 Pennsylvania State University, University Park, PA Advisor: Prof. Darrell Velegol

Bachelor of Technology, Chemical Engineering; May 2011 Minor: Energy Science and Engineering Indian Institute of Technology, Bombay (India)

Publications 1. S. Das, A. Garg, A. I. Campbell, J. Howse, A. Sen, D. Velegol, R. Golestanian and S. J. Ebbens, Boundaries can Steer Active Janus Spheres. Nature Communications, 6:8999 (2015). 2. A. Garg, C. Cartier, K. Bishop, D. Velegol, Particle Zeta Potentials Remain Finite in Saturated Salt Solutions. Langmuir, 32, 11837 (2016). 3. D. Velegol, A. Garg, R. Guha, A. Kar, M. Kumar, Origins of Concentration Gradients for Diffusiophoresis. Soft Matter, 12, 4686 (2016). 4. A. Garg, C. Gorski, D. Velegol, Chemical Micro-fracking : Creating porosity in calcite using pseudomorphic mineral replacement (To be submitted to Environmental Science & Technology, May 2017). 5. A. Garg, J. Wu, I. Al’Abri, C. Gorski, D. Velegol, Controlling Pseudomorphic Mineral Replacement through Transport (To be submitted to Chemical Geology, May 2017).

Presentations 1. A. Garg, C. Cartier, K. Bishop, D. Velegol, Particle zeta potentials remain finite in saturated salt solutions. ACS National Meeting 2017 (Talk). 2. A. Garg, C. Cartier, K. Bishop, D. Velegol, Finite Zeta Potential at High Ionic Strength. ACS Colloid and Surface Science Symposium, 2016 (Talk).

Work Experience Teaching Assistant, Penn State University Jan 2014 – Jul 2016 Research Intern, BASF SE, Ludwigshafen, Germany Oct 2011 – Apr 2012

Awards Gordon Research Seminar Travel Award (2016) Prevention of Accidents with Safety (PAWS) Award by Chemical Engineering, Penn State (2013)

Professional Affiliation Member of American Chemical Society