ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/iceWater2/iceWater4.tex
Revision: 4268
Committed: Wed Dec 31 17:25:13 2014 UTC (9 years, 8 months ago) by gezelter
Content type: application/x-tex
File size: 51470 byte(s)
Log Message:
JCP version now that PNAS said no

File Contents

# Content
1 %% ****** Start of file template.aps ****** %
2 %%
3 %%
4 %% This file is part of the APS files in the REVTeX 4 distribution.
5 %% Version 4.0 of REVTeX, August 2001
6 %%
7 %%
8 %% Copyright (c) 2001 The American Physical Society.
9 %%
10 %% See the REVTeX 4 README file for restrictions and more information.
11 %%
12 %
13 % This is a template for producing manuscripts for use with REVTEX 4.0
14 % Copy this file to another name and then work on that file.
15 % That way, you always have this original template file to use.
16 %
17 % Group addresses by affiliation; use superscriptaddress for long
18 % author lists, or if there are many overlapping affiliations.
19 % For Phys. Rev. appearance, change preprint to twocolumn.
20 % Choose pra, prb, prc, prd, pre, prl, prstab, or rmp for journal
21 % Add 'draft' option to mark overfull boxes with black boxes
22 % Add 'showpacs' option to make PACS codes appear
23 %\documentclass[aps,jcp,twocolumn,showpacs,superscriptaddress,groupedaddress]{revtex4} % for review and submission
24 \documentclass[aps,jcp,preprint,showpacs,superscriptaddress,groupedaddress]{revtex4} % for double-spaced preprint
25 \usepackage{graphicx} % needed for figures
26 \usepackage{dcolumn} % needed for some tables
27 \usepackage{bm} % for math
28 \usepackage{amssymb} % for math
29 %\usepackage{booktabs}
30 \usepackage{multirow}
31 \usepackage{tablefootnote}
32 \usepackage{times}
33 \usepackage{mathptm}
34 \usepackage[version=3]{mhchem}
35
36
37
38 \begin{document}
39
40 \title{The different facets of ice have different hydrophilicities: \\
41 Friction at water / ice-I$_\mathrm{h}$ interfaces}
42 \author{Patrick B. Louden}
43 \author{J. Daniel Gezelter}
44 \email{gezelter@nd.edu}
45 \affiliation{Department of Chemistry and Biochemistry, University of
46 Notre Dame, Notre Dame, IN 46556}
47
48 \date{\today}
49
50
51 \begin{abstract}
52 We present evidence that the prismatic and secondary prism facets of
53 ice-I$_\mathrm{h}$ crystals possess structural features that can
54 reduce the effective hydrophilicity of the ice/water interface. The
55 spreading dynamics of liquid water droplets on ice facets exhibits
56 long-time behavior that differs for the prismatic
57 $\{10\bar{1}0\}$ and secondary prism $\{11\bar{2}0\}$ facets
58 when compared with the basal $\{0001\}$ and pyramidal
59 $\{20\bar{2}1\}$ facets. We also present the results of
60 simulations of solid-liquid friction of the same four crystal facets
61 being drawn through liquid water, and find that the two prismatic
62 facets exhibit roughly half the solid-liquid friction of the basal
63 and pyramidal facets. These simulations provide evidence that the
64 two prismatic faces have a significantly smaller effective surface
65 area in contact with the liquid water. The ice / water interfacial
66 widths for all four crystal facets are similar (using both
67 structural and dynamic measures), and were found to be independent
68 of the shear rate. Additionally, decomposition of orientational
69 time correlation functions show position-dependence for the short-
70 and longer-time decay components close to the interface.
71 \end{abstract}
72
73 \pacs{68.08.Bc, 68.08.De, 66.20.Cy}
74 \keywords{ice; water; interfaces; hydrophobicity}
75 \maketitle
76
77 \section{Introduction}
78 Surfaces can be characterized as hydrophobic or hydrophilic
79 based on the strength of the interactions with water. Hydrophobic
80 surfaces do not have strong enough interactions with water to overcome
81 the internal attraction between molecules in the liquid phase, and the
82 degree of hydrophilicity of a surface can be described by the extent a
83 droplet can spread out over the surface. The contact angle, $\theta$,
84 formed between the solid and the liquid depends on the free energies
85 of the three interfaces involved, and is given by Young's
86 equation~\cite{Young05},
87 \begin{equation}\label{young}
88 \cos\theta = (\gamma_{sv} - \gamma_{sl})/\gamma_{lv} .
89 \end{equation}
90 Here $\gamma_{sv}$, $\gamma_{sl}$, and $\gamma_{lv}$ are the free
91 energies of the solid/vapor, solid/liquid, and liquid/vapor interfaces,
92 respectively. Large contact angles, $\theta > 90^{\circ}$, correspond
93 to hydrophobic surfaces with low wettability, while small contact
94 angles, $\theta < 90^{\circ}$, correspond to hydrophilic surfaces.
95 Experimentally, measurements of the contact angle of sessile drops is
96 often used to quantify the extent of wetting on surfaces with
97 thermally selective wetting
98 characteristics~\cite{Tadanaga00,Liu04,Sun04}.
99
100 Nanometer-scale structural features of a solid surface can influence
101 the hydrophilicity to a surprising degree. Small changes in the
102 heights and widths of nano-pillars can change a surface from
103 superhydrophobic, $\theta \ge 150^{\circ}$, to hydrophilic, $\theta
104 \sim 0^{\circ}$~\cite{Koishi09}. This is often referred to as the
105 Cassie-Baxter to Wenzel transition. Nano-pillared surfaces with
106 electrically tunable Cassie-Baxter and Wenzel states have also been
107 observed~\cite{Herbertson06,Dhindsa06,Verplanck07,Ahuja08,Manukyan11}.
108 Luzar and coworkers have modeled these transitions on nano-patterned
109 surfaces~\cite{Daub07,Daub10,Daub11,Ritchie12}, and have found the
110 change in contact angle is due to the field-induced perturbation of
111 hydrogen bonding at the liquid/vapor interface~\cite{Daub07}.
112
113 One would expect the interfaces of ice to be highly hydrophilic (and
114 possibly the most hydrophilic of all solid surfaces). In this paper we
115 present evidence that some of the crystal facets of ice-I$_\mathrm{h}$
116 have structural features that can reduce the effective hydrophilicity.
117 Our evidence for this comes from molecular dynamics (MD) simulations
118 of the spreading dynamics of liquid droplets on these facets, as well
119 as reverse non-equilibrium molecular dynamics (RNEMD) simulations of
120 solid-liquid friction.
121
122 Quiescent ice-I$_\mathrm{h}$/water interfaces have been studied
123 extensively using computer simulations. Hayward and Haymet
124 characterized and measured the widths of these
125 interfaces~\cite{Hayward01,Hayward02}. Nada and Furukawa have also
126 modeled the width of basal/water and prismatic/water
127 interfaces~\cite{Nada95} as well as crystal restructuring at
128 temperatures approaching the melting point~\cite{Nada00}.
129
130 The surface of ice exhibits a pre-melting layer, often called a
131 quasi-liquid layer (QLL), at temperatures near the melting point. MD
132 simulations of the facets of ice-I$_\mathrm{h}$ exposed to vacuum have
133 found QLL widths of approximately 10 \AA\ at 3 K below the melting
134 point~\cite{Conde08}. Similarly, Limmer and Chandler have used the mW
135 water model~\cite{Molinero09} and statistical field theory to estimate
136 QLL widths at similar temperatures to be about 3 nm~\cite{Limmer14}.
137
138 Recently, Sazaki and Furukawa have developed a technique using laser
139 confocal microscopy combined with differential interference contrast
140 microscopy that has sufficient spatial and temporal resolution to
141 visualize and quantitatively analyze QLLs on ice crystals at
142 temperatures near melting~\cite{Sazaki10}. They have found the width of
143 the QLLs perpendicular to the surface at -2.2$^{o}$C to be 3-4 \AA\
144 wide. They have also seen the formation of two immiscible QLLs, which
145 displayed different dynamics on the crystal surface~\cite{Sazaki12}.
146
147 % There is now significant interest in the \textit{tribological}
148 % properties of ice/ice and ice/water interfaces in the geophysics
149 % community. Understanding the dynamics of solid-solid shearing that is
150 % mediated by a liquid layer~\cite{Cuffey99, Bell08} will aid in
151 % understanding the macroscopic motion of large ice
152 % masses~\cite{Casassa91, Sukhorukov13, Pritchard12, Lishman13}.
153
154 Using molecular dynamics simulations, Samadashvili has recently shown
155 that when two smooth ice slabs slide past one another, a stable
156 liquid-like layer develops between them~\cite{Samadashvili13}. In a
157 previous study, our RNEMD simulations of ice-I$_\mathrm{h}$ shearing
158 through liquid water have provided quantitative estimates of the
159 solid-liquid kinetic friction coefficients~\cite{Louden13}. These
160 displayed a factor of two difference between the basal and prismatic
161 facets. The friction was found to be independent of shear direction
162 relative to the surface orientation. We attributed facet-based
163 difference in liquid-solid friction to the 6.5 \AA\ corrugation of the
164 prismatic face which reduces the effective surface area of the ice
165 that is in direct contact with liquid water.
166
167 In the sections that follow, we describe the simulations of
168 droplet-spreading dynamics using standard MD as well as simulations of
169 tribological properties using RNEMD. These simulations give
170 complementary results that point to the prismatic and secondary prism
171 facets having roughly half of their surface area in direct contact
172 with the liquid.
173
174 \section{Methodology}
175 \subsection{Construction of the Ice / Water interfaces}
176 Ice I$_\mathrm{h}$ crystallizes in the hexagonal space group
177 P$6_3/mmc$, and common ice crystals form hexagonal plates with the
178 basal face, $\{0001\}$, forming the top and bottom of each plate, and
179 the prismatic facet, $\{10\bar{1}0\}$, forming the sides. In extreme
180 temperatures or low water saturation conditions, ice crystals can
181 easily form as hollow columns, needles and dendrites. These are
182 structures that expose other crystalline facets of the ice to the
183 surroundings. Among the more common facets are the secondary prism,
184 $\{11\bar{2}0\}$, and pyramidal, $\{20\bar{2}1\}$, faces.
185
186 We found it most useful to work with proton-ordered, zero-dipole
187 crystals that expose strips of dangling H-atoms and lone
188 pairs~\cite{Buch:2008fk}. Our structures were created starting from
189 Structure 6 of Hirsch and Ojam\"{a}e's set of orthorhombic
190 representations for ice-I$_{h}$~\cite{Hirsch04}. This crystal
191 structure was cleaved along the four different facets. The exposed
192 face was reoriented normal to the $z$-axis of the simulation cell, and
193 the structures were extended to form large exposed facets in
194 rectangular box geometries. Liquid water boxes were created with
195 identical dimensions (in $x$ and $y$) as the ice, with a $z$ dimension
196 of three times that of the ice block, and a density corresponding to 1
197 g / cm$^3$. Each of the ice slabs and water boxes were independently
198 equilibrated at a pressure of 1 atm, and the resulting systems were
199 merged by carving out any liquid water molecules within 3 \AA\ of any
200 atoms in the ice slabs. Each of the combined ice/water systems were
201 then equilibrated at 225K, which is the liquid-ice coexistence
202 temperature for SPC/E water~\cite{Bryk02}. Reference
203 \citealp{Louden13} contains a more detailed explanation of the
204 construction of similar ice/water interfaces. The resulting dimensions
205 as well as the number of ice and liquid water molecules contained in
206 each of these systems are shown in Table \ref{tab:method}.
207
208 The SPC/E water model~\cite{Berendsen87} has been extensively
209 characterized over a wide range of liquid
210 conditions~\cite{Arbuckle02,Kuang12}, and its phase diagram has been
211 well studied~\cite{Baez95,Bryk04b,Sanz04b,Fennell:2005fk}. With longer
212 cutoff radii and careful treatment of electrostatics, SPC/E mostly
213 avoids metastable crystalline morphologies like
214 ice-\textit{i}~\cite{Fennell:2005fk} and ice-B~\cite{Baez95}. The
215 free energies and melting
216 points~\cite{Baez95,Arbuckle02,Gay02,Bryk02,Bryk04b,Sanz04b,Fennell:2005fk,Fernandez06,Abascal07,Vrbka07}
217 of various other crystalline polymorphs have also been calculated.
218 Haymet \textit{et al.} have studied quiescent Ice-I$_\mathrm{h}$/water
219 interfaces using the SPC/E water model, and have seen structural and
220 dynamic measurements of the interfacial width that agree well with
221 more expensive water models, although the coexistence temperature for
222 SPC/E is still well below the experimental melting point of real
223 water~\cite{Bryk02}. Given the extensive data and speed of this model,
224 it is a reasonable choice even though the temperatures required are
225 somewhat lower than real ice / water interfaces.
226
227 \section{Droplet Simulations}
228 Ice surfaces with a thickness of $\sim$~20~\AA\ were created as
229 described above, but were not solvated in a liquid box. The crystals
230 were then replicated along the $x$ and $y$ axes (parallel to the
231 surface) until a large surface ($>$ 126 nm\textsuperscript{2}) had
232 been created. The sizes and numbers of molecules in each of the
233 surfaces is given in Table S1. Weak translational restraining
234 potentials with spring constants of 1.5~$\mathrm{kcal\
235 mol}^{-1}\mathrm{~\AA}^{-2}$ (prismatic and pyramidal facets) or
236 4.0~$\mathrm{kcal\ mol}^{-1}\mathrm{~\AA}^{-2}$ (basal facet) were
237 applied to the centers of mass of each molecule in order to prevent
238 surface melting, although the molecules were allowed to reorient
239 freely. A water droplet containing 2048 SPC/E molecules was created
240 separately. Droplets of this size can produce agreement with the Young
241 contact angle extrapolated to an infinite drop size~\cite{Daub10}. The
242 surfaces and droplet were independently equilibrated to 225~K, at
243 which time the droplet was placed 3-5~\AA\ above the surface. Five
244 statistically independent simulations were carried out for each facet,
245 and the droplet was placed at unique $x$ and $y$ locations for each of
246 these simulations. Each simulation was 5~ns in length and was
247 conducted in the microcanonical (NVE) ensemble. Representative
248 configurations for the droplet on the prismatic facet are shown in
249 figure \ref{fig:Droplet}.
250
251 \section{Shearing Simulations (Interfaces in Bulk Water)}
252 To perform the shearing simulations, the velocity shearing and scaling
253 variant of reverse non-equilibrium molecular dynamics (VSS-RNEMD) was
254 employed \cite{Kuang12}. This method performs a series of simultaneous
255 non-equilibrium exchanges of linear momentum and kinetic energy
256 between two physically-separated regions of the simulation cell. The
257 system responds to this unphysical flux with velocity and temperature
258 gradients. When VSS-RNEMD is applied to bulk liquids, transport
259 properties like the thermal conductivity and the shear viscosity are
260 easily extracted assuming a linear response between the flux and the
261 gradient. At the interfaces between dissimilar materials, the same
262 method can be used to extract \textit{interfacial} transport
263 properties (e.g. the interfacial thermal conductance and the
264 hydrodynamic slip length).
265
266 The kinetic energy flux (producing a thermal gradient) is necessary
267 when performing shearing simulations at the ice-water interface in
268 order to prevent the frictional heating due to the shear from melting
269 the crystal. Reference \citealp{Louden13} provides more details on the
270 VSS-RNEMD method as applied to ice-water interfaces. A representative
271 configuration of the solvated prismatic facet being sheared through
272 liquid water is shown in figure \ref{fig:Shearing}.
273
274 All simulations were performed using OpenMD~\cite{OOPSE,openmd}, with
275 a time step of 2 fs and periodic boundary conditions in all three
276 dimensions. Electrostatics were handled using the damped-shifted
277 force real-space electrostatic kernel~\cite{Ewald}.
278
279 The interfaces were equilibrated for a total of 10 ns at equilibrium
280 conditions before being exposed to either a shear or thermal gradient.
281 This consisted of 5 ns under a constant temperature (NVT) integrator
282 set to 225~K followed by 5 ns under a microcanonical (NVE) integrator.
283 Weak thermal gradients were allowed to develop using the VSS-RNEMD
284 (NVE) integrator using a small thermal flux ($-2.0\times 10^{-6}$
285 kcal/mol/\AA$^2$/fs) for a duration of 5 ns to allow the gradient to
286 stabilize. The resulting temperature gradient was $\approx$ 10K over
287 the entire box length, which was sufficient to keep the temperature at
288 the interface within $\pm 1$ K of the 225~K target.
289
290 Velocity gradients were then imposed using the VSS-RNEMD (NVE)
291 integrator with a range of momentum fluxes. The systems were divided
292 into 100 bins along the $z$-axis for the VSS-RNEMD moves, which were
293 attempted every time step. Although computationally expensive, this
294 was done to minimize the magnitude of each individual momentum
295 exchange. Because individual VSS-RNEMD exchange moves conserve both
296 total energy and linear momentum, the method can be ``bolted-on'' to
297 simulations in any ensemble. The simulations of the pyramidal
298 interface were performed under the canonical (NVT) ensemble. When
299 time correlation functions were computed, the RNEMD simulations were
300 done in the microcanonical (NVE) ensemble. All simulations of the
301 other interfaces were carried out in the microcanonical ensemble.
302
303 These gradients were allowed to stabilize for 1~ns before data
304 collection started. Once established, four successive 0.5~ns runs were
305 performed for each shear rate. During these simulations,
306 configurations of the system were stored every 1~ps, and statistics on
307 the structure and dynamics in each bin were accumulated throughout the
308 simulations. Although there was some small variation in the measured
309 interfacial width between succcessive runs, no indication of bulk
310 melting or crystallization was observed. That is, no large scale
311 changes in the positions of the top and bottom interfaces occurred
312 during the simulations.
313
314 \section{Results}
315 \subsection{Ice - Water Contact Angles}
316 To determine the extent of wetting for each of the four crystal
317 facets, contact angles for liquid droplets on the ice surfaces were
318 computed using two methods. In the first method, the droplet is
319 assumed to form a spherical cap, and the contact angle is estimated
320 from the $z$-axis location of the droplet's center of mass
321 ($z_\mathrm{cm}$). This procedure was first described by Hautman and
322 Klein~\cite{Hautman91}, and was utilized by Hirvi and Pakkanen in
323 their investigation of water droplets on polyethylene and poly(vinyl
324 chloride) surfaces~\cite{Hirvi06}. For each stored configuration, the
325 contact angle, $\theta$, was found by inverting the expression for the
326 location of the droplet center of mass,
327 \begin{equation}\label{contact_1}
328 \langle z_\mathrm{cm}\rangle = 2^{-4/3}R_{0}\bigg(\frac{1-cos\theta}{2+cos\theta}\bigg)^{1/3}\frac{3+cos\theta}{2+cos\theta} ,
329 \end{equation}
330 where $R_{0}$ is the radius of the free water droplet.
331
332 In addition to the spherical cap method outlined above, a second
333 method for obtaining the contact angle was described by Ruijter,
334 Blake, and Coninck~\cite{Ruijter99}. This method uses a cylindrical
335 averaging of the droplet's density profile. A threshold density of
336 0.5 g cm\textsuperscript{-3} is used to estimate the location of the
337 edge of the droplet. The $r$ and $z$-dependence of the droplet's edge
338 is then fit to a circle, and the contact angle is computed from the
339 intersection of the fit circle with the $z$-axis location of the solid
340 surface. Again, for each stored configuration, the density profile in
341 a set of annular shells was computed. Due to large density
342 fluctuations close to the ice, all shells located within 2 \AA\ of the
343 ice surface were left out of the circular fits. The height of the
344 solid surface ($z_\mathrm{suface}$) along with the best fitting origin
345 ($z_\mathrm{droplet}$) and radius ($r_\mathrm{droplet}$) of the
346 droplet can then be used to compute the contact angle,
347 \begin{equation}
348 \theta = 90 + \frac{180}{\pi} \sin^{-1}\left(\frac{z_\mathrm{droplet} -
349 z_\mathrm{surface}}{r_\mathrm{droplet}} \right).
350 \end{equation}
351 Both methods provided similar estimates of the dynamic contact angle,
352 although the spherical cap method is significantly less prone to
353 noise, and is the method used to compute the contact angles in table
354 \ref{tab:kappa}.
355
356 Because the initial droplet was placed above the surface, the initial
357 value of 180$^{\circ}$ decayed over time (See figure 1 in the
358 SI). Each of these profiles were fit to a
359 biexponential decay, with a short-time contribution ($\tau_c$) that
360 describes the initial contact with the surface, a long time
361 contribution ($\tau_s$) that describes the spread of the droplet over
362 the surface, and a constant ($\theta_\infty$) to capture the
363 infinite-time estimate of the equilibrium contact angle,
364 \begin{equation}
365 \theta(t) = \theta_\infty + (180-\theta_\infty) \left[ a e^{-t/\tau_c} +
366 (1-a) e^{-t/\tau_s} \right]
367 \end{equation}
368 We have found that the rate for water droplet spreading across all
369 four crystal facets, $k_\mathrm{spread} = 1/\tau_s \approx$ 0.7
370 ns$^{-1}$. However, the basal and pyramidal facets produced estimated
371 equilibrium contact angles, $\theta_\infty \approx$ 35$^{\circ}$, while
372 prismatic and secondary prismatic had values for $\theta_\infty$ near
373 43$^{\circ}$ as seen in Table \ref{tab:kappa}.
374
375 These results indicate that by traditional measures, the basal and
376 pyramidal facets are more hydrophilic than the prismatic and secondary
377 prism facets, and surprisingly, that the differential hydrophilicities
378 of the crystal facets is not reflected in the spreading rate of the
379 droplet.
380
381 % This is in good agreement with our calculations of friction
382 % coefficients, in which the basal
383 % and pyramidal had a higher coefficient of kinetic friction than the
384 % prismatic and secondary prismatic. Due to this, we believe that the
385 % differences in friction coefficients can be attributed to the varying
386 % hydrophilicities of the facets.
387
388 \subsection{Solid-liquid friction of the interfaces}
389 In a bulk fluid, the shear viscosity, $\eta$, can be determined
390 assuming a linear response to a shear stress,
391 \begin{equation}\label{Shenyu-11}
392 j_{z}(p_{x}) = \eta \frac{\partial v_{x}}{\partial z}.
393 \end{equation}
394 Here $j_{z}(p_{x})$ is the flux (in $x$-momentum) that is transferred
395 in the $z$ direction (i.e. the shear stress). The RNEMD simulations
396 impose an artificial momentum flux between two regions of the
397 simulation, and the velocity gradient is the fluid's response. This
398 technique has now been applied quite widely to determine the
399 viscosities of a number of bulk fluids~\cite{Muller99,Bordat02,Cavalcanti07}.
400
401 At the interface between two phases (e.g. liquid / solid) the same
402 momentum flux creates a velocity difference between the two materials,
403 and this can be used to define an interfacial friction coefficient
404 ($\kappa$),
405 \begin{equation}\label{Shenyu-13}
406 j_{z}(p_{x}) = \kappa \left[ v_{x}(liquid) - v_{x}(solid) \right]
407 \end{equation}
408 where $v_{x}(solid)$ is the velocity of the solid and $v_{x}(liquid)$
409 is the velocity of the liquid measured at the hydrodynamic boundary
410 layer.
411
412 The simulations described here contain significant quantities of both
413 liquid and solid phases, and the momentum flux must traverse a region
414 of the liquid that is simultaneously under a thermal gradient. Since
415 the liquid has a temperature-dependent shear viscosity, $\eta(T)$,
416 estimates of the solid-liquid friction coefficient can be obtained if
417 one knows the viscosity of the liquid at the interface (i.e. at the
418 melting temperature, $T_m$),
419 \begin{equation}\label{kappa-2}
420 \kappa = \frac{\eta(T_{m})}{\left[v_{x}(fluid)-v_{x}(solid)\right]}\left(\frac{\partial v_{x}}{\partial z}\right).
421 \end{equation}
422 For SPC/E, the melting temperature of Ice-I$_\mathrm{h}$ is estimated
423 to be 225~K~\cite{Bryk02}. To obtain the value of
424 $\eta(225\mathrm{~K})$ for the SPC/E model, a $31.09 \times 29.38
425 \times 124.39$ \AA\ box with 3744 water molecules in a disordered
426 configuration was equilibrated to 225~K, and five
427 statistically-independent shearing simulations were performed (with
428 imposed fluxes that spanned a range of $3 \rightarrow 13
429 \mathrm{~m~s}^{-1}$ ). Each simulation was conducted in the
430 microcanonical ensemble with total simulation times of 5 ns. The
431 VSS-RNEMD exchanges were carried out every 2 fs. We estimate
432 $\eta(225\mathrm{~K})$ to be 0.0148 $\pm$ 0.0007 Pa s for SPC/E,
433 roughly ten times larger than the shear viscosity previously computed
434 at 280~K~\cite{Kuang12}.
435
436 The interfacial friction coefficient can equivalently be expressed as
437 the ratio of the viscosity of the fluid to the hydrodynamic slip
438 length, $\kappa = \eta / \delta$. The slip length is an indication of
439 strength of the interactions between the solid and liquid phases,
440 although the connection between slip length and surface hydrophobicity
441 is not yet clear. In some simulations, the slip length has been found
442 to have a link to the effective surface
443 hydrophobicity~\cite{Sendner:2009uq}, although Ho \textit{et al.} have
444 found that liquid water can also slip on hydrophilic
445 surfaces~\cite{Ho:2011zr}. Experimental evidence for a direct tie
446 between slip length and hydrophobicity is also not
447 definitive. Total-internal reflection particle image velocimetry
448 (TIR-PIV) studies have suggested that there is a link between slip
449 length and effective
450 hydrophobicity~\cite{Lasne:2008vn,Bouzigues:2008ys}. However, recent
451 surface sensitive cross-correlation spectroscopy (TIR-FCCS)
452 measurements have seen similar slip behavior for both hydrophobic and
453 hydrophilic surfaces~\cite{Schaeffel:2013kx}.
454
455 In each of the systems studied here, the interfacial temperature was
456 kept fixed to 225~K, which ensured the viscosity of the fluid at the
457 interace was identical. Thus, any significant variation in $\kappa$
458 between the systems is a direct indicator of the slip length and the
459 effective interaction strength between the solid and liquid phases.
460
461 The calculated $\kappa$ values found for the four crystal facets of
462 Ice-I$_\mathrm{h}$ are shown in Table \ref{tab:kappa}. The basal and
463 pyramidal facets were found to have similar values of $\kappa \approx
464 6$ ($\times 10^{-4} \mathrm{amu~\AA}^{-2}\mathrm{fs}^{-1}$), while the
465 prismatic and secondary prism facets exhibited $\kappa \approx 3$
466 ($\times 10^{-4} \mathrm{amu~\AA}^{-2}\mathrm{fs}^{-1}$). These
467 results are also essentially independent of the direction of the shear
468 relative to channels on the surfaces of the facets. The friction
469 coefficients indicate that the basal and pyramidal facets have
470 significantly stronger interactions with liquid water than either of
471 the two prismatic facets. This is in agreement with the contact angle
472 results above - both of the high-friction facets exhibited smaller
473 contact angles, suggesting that the solid-liquid friction (and inverse
474 slip length) is correlated with the hydrophilicity of these facets.
475
476 \subsection{Structural measures of interfacial width under shear}
477 One of the open questions about ice/water interfaces is whether the
478 thickness of the 'slush-like' quasi-liquid layer (QLL) depends on the
479 facet of ice presented to the water. In the QLL region, the water
480 molecules are ordered differently than in either the solid or liquid
481 phases, and also exhibit distinct dynamical behavior. The width of
482 this quasi-liquid layer has been estimated by finding the distance
483 over which structural order parameters or dynamic properties change
484 from their bulk liquid values to those of the solid ice. The
485 properties used to find interfacial widths have included the local
486 density, the diffusion constant, and both translational and
487 orientational order
488 parameters~\cite{Karim88,Karim90,Hayward01,Hayward02,Bryk02,Gay02,Louden13}.
489
490 The VSS-RNEMD simulations impose thermal and velocity gradients.
491 These gradients perturb the momenta of the water molecules, so
492 parameters that depend on translational motion are often measuring the
493 momentum exchange, and not physical properties of the interface. As a
494 structural measure of the interface, we have used the local
495 tetrahedral order parameter, which measures the match of the local
496 molecular environments (e.g. the angles between nearest neighbor
497 molecules) to perfect tetrahedral ordering. This quantity was
498 originally described by Errington and Debenedetti~\cite{Errington01}
499 and has been used in bulk simulations by Kumar \textit{et
500 al.}~\cite{Kumar09} It has previously been used in ice/water
501 interfaces by by Bryk and Haymet~\cite{Bryk04b}.
502
503 To determine the structural widths of the interfaces under shear, each
504 of the systems was divided into 100 bins along the $z$-dimension, and
505 the local tetrahedral order parameter (Eq. 5 in Reference
506 \citealp{Louden13}) was time-averaged in each bin for the duration of
507 the shearing simulation. The spatial dependence of this order
508 parameter, $q(z)$, is the tetrahedrality profile of the interface.
509 The lower panels in figures 2-5 in the supporting information show
510 tetrahedrality profiles (in circles) for each of the four interfaces.
511 The $q(z)$ function has a range of $(0,1)$, where a value of unity
512 indicates a perfectly tetrahedral environment. The $q(z)$ for the
513 bulk liquid was found to be $\approx~0.77$, while values of
514 $\approx~0.92$ were more common in the ice. The tetrahedrality
515 profiles were fit using a hyperbolic tangent function (see Eq. 6 in
516 Reference \citealp{Louden13}) designed to smoothly fit the bulk to ice
517 transition while accounting for the weak thermal gradient. In panels
518 $b$ and $c$ of the same figures, the resulting thermal and velocity
519 gradients from an imposed kinetic energy and momentum fluxes can be
520 seen. The vertical dotted lines traversing these figures indicate the
521 midpoints of the interfaces as determined by the tetrahedrality
522 profiles. The hyperbolic tangent fit provides an estimate of
523 $d_\mathrm{struct}$, the structural width of the interface.
524
525 We find the interfacial width to be $3.2 \pm 0.2$ \AA\ (pyramidal) and
526 $3.2 \pm 0.2$ \AA\ (secondary prism) for the control systems with no
527 applied momentum flux. This is similar to our previous results for the
528 interfacial widths of the quiescent basal ($3.2 \pm 0.4$ \AA) and
529 prismatic systems ($3.6 \pm 0.2$ \AA).
530
531 Over the range of shear rates investigated, $0.4 \rightarrow
532 6.0~\mathrm{~m~s}^{-1}$ for the pyramidal system and $0.6 \rightarrow
533 5.2~\mathrm{~m~s}^{-1}$ for the secondary prism, we found no
534 significant change in the interfacial width. The mean interfacial
535 widths are collected in table \ref{tab:kappa}. This follows our
536 previous findings of the basal and prismatic systems, in which the
537 interfacial widths of the basal and prismatic facets were also found
538 to be insensitive to the shear rate~\cite{Louden13}.
539
540 The similarity of these interfacial width estimates indicate that the
541 particular facet of the exposed ice crystal has little to no effect on
542 how far into the bulk the ice-like structural ordering persists. Also,
543 it appears that for the shearing rates imposed in this study, the
544 interfacial width is not structurally modified by the movement of
545 water over the ice.
546
547 \subsection{Dynamic measures of interfacial width under shear}
548 The spatially-resolved orientational time correlation function,
549 \begin{equation}\label{C(t)1}
550 C_{2}(z,t)=\langle P_{2}(\mathbf{u}_i(0)\cdot \mathbf{u}_i(t))
551 \delta(z_i(0) - z) \rangle,
552 \end{equation}
553 provides local information about the decorrelation of molecular
554 orientations in time. Here, $P_{2}$ is the second-order Legendre
555 polynomial, and $\mathbf{u}_i$ is the molecular vector that bisects
556 the HOH angle of molecule $i$. The angle brackets indicate an average
557 over all the water molecules and time origins, and the delta function
558 restricts the average to specific regions. In the crystal, decay of
559 $C_2(z,t)$ is incomplete, while in the liquid, correlation times are
560 typically measured in ps. Observing the spatial-transition between
561 the decay regimes can define a dynamic measure of the interfacial
562 width.
563
564 To determine the dynamic widths of the interfaces under shear, each of
565 the systems was divided into bins along the $z$-dimension ($\approx$ 3
566 \AA\ wide) and $C_2(z,t)$ was computed using only those molecules that
567 were in the bin at the initial time. To compute these correlation
568 functions, each of the 0.5 ns simulations was followed by a shorter
569 200 ps microcanonical (NVE) simulation in which the positions and
570 orientations of every molecule in the system were recorded every 0.1
571 ps.
572
573 The time-dependence was fit to a triexponential decay, with three time
574 constants: $\tau_{short}$, measuring the librational motion of the
575 water molecules, $\tau_{middle}$, measuring the timescale for breaking
576 and making of hydrogen bonds, and $\tau_{long}$, corresponding to the
577 translational motion of the water molecules. An additional constant
578 was introduced in the fits to describe molecules in the crystal which
579 do not experience long-time orientational decay.
580
581 In Figures 6-9 in the supporting information, the $z$-coordinate
582 profiles for the three decay constants, $\tau_{short}$,
583 $\tau_{middle}$, and $\tau_{long}$ for the different interfaces are
584 shown. (Figures 6 \& 7 are new results, and Figures 8 \& 9 are
585 updated plots from Ref \citealp{Louden13}.) In the liquid regions of
586 all four interfaces, we observe $\tau_{middle}$ and $\tau_{long}$ to
587 have approximately consistent values of $3-6$ ps and $30-40$ ps,
588 respectively. Both of these times increase in value approaching the
589 interface. Approaching the interface, we also observe that
590 $\tau_{short}$ decreases from its liquid-state value of $72-76$ fs.
591 The approximate values for the decay constants and the trends
592 approaching the interface match those reported previously for the
593 basal and prismatic interfaces.
594
595 We have estimated the dynamic interfacial width $d_\mathrm{dyn}$ by
596 fitting the profiles of all the three orientational time constants
597 with an exponential decay to the bulk-liquid behavior,
598 \begin{equation}\label{tauFit}
599 \tau(z)\approx\tau_{liquid}+(\tau_{wall}-\tau_{liquid})e^{-(z-z_{wall})/d_\mathrm{dyn}}
600 \end{equation}
601 where $\tau_{liquid}$ and $\tau_{wall}$ are the liquid and projected
602 wall values of the decay constants, $z_{wall}$ is the location of the
603 interface, as measured by the structural order parameter. These
604 values are shown in table \ref{tab:kappa}. Because the bins must be
605 quite wide to obtain reasonable profiles of $C_2(z,t)$, the error
606 estimates for the dynamic widths of the interface are significantly
607 larger than for the structural widths. However, all four interfaces
608 exhibit dynamic widths that are significantly below 1~nm, and are in
609 reasonable agreement with the structural width above.
610
611 \section{Conclusions}
612 In this work, we used MD simulations to measure the advancing contact
613 angles of water droplets on the basal, prismatic, pyramidal, and
614 secondary prism facets of Ice-I$_\mathrm{h}$. Although we saw no
615 significant change in the \textit{rate} at which the droplets spread
616 over the surface, the long-time behavior predicts larger equilibrium
617 contact angles for the two prismatic facets.
618
619 We have also used RNEMD simulations of water interfaces with the same
620 four crystal facets to compute solid-liquid friction coefficients. We
621 have observed coefficients of friction that differ by a factor of two
622 between the two prismatic facets and the basal and pyramidal facets.
623 Because the solid-liquid friction coefficient is directly tied to the
624 inverse of the hydrodynamic slip length, this suggests that there are
625 significant differences in the overall interaction strengths between
626 these facets and the liquid layers immediately in contact with them.
627
628 The agreement between these two measures have lead us to conclude that
629 the two prismatic facets have a lower hydrophilicity than either the
630 basal or pyramidal facets. One possible explanation of this behavior
631 is that the face presented by both prismatic facets consists of deep,
632 narrow channels (i.e. stripes of adjacent rows of pairs of
633 hydrogen-bound water molecules). At the surfaces of these facets,
634 the channels are 6.35 \AA\ wide and the sub-surface ice layer is 2.25
635 \AA\ below (and therefore blocked from hydrogen bonding with the
636 liquid). This means that only 1/2 of the surface molecules can form
637 hydrogen bonds with liquid-phase molecules.
638
639 In the basal plane, the surface features are shallower (1.3 \AA), with
640 no blocked subsurface layer. The pyramidal face has much wider
641 channels (8.65 \AA) which are also quite shallow (1.37 \AA). These
642 features allow liquid phase molecules to form hydrogen bonds with all
643 of the surface molecules in the basal and pyramidal facets. This
644 means that for similar surface areas, the two prismatic facets have an
645 effective hydrogen bonding surface area of half of the basal and
646 pyramidal facets. The reduction in the effective surface area would
647 explain much of the behavior observed in our simulations.
648
649 \begin{acknowledgments}
650 Support for this project was provided by the National
651 Science Foundation under grant CHE-1362211. Computational time was
652 provided by the Center for Research Computing (CRC) at the
653 University of Notre Dame.
654 \end{acknowledgments}
655 \bibliographystyle{aip}
656 \newpage
657 \bibliography{iceWater}
658 % *****************************************
659 % There is significant interest in the properties of ice/ice and ice/water
660 % interfaces in the geophysics community. Most commonly, the results of shearing
661 % two ice blocks past one
662 % another\cite{Casassa91, Sukhorukov13, Pritchard12, Lishman13} or the shearing
663 % of ice through water\cite{Cuffey99, Bell08}. Using molecular dynamics
664 % simulations, Samadashvili has recently shown that when two smooth ice slabs
665 % slide past one another, a stable liquid-like layer develops between
666 % them\cite{Samadashvili13}. To fundamentally understand these processes, a
667 % molecular understanding of the ice/water interfaces is needed.
668
669 % Investigation of the ice/water interface is also crucial in understanding
670 % processes such as nucleation, crystal
671 % growth,\cite{Han92, Granasy95, Vanfleet95} and crystal
672 % melting\cite{Weber83, Han92, Sakai96, Sakai96B}. Insight gained to these
673 % properties can also be applied to biological systems of interest, such as
674 % the behavior of the antifreeze protein found in winter
675 % flounder,\cite{Wierzbicki07,Chapsky97} and certain terrestial
676 % arthropods.\cite{Duman:2001qy,Meister29012013} Elucidating the properties which
677 % give rise to these processes through experimental techniques can be expensive,
678 % complicated, and sometimes infeasible. However, through the use of molecular
679 % dynamics simulations much of the problems of investigating these properties
680 % are alleviated.
681
682 % Understanding ice/water interfaces inherently begins with the isolated
683 % systems. There has been extensive work parameterizing models for liquid water,
684 % such as the SPC\cite{Berendsen81}, SPC/E\cite{Berendsen87},
685 % TIP4P\cite{Jorgensen85}, TIP4P/2005\cite{Abascal05},
686 % ($\dots$), and more recently, models for simulating
687 % the solid phases of water, such as the TIP4P/Ice\cite{Abascal05b} model. The
688 % melting point of various crystal structures of ice have been calculated for
689 % many of these models
690 % (SPC\cite{Karim90,Abascal07}, SPC/E\cite{Baez95,Arbuckle02,Gay02,Bryk02,Bryk04b,Sanz04b,Gernandez06,Abascal07,Vrbka07}, TIP4P\cite{Karim88,Gao00,Sanz04,Sanz04b,Koyama04,Wang05,Fernandez06,Abascal07}, TIP5P\cite{Sanz04,Koyama04,Wang05,Fernandez06,Abascal07}),
691 % and the partial or complete phase diagram for the model has been determined
692 % (SPC/E\cite{Baez95,Bryk04b,Sanz04b}, TIP4P\cite{Sanz04,Sanz04b,Koyama04}, TIP5P\cite{Sanz04,Koyama04}).
693 % Knowing the behavior and melting point for these models has enabled an initial
694 % investigation of ice/water interfaces.
695
696 % The Ice-I$_\mathrm{h}$/water quiescent interface has been extensively studied
697 % over the past 30 years by theory and experiment. Haymet \emph{et al.} have
698 % done significant work characterizing and quantifying the width of these
699 % interfaces for the SPC,\cite{Karim90} SPC/E,\cite{Gay02,Bryk02},
700 % CF1,\cite{Hayward01,Hayward02} and TIP4P\cite{Karim88} models for water. In
701 % recent years, Haymet has focused on investigating the effects cations and
702 % anions have on crystal nucleaion and
703 % melting.\cite{Bryk04,Smith05,Wilson08,Wilson10} Nada and Furukawa have studied
704 % the the basal- and prismatic-water interface width\cite{Nada95}, crystal
705 % surface restructuring at temperatures approaching the melting
706 % point\cite{Nada00}, and the mechanism of ice growth inhibition by antifreeze
707 % proteins\cite{Nada08,Nada11,Nada12}. Nada has developed a six-site water model
708 % for ice/water interfaces near the melting point\cite{Nada03}, and studied the
709 % dependence of crystal growth shape on applied pressure\cite{Nada11b}. Using
710 % this model, Nada and Furukawa have established differential
711 % growth rates for the basal, prismatic, and secondary prismatic facets of
712 % Ice-I$_\mathrm{h}$ and found their origins due to a reordering of the hydrogen
713 % bond network in water near the interface\cite{Nada05}. While the work
714 % described so far has mainly focused on bulk water on ice, there is significant
715 % interest in thin films of water on ice surfaces as well.
716
717 % It is well known that the surface of ice exhibits a premelting layer at
718 % temperatures near the melting point, often called a quasi-liquid layer (QLL).
719 % Molecular dynamics simulations of the facets of ice-I$_\mathrm{h}$ exposed
720 % to vacuum performed by Conde, Vega and Patrykiejew have found QLL widths of
721 % approximately 10 \AA\ at 3 K below the melting point\cite{Conde08}.
722 % Similarly, Limmer and Chandler have used course grain simulations and
723 % statistical field theory to estimated QLL widths at the same temperature to
724 % be about 3 nm\cite{Limmer14}.
725 % Recently, Sazaki and Furukawa have developed an experimental technique with
726 % sufficient spatial and temporal resolution to visulaize and quantitatively
727 % analyze QLLs on ice crystals at temperatures near melting\cite{Sazaki10}. They
728 % have found the width of the QLLs perpindicular to the surface at -2.2$^{o}$C
729 % to be $\mathcal{O}$(\AA). They have also seen the formation of two immiscible
730 % QLLs, which displayed different stabilities and dynamics on the crystal
731 % surface\cite{Sazaki12}. Knowledge of the hydrophilicities of each
732 % of the crystal facets would help further our understanding of the properties
733 % and dynamics of the QLLs.
734
735 % Presented here is the follow up to our previous paper\cite{Louden13}, in which
736 % the basal and prismatic facets of an ice-I$_\mathrm{h}$/water interface were
737 % investigated where the ice was sheared relative to the liquid. By using a
738 % recently developed velocity shearing and scaling approach to reverse
739 % non-equilibrium molecular dynamics (VSS-RNEMD), simultaneous temperature and
740 % velocity gradients can be applied to the system, which allows for measurment
741 % of friction and thermal transport properties while maintaining a stable
742 % interfacial temperature\cite{Kuang12}. Structural analysis and dynamic
743 % correlation functions were used to probe the interfacial response to a shear,
744 % and the resulting solid/liquid kinetic friction coefficients were reported.
745 % In this paper we present the same analysis for the pyramidal and secondary
746 % prismatic facets, and show that the differential interfacial friction
747 % coefficients for the four facets of ice-I$_\mathrm{h}$ are determined by their
748 % relative hydrophilicity by means of dynamics water contact angle
749 % simulations.
750
751 % The local tetrahedral order parameter, $q(z)$, is given by
752 % \begin{equation}
753 % q(z) = \int_0^L \sum_{k=1}^{N} \Bigg(1 -\frac{3}{8}\sum_{i=1}^{3}
754 % \sum_{j=i+1}^{4} \bigg(\cos\psi_{ikj}+\frac{1}{3}\bigg)^2\Bigg)
755 % \delta(z_{k}-z)\mathrm{d}z \Bigg/ N_z
756 % \label{eq:qz}
757 % \end{equation}
758 % where $\psi_{ikj}$ is the angle formed between the oxygen sites of molecules
759 % $i$,$k$, and $j$, where the centeral oxygen is located within molecule $k$ and
760 % molecules $i$ and $j$ are two of the closest four water molecules
761 % around molecule $k$. All four closest neighbors of molecule $k$ are also
762 % required to reside within the first peak of the pair distribution function
763 % for molecule $k$ (typically $<$ 3.41 \AA\ for water).
764 % $N_z = \int\delta(z_k - z) \mathrm{d}z$ is a normalization factor to account
765 % for the varying population of molecules within each finite-width bin.
766
767
768 % The hydrophobicity or hydrophilicity of a surface can be described by the
769 % extent a droplet of water wets the surface. The contact angle formed between
770 % the solid and the liquid, $\theta$, which relates the free energies of the
771 % three interfaces involved, is given by Young's equation.
772 % \begin{equation}\label{young}
773 % \cos\theta = (\gamma_{sv} - \gamma_{sl})/\gamma_{lv}
774 % \end{equation}
775 % Here $\gamma_{sv}$, $\gamma_{sl}$, and $\gamma_{lv}$ are the free energies
776 % of the solid/vapor, solid/liquid, and liquid/vapor interfaces respectively.
777 % Large contact angles ($\theta$ $\gg$ 90\textsuperscript{o}) correspond to low
778 % wettability and hydrophobic surfaces, while small contact angles
779 % ($\theta$ $\ll$ 90\textsuperscript{o}) correspond to high wettability and
780 % hydrophilic surfaces. Experimentally, measurements of the contact angle
781 % of sessile drops has been used to quantify the extent of wetting on surfaces
782 % with thermally selective wetting charactaristics\cite{Tadanaga00,Liu04,Sun04},
783 % as well as nano-pillared surfaces with electrically tunable Cassie-Baxter and
784 % Wenzel states\cite{Herbertson06,Dhindsa06,Verplanck07,Ahuja08,Manukyan11}.
785 % Luzar and coworkers have done significant work modeling these transitions on
786 % nano-patterned surfaces\cite{Daub07,Daub10,Daub11,Ritchie12}, and have found
787 % the change in contact angle to be due to the external field perturbing the
788 % hydrogen bonding of the liquid/vapor interface\cite{Daub07}.
789
790 % SI stuff:
791
792 % Correlation functions:
793 % To compute eq. \eqref{C(t)1}, each 0.5 ns simulation was
794 % followed by an additional 200 ps NVE simulation during which the
795 % position and orientations of each molecule were recorded every 0.1 ps.
796
797
798 %\end{article}
799 \newpage
800 \begin{figure}
801 \includegraphics[width=\linewidth]{Droplet}
802 \caption{\label{fig:Droplet} Computational model of a droplet of
803 liquid water spreading over the prismatic $\{10\bar{1}0\}$ facet
804 of ice, before (left) and 2.6 ns after (right) being introduced to the
805 surface. The contact angle ($\theta$) shrinks as the simulation
806 proceeds, and the long-time behavior of this angle is used to
807 estimate the hydrophilicity of the facet.}
808 \end{figure}
809 \newpage
810 \begin{figure}
811 \includegraphics[width=1.9in]{Shearing}
812 \caption{\label{fig:Shearing} Computational model of a slab of ice
813 being sheared through liquid water. In this figure, the ice is
814 presenting two copies of the prismatic $\{10\bar{1}0\}$ facet
815 towards the liquid phase. The RNEMD simulation exchanges both
816 linear momentum (indicated with arrows) and kinetic energy between
817 the central box and the box that spans the cell boundary. The
818 system responds with weak thermal gradient and a velocity profile
819 that shears the ice relative to the surrounding liquid.}
820 \end{figure}
821
822 % \begin{figure}
823 % \includegraphics[width=\linewidth]{ContactAngle}
824 % \caption{\label{fig:ContactAngle} The dynamic contact angle of a
825 % droplet after approaching each of the four ice facets. The decay to
826 % an equilibrium contact angle displays similar dynamics. Although
827 % all the surfaces are hydrophilic, the long-time behavior stabilizes
828 % to significantly flatter droplets for the basal and pyramidal
829 % facets. This suggests a difference in hydrophilicity for these
830 % facets compared with the two prismatic facets.}
831 % \end{figure}
832
833 % \begin{figure}
834 % \includegraphics[width=\linewidth]{Pyr_comic_strip}
835 % \caption{\label{fig:pyrComic} Properties of the pyramidal interface
836 % being sheared through water at 3.8 ms\textsuperscript{-1}. Lower
837 % panel: the local tetrahedral order parameter, $q(z)$, (circles) and
838 % the hyperbolic tangent fit (turquoise line). Middle panel: the
839 % imposed thermal gradient required to maintain a fixed interfacial
840 % temperature of 225 K. Upper panel: the transverse velocity gradient
841 % that develops in response to an imposed momentum flux. The vertical
842 % dotted lines indicate the locations of the midpoints of the two
843 % interfaces.}
844 % \end{figure}
845
846 % \begin{figure}
847 % \includegraphics[width=\linewidth]{SP_comic_strip}
848 % \caption{\label{fig:spComic} The secondary prismatic interface with a shear
849 % rate of 3.5 \
850 % ms\textsuperscript{-1}. Panel descriptions match those in figure \ref{fig:pyrComic}.}
851 % \end{figure}
852
853 % \begin{figure}
854 % \includegraphics[width=\linewidth]{Pyr-orient}
855 % \caption{\label{fig:PyrOrient} The three decay constants of the
856 % orientational time correlation function, $C_2(z,t)$, for water as a
857 % function of distance from the center of the ice slab. The vertical
858 % dashed line indicates the edge of the pyramidal ice slab determined
859 % by the local order tetrahedral parameter. The control (circles) and
860 % sheared (squares) simulations were fit using shifted-exponential
861 % decay (see Eq. 9 in Ref. \citealp{Louden13}).}
862 % \end{figure}
863
864 % \begin{figure}
865 % \includegraphics[width=\linewidth]{SP-orient-less}
866 % \caption{\label{fig:SPorient} Decay constants for $C_2(z,t)$ at the secondary
867 % prismatic face. Panel descriptions match those in \ref{fig:PyrOrient}.}
868 % \end{figure}
869
870 \newpage
871 \begin{table}[h]
872 \centering
873 \caption{Sizes of the droplet and shearing simulations. Cell
874 dimensions are measured in \AA. \label{tab:method}}
875 \begin{tabular}{r|cccc|ccccc}
876 \toprule
877 \multirow{2}{*}{Interface} & \multicolumn{4}{c|}{Droplet} & \multicolumn{5}{c}{
878 Shearing} \\
879 & $N_\mathrm{ice}$ & $N_\mathrm{droplet}$ & $L_x$ & $L_y$ & $N_\mathrm{ice}$ &
880 $N_\mathrm{liquid}$ & $L_x$ & $L_y$ & $L_z$ \\
881 \colrule
882 Basal $\{0001\}$ & 12960 & 2048 & 134.70 & 140.04 & 900 & 1846 & 23.87 & 35.83 & 98.64 \\
883 Pyramidal $\{20\bar{2}1\}$ & 11136 & 2048 & 143.75 & 121.41 & 1216 & 2203 & 37.47 & 29.50 & 93.02 \\
884 Prismatic $\{10\bar{1}0\}$ & 9900 & 2048 & 110.04 & 115.00 & 3000 & 5464 & 35.95 & 35.65 & 205.77 \\
885 Secondary Prism $\{11\bar{2}0\}$ & 11520 & 2048 & 146.72 & 124.48 & 3840 & 8176 & 71.87 & 31.66 & 161.55 \\
886 \botrule
887 \end{tabular}
888 \end{table}
889
890 \newpage
891 \begin{table}[h]
892 \centering
893 \caption{Structural and dynamic properties of the interfaces of
894 Ice-I$_\mathrm{h}$ with water.\label{tab:kappa}}
895 \begin{tabular}{r|cc|cc|cccc}
896 \toprule
897 \multirow{2}{*}{Interface} & \multicolumn{2}{c|}{Channel Size} &\multicolumn{2}{c|}{Droplet} & \multicolumn{4}{c}{Shearing\footnotemark[1]}\\
898 & Width (\AA) & Depth (\AA) & $\theta_{\infty}$ ($^\circ$) & $k_\mathrm{spread}$ (ns\textsuperscript{-1}) &
899 $\kappa_{x}$ & $\kappa_{y}$ & $d_\mathrm{struct}$ (\AA) & $d_\mathrm{dyn}$ (\AA) \\
900 \colrule
901 Basal $\{0001\}$ & 4.49 & 1.30 & $34.1(9)$ &$0.60(7)$
902 & $5.9(3)$ & $6.5(8)$ & $3.2(4)$ & $2(1)$ \\
903 Pyramidal $\{20\bar{2}1\}$ & 8.65 & 1.37 & $35(3)$ & $0.7(1)$ &
904 $5.8(4)$ & $6.1(5)$ & $3.2(2)$ & $2.5(3)$\\
905 Prismatic $\{10\bar{1}0\}$ & 6.35 & 2.25 & $45(3)$ & $0.75(9)$ &
906 $3.0(2)$ & $3.0(1)$ & $3.6(2)$ & $4(2)$ \\
907 Secondary Prism $\{11\bar{2}0\}$ & 6.35 & 2.25 & $43(2)$ & $0.69(3)$ &
908 $3.5(1)$ & $3.3(2)$ & $3.2(2)$ & $5(3)$ \\
909 \botrule
910 \end{tabular}
911 \begin{flushleft}
912 \footnotemark[1]\footnotesize{Liquid-solid friction coefficients ($\kappa_x$ and
913 $\kappa_y$) are expressed in 10\textsuperscript{-4} amu
914 \AA\textsuperscript{-2} fs\textsuperscript{-1}.} \\
915 \footnotemark[2]\footnotesize{Uncertainties in
916 the last digits are given in parentheses.}
917 \end{flushleft}
918 \end{table}
919
920 % Basal $\{0001\}$ & 4.49 & 1.30 & $34.1 \pm 0.9$ &$0.60 \pm 0.07$
921 % & $5.9 \pm 0.3$ & $6.5 \pm 0.8$ & $3.2 \pm 0.4$ & $2 \pm 1$ \\
922 % Pyramidal $\{2~0~\bar{2}~1\}$ & 8.65 & 1.37 & $35 \pm 3$ & $0.7 \pm 0.1$ &
923 % $5.8 \pm 0.4$ & $6.1 \pm 0.5$ & $3.2 \pm 0.2$ & $2.5 \pm 0.3$\\
924 % Prismatic $\{1~0~\bar{1}~0\}$ & 6.35 & 2.25 & $45 \pm 3$ & $0.75 \pm 0.09$ &
925 % $3.0 \pm 0.2$ & $3.0 \pm 0.1$ & $3.6 \pm 0.2$ & $4 \pm 2$ \\
926 % Secondary Prism $\{1~1~\bar{2}~0\}$ & 6.35 & 2.25 & $43 \pm 2$ & $0.69 \pm 0.03$ &
927 % $3.5 \pm 0.1$ & $3.3 \pm 0.2$ & $3.2 \pm 0.2$ & $5 \pm 3$ \\
928
929
930 \end{document}