ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/iceWater2/iceWater3.tex
Revision: 4257
Committed: Mon Dec 15 20:45:28 2014 UTC (9 years, 6 months ago) by gezelter
Content type: application/x-tex
File size: 47912 byte(s)
Log Message:
Latest

File Contents

# User Rev Content
1 gezelter 4243 %% PNAStwoS.tex
2     %% Sample file to use for PNAS articles prepared in LaTeX
3     %% For two column PNAS articles
4     %% Version1: Apr 15, 2008
5     %% Version2: Oct 04, 2013
6    
7     %% BASIC CLASS FILE
8     \documentclass{pnastwo}
9    
10     %% ADDITIONAL OPTIONAL STYLE FILES Font specification
11    
12     %\usepackage{PNASTWOF}
13     \usepackage[version=3]{mhchem}
14 gezelter 4245 \usepackage[round,numbers,sort&compress]{natbib}
15     \usepackage{fixltx2e}
16 gezelter 4247 \usepackage{booktabs}
17     \usepackage{multirow}
18 gezelter 4245 \bibpunct{(}{)}{,}{n}{,}{,}
19     \bibliographystyle{pnas2011}
20 gezelter 4243
21     %% OPTIONAL MACRO DEFINITIONS
22     \def\s{\sigma}
23     %%%%%%%%%%%%
24     %% For PNAS Only:
25     %\url{www.pnas.org/cgi/doi/10.1073/pnas.0709640104}
26     \copyrightyear{2014}
27     \issuedate{Issue Date}
28     \volume{Volume}
29     \issuenumber{Issue Number}
30     %\setcounter{page}{2687} %Set page number here if desired
31     %%%%%%%%%%%%
32    
33     \begin{document}
34    
35 gezelter 4254 \title{The different facets of ice have different hydrophilicities:
36     Friction at water / ice-I\textsubscript{h} interfaces}
37 gezelter 4243
38     \author{Patrick B. Louden\affil{1}{Department of Chemistry and Biochemistry, University of Notre Dame, Notre Dame,
39     IN 46556}
40     \and
41     J. Daniel Gezelter\affil{1}{}}
42    
43     \contributor{Submitted to Proceedings of the National Academy of Sciences
44     of the United States of America}
45    
46     %%%Newly updated.
47     %%% If significance statement need, then can use the below command otherwise just delete it.
48 gezelter 4257 \significancetext{Surface hydrophilicity is a measure of the
49     interaction strength between a solid surface and liquid water. Our
50     simulations show that the solid that is thought to be extremely
51     hydrophilic (ice) displays different behavior depending on which
52     crystal facet is presented to the liquid. This behavior is
53     potentially important in geophysics, in recognition of ice surfaces
54     by anti-freeze proteins, and in understanding how the friction
55     between ice and other solids may be mediated by a quasi-liquid layer
56     of water.}
57 gezelter 4243
58     \maketitle
59    
60     \begin{article}
61 gezelter 4245 \begin{abstract}
62 gezelter 4257 We present evidence that the prismatic and secondary prism facets
63     of ice-I$_\mathrm{h}$ crystals posess structural features that can
64     reduce the effective hydrophilicity of the ice/water
65     interface. The spreading dynamics of liquid water droplets on ice
66     facets exhibits long-time behavior that differs substantially for
67     the prismatic $\{1~0~\bar{1}~0\}$ and secondary prism
68     $\{1~1~\bar{2}~0\}$ facets when compared with the basal $\{0001\}$
69     and pyramidal $\{2~0~\bar{2}~1\}$ facets. We also present the
70     results of simulations of solid-liquid friction of the same four
71     crystal facets being drawn through liquid water. These simulation
72 gezelter 4245 techniques provide evidence that the two prismatic faces have an
73     effective surface area in contact with the liquid water of
74 gezelter 4257 approximately half of the total surface area of the crystal. The
75 gezelter 4245 ice / water interfacial widths for all four crystal facets are
76     similar (using both structural and dynamic measures), and were
77     found to be independent of the shear rate. Additionally,
78     decomposition of orientational time correlation functions show
79     position-dependence for the short- and longer-time decay
80     components close to the interface.
81 gezelter 4243 \end{abstract}
82    
83     \keywords{ice | water | interfaces | hydrophobicity}
84     \abbreviations{QLL, quasi liquid layer; MD, molecular dynamics; RNEMD,
85     reverse non-equilibrium molecular dynamics}
86    
87 gezelter 4245 \dropcap{S}urfaces can be characterized as hydrophobic or hydrophilic
88     based on the strength of the interactions with water. Hydrophobic
89     surfaces do not have strong enough interactions with water to overcome
90     the internal attraction between molecules in the liquid phase, and the
91     degree of hydrophilicity of a surface can be described by the extent a
92 gezelter 4254 droplet can spread out over the surface. The contact angle, $\theta$,
93     formed between the solid and the liquid depends on the free energies
94     of the three interfaces involved, and is given by Young's
95     equation~\cite{Young05},
96 gezelter 4245 \begin{equation}\label{young}
97     \cos\theta = (\gamma_{sv} - \gamma_{sl})/\gamma_{lv} .
98     \end{equation}
99     Here $\gamma_{sv}$, $\gamma_{sl}$, and $\gamma_{lv}$ are the free
100 gezelter 4254 energies of the solid/vapor, solid/liquid, and liquid/vapor interfaces,
101 gezelter 4245 respectively. Large contact angles, $\theta > 90^{\circ}$, correspond
102     to hydrophobic surfaces with low wettability, while small contact
103     angles, $\theta < 90^{\circ}$, correspond to hydrophilic surfaces.
104     Experimentally, measurements of the contact angle of sessile drops is
105     often used to quantify the extent of wetting on surfaces with
106     thermally selective wetting
107 gezelter 4254 characteristics~\cite{Tadanaga00,Liu04,Sun04}.
108 gezelter 4243
109 gezelter 4245 Nanometer-scale structural features of a solid surface can influence
110     the hydrophilicity to a surprising degree. Small changes in the
111     heights and widths of nano-pillars can change a surface from
112     superhydrophobic, $\theta \ge 150^{\circ}$, to hydrophilic, $\theta
113 plouden 4246 \sim 0^{\circ}$.\cite{Koishi09} This is often referred to as the
114 gezelter 4245 Cassie-Baxter to Wenzel transition. Nano-pillared surfaces with
115     electrically tunable Cassie-Baxter and Wenzel states have also been
116 gezelter 4254 observed~\cite{Herbertson06,Dhindsa06,Verplanck07,Ahuja08,Manukyan11}.
117 gezelter 4245 Luzar and coworkers have modeled these transitions on nano-patterned
118 gezelter 4254 surfaces~\cite{Daub07,Daub10,Daub11,Ritchie12}, and have found the
119 gezelter 4245 change in contact angle is due to the field-induced perturbation of
120 gezelter 4254 hydrogen bonding at the liquid/vapor interface~\cite{Daub07}.
121 gezelter 4245
122     One would expect the interfaces of ice to be highly hydrophilic (and
123     possibly the most hydrophilic of all solid surfaces). In this paper we
124     present evidence that some of the crystal facets of ice-I$_\mathrm{h}$
125 gezelter 4254 have structural features that can reduce the effective hydrophilicity.
126 gezelter 4245 Our evidence for this comes from molecular dynamics (MD) simulations
127     of the spreading dynamics of liquid droplets on these facets, as well
128     as reverse non-equilibrium molecular dynamics (RNEMD) simulations of
129     solid-liquid friction.
130    
131     Quiescent ice-I$_\mathrm{h}$/water interfaces have been studied
132     extensively using computer simulations. Haymet \textit{et al.}
133     characterized and measured the width of these interfaces for the
134     SPC~\cite{Karim90}, SPC/E~\cite{Gay02,Bryk02},
135     CF1~\cite{Hayward01,Hayward02} and TIP4P~\cite{Karim88} models, in
136     both neat water and with solvated
137     ions~\cite{Bryk04,Smith05,Wilson08,Wilson10}. Nada and Furukawa have
138     studied the width of basal/water and prismatic/water
139     interfaces~\cite{Nada95} as well as crystal restructuring at
140     temperatures approaching the melting point~\cite{Nada00}.
141    
142 gezelter 4243 The surface of ice exhibits a premelting layer, often called a
143 gezelter 4245 quasi-liquid layer (QLL), at temperatures near the melting point. MD
144     simulations of the facets of ice-I$_\mathrm{h}$ exposed to vacuum have
145     found QLL widths of approximately 10 \AA\ at 3 K below the melting
146 gezelter 4254 point~\cite{Conde08}. Similarly, Limmer and Chandler have used the mW
147 gezelter 4245 water model~\cite{Molinero09} and statistical field theory to estimate
148 gezelter 4254 QLL widths at similar temperatures to be about 3 nm~\cite{Limmer14}.
149 gezelter 4243
150 gezelter 4245 Recently, Sazaki and Furukawa have developed a technique using laser
151     confocal microscopy combined with differential interference contrast
152     microscopy that has sufficient spatial and temporal resolution to
153     visulaize and quantitatively analyze QLLs on ice crystals at
154 gezelter 4254 temperatures near melting~\cite{Sazaki10}. They have found the width of
155 gezelter 4245 the QLLs perpindicular to the surface at -2.2$^{o}$C to be 3-4 \AA\
156     wide. They have also seen the formation of two immiscible QLLs, which
157 gezelter 4254 displayed different dynamics on the crystal surface~\cite{Sazaki12}.
158 gezelter 4243
159 gezelter 4257 % There is now significant interest in the \textit{tribological}
160     % properties of ice/ice and ice/water interfaces in the geophysics
161     % community. Understanding the dynamics of solid-solid shearing that is
162     % mediated by a liquid layer~\cite{Cuffey99, Bell08} will aid in
163     % understanding the macroscopic motion of large ice
164     % masses~\cite{Casassa91, Sukhorukov13, Pritchard12, Lishman13}.
165 gezelter 4243
166     Using molecular dynamics simulations, Samadashvili has recently shown
167     that when two smooth ice slabs slide past one another, a stable
168 gezelter 4254 liquid-like layer develops between them~\cite{Samadashvili13}. In a
169 gezelter 4245 previous study, our RNEMD simulations of ice-I$_\mathrm{h}$ shearing
170     through liquid water have provided quantitative estimates of the
171 gezelter 4254 solid-liquid kinetic friction coefficients~\cite{Louden13}. These
172 gezelter 4245 displayed a factor of two difference between the basal and prismatic
173     facets. The friction was found to be independent of shear direction
174     relative to the surface orientation. We attributed facet-based
175     difference in liquid-solid friction to the 6.5 \AA\ corrugation of the
176     prismatic face which reduces the effective surface area of the ice
177     that is in direct contact with liquid water.
178 gezelter 4243
179 gezelter 4245 In the sections that follow, we outline the methodology used to
180     simulate droplet-spreading dynamics using standard MD and tribological
181     properties using RNEMD simulations. These simulation methods give
182     complementary results that point to the prismatic and secondary prism
183     facets having roughly half of their surface area in direct contact
184     with the liquid.
185 gezelter 4243
186 gezelter 4245 \section{Methodology}
187     \subsection{Construction of the Ice / Water Interfaces}
188     To construct the four interfacial ice/water systems, a proton-ordered,
189     zero-dipole crystal of ice-I$_\mathrm{h}$ with exposed strips of
190 gezelter 4247 H-atoms was created using Structure 6 of Hirsch and Ojam\"{a}e's set
191     of orthorhombic representations for ice-I$_{h}$~\cite{Hirsch04}. This
192     crystal structure was cleaved along the four different facets. The
193     exposed face was reoriented normal to the $z$-axis of the simulation
194     cell, and the structures were and extended to form large exposed
195     facets in rectangular box geometries. Liquid water boxes were created
196     with identical dimensions (in $x$ and $y$) as the ice, with a $z$
197     dimension of three times that of the ice block, and a density
198     corresponding to 1 g / cm$^3$. Each of the ice slabs and water boxes
199     were independently equilibrated at a pressure of 1 atm, and the
200     resulting systems were merged by carving out any liquid water
201     molecules within 3 \AA\ of any atoms in the ice slabs. Each of the
202     combined ice/water systems were then equilibrated at 225K, which is
203     the liquid-ice coexistence temperature for SPC/E
204 gezelter 4254 water~\cite{Bryk02}. Reference \citealp{Louden13} contains a more
205     detailed explanation of the construction of similar ice/water
206     interfaces. The resulting dimensions as well as the number of ice and
207     liquid water molecules contained in each of these systems are shown in
208     Table \ref{tab:method}.
209 gezelter 4243
210 gezelter 4247 The SPC/E water model~\cite{Berendsen87} has been extensively
211 plouden 4246 characterized over a wide range of liquid
212 gezelter 4254 conditions~\cite{Arbuckle02,Kuang12}, and its phase diagram has been
213     well studied~\cite{Baez95,Bryk04b,Sanz04b,Fennell:2005fk}, With longer
214 gezelter 4247 cutoff radii and careful treatment of electrostatics, SPC/E mostly
215     avoids metastable crystalline morphologies like
216     ice-\textit{i}~\cite{Fennell:2005fk} and ice-B~\cite{Baez95}. The
217 gezelter 4254 free energies and melting
218     points~\cite{Baez95,Arbuckle02,Gay02,Bryk02,Bryk04b,Sanz04b,Fennell:2005fk,Fernandez06,Abascal07,Vrbka07}
219 gezelter 4247 of various other crystalline polymorphs have also been calculated.
220     Haymet \textit{et al.} have studied quiescent Ice-I$_\mathrm{h}$/water
221     interfaces using the SPC/E water model, and have seen structural and
222     dynamic measurements of the interfacial width that agree well with
223     more expensive water models, although the coexistence temperature for
224     SPC/E is still well below the experimental melting point of real
225     water~\cite{Bryk02}. Given the extensive data and speed of this model,
226     it is a reasonable choice even though the temperatures required are
227     somewhat lower than real ice / water interfaces.
228 gezelter 4245
229 gezelter 4247 \subsection{Droplet Simulations}
230     Ice interfaces with a thickness of $\sim$~20~\AA\ were created as
231 gezelter 4245 described above, but were not solvated in a liquid box. The crystals
232     were then replicated along the $x$ and $y$ axes (parallel to the
233 gezelter 4247 surface) until a large surface ($>$ 126 nm\textsuperscript{2}) had
234     been created. The sizes and numbers of molecules in each of the
235     surfaces is given in Table \ref{tab:method}. Weak translational
236 gezelter 4254 restraining potentials with spring constants of 1.5~$\mathrm{kcal\
237     mol}^{-1}\mathrm{~\AA}^{-2}$ (prismatic and pyramidal facets) or
238     4.0~$\mathrm{kcal\ mol}^{-1}\mathrm{~\AA}^{-2}$ (basal facet) were
239     applied to the centers of mass of each molecule in order to prevent
240     surface melting, although the molecules were allowed to reorient
241     freely. A water doplet containing 2048 SPC/E molecules was created
242     separately. Droplets of this size can produce agreement with the Young
243     contact angle extrapolated to an infinite drop size~\cite{Daub10}. The
244     surfaces and droplet were independently equilibrated to 225 K, at
245     which time the droplet was placed 3-5~\AA\ above the surface. Five
246     statistically independent simulations were carried out for each facet,
247     and the droplet was placed at unique $x$ and $y$ locations for each of
248     these simulations. Each simulation was 5~ns in length and was
249     conducted in the microcanonical (NVE) ensemble. Representative
250     configurations for the droplet on the prismatic facet are shown in
251     figure \ref{fig:Droplet}.
252 gezelter 4243
253 gezelter 4247 \subsection{Shearing Simulations (Interfaces in Bulk Water)}
254    
255     To perform the shearing simulations, the velocity shearing and scaling
256     variant of reverse non-equilibrium molecular dynamics (VSS-RNEMD) was
257     employed \cite{Kuang12}. This method performs a series of simultaneous
258     non-equilibrium exchanges of linear momentum and kinetic energy
259     between two physically-separated regions of the simulation cell. The
260     system responds to this unphysical flux with velocity and temperature
261     gradients. When VSS-RNEMD is applied to bulk liquids, transport
262     properties like the thermal conductivity and the shear viscosity are
263     easily extracted assuming a linear response between the flux and the
264     gradient. At the interfaces between dissimilar materials, the same
265     method can be used to extract \textit{interfacial} transport
266     properties (e.g. the interfacial thermal conductance and the
267     hydrodynamic slip length).
268    
269     The kinetic energy flux (producing a thermal gradient) is necessary
270     when performing shearing simulations at the ice-water interface in
271     order to prevent the frictional heating due to the shear from melting
272 gezelter 4254 the crystal. Reference \citealp{Louden13} provides more details on the
273     VSS-RNEMD method as applied to ice-water interfaces. A representative
274     configuration of the solvated prismatic facet being sheared through
275     liquid water is shown in figure \ref{fig:Shearing}.
276 gezelter 4247
277 gezelter 4254 The exchanges between the two regions were carried out every 2 fs
278 gezelter 4257 (i.e. every time step). Although computationally expensive, this was
279     done to minimize the magnitude of each individual momentum exchange.
280     Because individual VSS-RNEMD exchanges conserve both total energy and
281     linear momentum, the method can be ``bolted-on'' to simulations in any
282     ensemble. The simulations of the pyramidal interface were performed
283     under the canonical (NVT) ensemble. When time correlation functions
284     were computed, the RNEMD simulations were done in the microcanonical
285     (NVE) ensemble. All simulations of the other interfaces were carried
286     out in the microcanonical ensemble.
287 gezelter 4247
288     \section{Results}
289     \subsection{Ice - Water Contact Angles}
290 plouden 4246
291     To determine the extent of wetting for each of the four crystal
292 gezelter 4247 facets, contact angles for liquid droplets on the ice surfaces were
293     computed using two methods. In the first method, the droplet is
294     assumed to form a spherical cap, and the contact angle is estimated
295     from the $z$-axis location of the droplet's center of mass
296     ($z_\mathrm{cm}$). This procedure was first described by Hautman and
297     Klein~\cite{Hautman91}, and was utilized by Hirvi and Pakkanen in
298     their investigation of water droplets on polyethylene and poly(vinyl
299     chloride) surfaces~\cite{Hirvi06}. For each stored configuration, the
300     contact angle, $\theta$, was found by inverting the expression for the
301     location of the droplet center of mass,
302 plouden 4246 \begin{equation}\label{contact_1}
303 gezelter 4247 \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} ,
304 plouden 4246 \end{equation}
305 gezelter 4247 where $R_{0}$ is the radius of the free water droplet.
306 plouden 4246
307 gezelter 4247 The second method for obtaining the contact angle was described by
308     Ruijter, Blake, and Coninck~\cite{Ruijter99}. This method uses a
309     cylindrical averaging of the droplet's density profile. A threshold
310     density of 0.5 g cm\textsuperscript{-3} is used to estimate the
311     location of the edge of the droplet. The $r$ and $z$-dependence of
312     the droplet's edge is then fit to a circle, and the contact angle is
313     computed from the intersection of the fit circle with the $z$-axis
314     location of the solid surface. Again, for each stored configuration,
315     the density profile in a set of annular shells was computed. Due to
316     large density fluctuations close to the ice, all shells located within
317     2 \AA\ of the ice surface were left out of the circular fits. The
318     height of the solid surface ($z_\mathrm{suface}$) along with the best
319 gezelter 4254 fitting origin ($z_\mathrm{droplet}$) and radius
320 gezelter 4247 ($r_\mathrm{droplet}$) of the droplet can then be used to compute the
321     contact angle,
322     \begin{equation}
323 gezelter 4254 \theta = 90 + \frac{180}{\pi} \sin^{-1}\left(\frac{z_\mathrm{droplet} -
324 gezelter 4247 z_\mathrm{surface}}{r_\mathrm{droplet}} \right).
325     \end{equation}
326     Both methods provided similar estimates of the dynamic contact angle,
327     although the first method is significantly less prone to noise, and
328     is the method used to report contact angles below.
329    
330     Because the initial droplet was placed above the surface, the initial
331 gezelter 4250 value of 180$^{\circ}$ decayed over time (See figure
332     \ref{fig:ContactAngle}). Each of these profiles were fit to a
333     biexponential decay, with a short-time contribution ($\tau_c$) that
334     describes the initial contact with the surface, a long time
335     contribution ($\tau_s$) that describes the spread of the droplet over
336     the surface, and a constant ($\theta_\infty$) to capture the
337     infinite-time estimate of the equilibrium contact angle,
338 gezelter 4247 \begin{equation}
339 gezelter 4250 \theta(t) = \theta_\infty + (180-\theta_\infty) \left[ a e^{-t/\tau_c} +
340     (1-a) e^{-t/\tau_s} \right]
341 gezelter 4247 \end{equation}
342 gezelter 4250 We have found that the rate for water droplet spreading across all
343     four crystal facets, $k_\mathrm{spread} = 1/\tau_s \approx$ 0.7
344     ns$^{-1}$. However, the basal and pyramidal facets produced estimated
345     equilibrium contact angles, $\theta_\infty \approx$ 35$^{o}$, while
346 gezelter 4247 prismatic and secondary prismatic had values for $\theta_\infty$ near
347 gezelter 4250 43$^{o}$ as seen in Table \ref{tab:kappa}.
348 gezelter 4247
349 gezelter 4254 These results indicate that the basal and pyramidal facets are more
350 gezelter 4257 hydrophilic by traditional measures than the prismatic and secondary
351     prism facets, and surprisingly, that the differential hydrophilicities
352     of the crystal facets is not reflected in the spreading rate of the
353     droplet.
354 gezelter 4247
355 plouden 4246 % This is in good agreement with our calculations of friction
356     % coefficients, in which the basal
357     % and pyramidal had a higher coefficient of kinetic friction than the
358     % prismatic and secondary prismatic. Due to this, we beleive that the
359     % differences in friction coefficients can be attributed to the varying
360     % hydrophilicities of the facets.
361    
362 gezelter 4257 \subsection{Solid-liquid friction of the interfaces}
363 gezelter 4254 In a bulk fluid, the shear viscosity, $\eta$, can be determined
364     assuming a linear response to a shear stress,
365 plouden 4246 \begin{equation}\label{Shenyu-11}
366 gezelter 4254 j_{z}(p_{x}) = \eta \frac{\partial v_{x}}{\partial z}.
367 plouden 4246 \end{equation}
368 gezelter 4254 Here $j_{z}(p_{x})$ is the flux (in $x$-momentum) that is transferred
369     in the $z$ direction (i.e. the shear stress). The RNEMD simulations
370     impose an artificial momentum flux between two regions of the
371     simulation, and the velocity gradient is the fluid's response. This
372     technique has now been applied quite widely to determine the
373     viscosities of a number of bulk fluids~\cite{}.
374    
375     At the interface between two phases (e.g. liquid / solid) the same
376     momentum flux creates a velocity difference between the two materials,
377     and this can be used to define an interfacial friction coefficient
378     ($\kappa$),
379     \begin{equation}\label{Shenyu-13}
380     j_{z}(p_{x}) = \kappa \left[ v_{x}(liquid) - v_{x}(solid) \right]
381 plouden 4246 \end{equation}
382 gezelter 4254 where $v_{x}(liquid)$ and $v_{x}(solid)$ are the velocities measured
383     directly adjacent to the interface.
384    
385     The simulations described here contain significant quantities of both
386     liquid and solid phases, and the momentum flux must traverse a region
387     of the liquid that is simultaneously under a thermal gradient. Since
388     the liquid has a temperature-dependent shear viscosity, $\eta(T)$,
389     estimates of the solid-liquid friction coefficient can be obtained if
390     one knows the viscosity of the liquid at the interface (i.e. at the
391     melting temperature, $T_m$),
392 plouden 4246 \begin{equation}\label{kappa-2}
393 gezelter 4254 \kappa = \frac{\eta(T_{m})}{\left[v_{x}(fluid)-v_{x}(solid)\right]}\left(\frac{\partial v_{x}}{\partial z}\right).
394 plouden 4246 \end{equation}
395 gezelter 4254 For SPC/E, the melting temperature of Ice-I$_\mathrm{h}$ is estimated
396     to be 225~K~\cite{Bryk02}. To obtain the value of
397     $\eta(225\mathrm{~K})$ for the SPC/E model, a $31.09 \times 29.38
398     \times 124.39$ \AA\ box with 3744 water molecules in a disordered
399     configuration was equilibrated to 225~K, and five
400     statistically-independent shearing simulations were performed (with
401 gezelter 4257 imposed fluxes that spanned a range of $3 \rightarrow 13
402     \mathrm{~m~s}^{-1}$ ). Each simulation was conducted in the
403     microcanonical ensemble with total simulation times of 5 ns. The
404     VSS-RNEMD exchanges were carried out every 2 fs. We estimate
405     $\eta(225\mathrm{~K})$ to be 0.0148 $\pm$ 0.0007 Pa s for SPC/E,
406     roughly ten times larger than the shear viscosity previously computed
407     at 280~K~\cite{Kuang12}.
408 plouden 4246
409 gezelter 4257 The interfacial friction coefficient can equivalently be expressed as
410     the ratio of the viscosity of the fluid to the hydrodynamic slip
411     length, $\kappa = \eta / \delta$. The slip length is an indication of
412     strength of the interactions between the solid and liquid phases,
413     although the connection between slip length and surface hydrophobicity
414     is not yet clear. In some simulations, the slip length has been found
415     to have a link to the effective surface
416     hydrophobicity~\cite{Sendner:2009uq}, although Ho \textit{et al.} have
417     found that liquid water can also slip on hydrophilic
418     surfaces~\cite{Ho:2011zr}. Experimental evidence for a direct tie
419     between slip length and hydrophobicity is also not
420 gezelter 4254 definitive. Total-internal reflection particle image velocimetry
421     (TIR-PIV) studies have suggested that there is a link between slip
422     length and effective
423     hydrophobicity~\cite{Lasne:2008vn,Bouzigues:2008ys}. However, recent
424     surface sensitive cross-correlation spectroscopy (TIR-FCCS)
425     measurements have seen similar slip behavior for both hydrophobic and
426     hydrophilic surfaces~\cite{Schaeffel:2013kx}.
427 plouden 4246
428 gezelter 4254 In each of the systems studied here, the interfacial temperature was
429     kept fixed to 225K, which ensured the viscosity of the fluid at the
430     interace was identical. Thus, any significant variation in $\kappa$
431     between the systems is a direct indicator of the slip length and the
432     effective interaction strength between the solid and liquid phases.
433 gezelter 4243
434 gezelter 4254 The calculated $\kappa$ values found for the four crystal facets of
435     Ice-I$_\mathrm{h}$ are shown in Table \ref{tab:kappa}. The basal and
436     pyramidal facets were found to have similar values of $\kappa \approx
437     6$ ($\times 10^{-4} \mathrm{amu~\AA}^{-2}\mathrm{fs}^{-1}$), while the
438     prismatic and secondary prism facets exhibited $\kappa \approx 3$
439     ($\times 10^{-4} \mathrm{amu~\AA}^{-2}\mathrm{fs}^{-1}$). These
440     results are also essentially independent of shearing direction
441     relative to features on the surface of the facets. The friction
442     coefficients indicate that the basal and pyramidal facets have
443     significantly stronger interactions with liquid water than either of
444     the two prismatic facets. This is in agreement with the contact angle
445     results above - both of the high-friction facets exhbited smaller
446     contact angles, suggesting that the solid-liquid friction is
447     correlated with the hydrophilicity of these facets.
448 gezelter 4243
449 gezelter 4254 \subsection{Structural measures of interfacial width under shear}
450     One of the open questions about ice/water interfaces is whether the
451     thickness of the 'slush-like' quasi-liquid layer (QLL) depends on the
452     facet of ice presented to the water. In the QLL region, the water
453     molecules are ordered differently than in either the solid or liquid
454     phases, and also exhibit distinct dynamical behavior. The width of
455     this quasi-liquid layer has been estimated by finding the distance
456     over which structural order parameters or dynamic properties change
457     from their bulk liquid values to those of the solid ice. The
458     properties used to find interfacial widths have included the local
459     density, the diffusion constant, and the translational and
460     orientational order
461     parameters~\cite{Karim88,Karim90,Hayward01,Hayward02,Bryk02,Gay02,Louden13}.
462    
463     The VSS-RNEMD simulations impose thermal and velocity gradients.
464     These gradients perturb the momenta of the water molecules, so
465     parameters that depend on translational motion are often measuring the
466     momentum exchange, and not physical properties of the interface. As a
467     structural measure of the interface, we have used the local
468     tetrahedral order parameter to estimate the width of the interface.
469     This quantity was originally described by Errington and
470     Debenedetti~\cite{Errington01} and has been used in bulk simulations
471     by Kumar \textit{et al.}~\cite{Kumar09}. It has previously been used
472     in ice/water interfaces by by Bryk and Haymet~\cite{Bryk04b}.
473    
474     To determine the structural widths of the interfaces under shear, each
475     of the systems was divided into 100 bins along the $z$-dimension, and
476     the local tetrahedral order parameter (Eq. 5 in Reference
477     \citealp{Louden13}) was time-averaged in each bin for the duration of
478     the shearing simulation. The spatial dependence of this order
479     parameter, $q(z)$, is the tetrahedrality profile of the interface. A
480     representative profile for the pyramidal facet is shown in circles in
481     panel $a$ of figure \ref{fig:pyrComic}. The $q(z)$ function has a
482     range of $(0,1)$, where a value of unity indicates a perfectly
483     tetrahedral environment. The $q(z)$ for the bulk liquid was found to
484     be $\approx~0.77$, while values of $\approx~0.92$ were more common in
485     the ice. The tetrahedrality profiles were fit using a hyperbolic
486     tangent function (see Eq. 6 in Reference \citealp{Louden13}) designed
487     to smoothly fit the bulk to ice transition while accounting for the
488     weak thermal gradient. In panels $b$ and $c$, the resulting thermal
489     and velocity gradients from an imposed kinetic energy and momentum
490     fluxes can be seen. The vertical dotted lines traversing all three
491     panels indicate the midpoints of the interface as determined by the
492     tetrahedrality profiles.
493 gezelter 4243
494 gezelter 4254 We find the interfacial width to be $3.2 \pm 0.2$ \AA\ (pyramidal) and
495     $3.2 \pm 0.2$ \AA\ (secondary prism) for the control systems with no
496     applied momentum flux. This is similar to our previous results for the
497     interfacial widths of the quiescent basal ($3.2 \pm 0.4$ \AA) and
498     prismatic systems ($3.6 \pm 0.2$ \AA).
499 gezelter 4243
500 gezelter 4254 Over the range of shear rates investigated, $0.4 \rightarrow
501     6.0~\mathrm{~m~s}^{-1}$ for the pyramidal system and $0.6 \rightarrow
502     5.2~\mathrm{~m~s}^{-1}$ for the secondary prism, we found no
503     significant change in the interfacial width. The mean interfacial
504     widths are collected in table \ref{tab:kappa}. This follows our
505     previous findings of the basal and prismatic systems, in which the
506     interfacial widths of the basal and prismatic facets were also found
507     to be insensitive to the shear rate~\cite{Louden13}.
508 gezelter 4243
509 gezelter 4254 The similarity of these interfacial width estimates indicate that the
510     particular facet of the exposed ice crystal has little to no effect on
511     how far into the bulk the ice-like structural ordering persists. Also,
512     it appears that for the shearing rates imposed in this study, the
513     interfacial width is not structurally modified by the movement of
514     water over the ice.
515 gezelter 4243
516 gezelter 4254 \subsection{Dynamic measures of interfacial width under shear}
517 gezelter 4257 The spatially-resolved orientational time correlation function,
518 gezelter 4243 \begin{equation}\label{C(t)1}
519 gezelter 4257 C_{2}(z,t)=\langle P_{2}(\mathbf{u}_i(0)\cdot \mathbf{u}_i(t))
520     \delta(z_i(0) - z) \rangle,
521 gezelter 4243 \end{equation}
522 gezelter 4257 provides local information about the decorrelation of molecular
523     orientations in time. Here, $P_{2}$ is the second-order Legendre
524     polynomial, and $\mathbf{u}_i$ is the molecular vector that bisects
525     the HOH angle of molecule $i$. The angle brackets indicate an average
526     over all the water molecules, and the delta function restricts the
527     average to specific regions. In the crystal, decay of $C_2(z,t)$ is
528     incomplete, while liquid water correlation times are typically
529     measured in ps. Observing the spatial-transition between the decay
530     regimes can define a dynamic measure of the interfacial width.
531 gezelter 4243
532 gezelter 4257 Each of the systems was divided into bins along the $z$-dimension
533     ($\approx$ 3 \AA\ wide) and $C_2(z,t)$ was computed using only those
534     molecules that were in the bin at the initial time. The
535     time-dependence was fit to a triexponential decay, with three time
536     constants: $\tau_{short}$, measuring the librational motion of the
537     water molecules, $\tau_{middle}$, measuring the timescale for breaking
538     and making of hydrogen bonds, and $\tau_{long}$, corresponding to the
539     translational motion of the water molecules. An additional constant
540     was introduced in the fits to describe molecules in the crystal which
541     do not experience long-time orientational decay.
542 gezelter 4243
543 gezelter 4257 In Figures S5-S8 in the supporting information, the $z$-coordinate
544     profiles for the three decay constants, $\tau_{short}$,
545     $\tau_{middle}$, and $\tau_{long}$ for the different interfaces are
546     shown. Figures S5 \& S6 are new results, and Figures S7 \& S8 are
547     updated plots from Ref \citealp{Louden13}. In the liquid regions of
548     all four interfaces, we observe $\tau_{middle}$ and $\tau_{long}$ to
549     have approximately consistent values of $3-6$ ps and $30-40$ ps,
550     respectively. Both of these times increase in value approaching the
551     interface. Approaching the interface, we also observe that
552     $\tau_{short}$ decreases from its liquid-state value of $72-76$ fs.
553     The approximate values for the decay constants and the trends
554     approaching the interface match those reported previously for the
555     basal and prismatic interfaces.
556 gezelter 4243
557 gezelter 4257 We have estimated the dynamic interfacial width $d_\mathrm{dyn}$ by
558     fitting the profiles of all the three orientational time constants
559     with an exponential decay to the bulk-liquid behavior,
560 gezelter 4243 \begin{equation}\label{tauFit}
561 gezelter 4257 \tau(z)\approx\tau_{liquid}+(\tau_{wall}-\tau_{liquid})e^{-(z-z_{wall})/d_\mathrm{dyn}}
562 gezelter 4243 \end{equation}
563 gezelter 4257 where $\tau_{liquid}$ and $\tau_{wall}$ are the liquid and projected
564     wall values of the decay constants, $z_{wall}$ is the location of the
565     interface, as measured by the structural order parameter. These
566     values are shown in table \ref{tab:kappa}. Because the bins must be
567     quite wide to obtain reasonable profiles of $C_2(z,t)$, the error
568     estimates for the dynamic widths of the interface are significantly
569     larger than for the structural widths. However, all four interfaces
570     exhibit dynamic widths that are significantly below 1~nm, and are in
571     reasonable agreement with the structural width above.
572 gezelter 4243
573 gezelter 4257 \section{Conclusions}
574     In this work, we used MD simulations to measure the advancing contact
575     angles of water droplets on the basal, prismatic, pyramidal, and
576     secondary prism facets of Ice-I$_\mathrm{h}$. Although there was no
577     significant change in the \textit{rate} at which the droplets spread
578     over the surface, the long-time behavior indicates that we should
579     expect to see larger equilibrium contact angles for the two prismatic
580     facets.
581 gezelter 4243
582 gezelter 4257 We have also used RNEMD simulations of water interfaces with the same
583     four crystal facets to compute solid-liquid friction coefficients. We
584     have observed coefficients of friction that differ by a factor of two
585     between the two prismatic facets and the basal and pyramidal facets.
586     Because the solid-liquid friction coefficient is directly tied to the
587     hydrodynamic slip length, this suggests that there are significant
588     differences in the overall interaction strengths between these facets
589     and the liquid layers immediately in contact with them.
590 gezelter 4243
591 gezelter 4257 The agreement between these two measures have lead us to conclude that
592     the two prismatic facets have a lower hydrophilicity than either the
593     basal or pyramidal facets. One possible explanation of this behavior
594     is that the face presented by both prismatic facets consists of deep,
595     narrow channels (i.e. stripes of adjacent rows of pairs of
596     hydrodgen-bound water molecules). At the surfaces of these facets,
597     the channels are 6.35 \AA\ wide and the sub-surface ice layer is 2.25
598     \AA\ below (and therefore blocked from hydrogen bonding with the
599     liquid). This means that only 1/2 of the surface molecules can form
600     hydrogen bonds with liquid-phase molecules.
601 gezelter 4243
602 gezelter 4257 In the basal plane, the surface features are narrower (4.49 \AA) and
603     shallower (1.3 \AA), while the pyramidal face has much wider channels
604     (8.65 \AA) which are also quite shallow (1.37 \AA). These features
605     allow liquid phase molecules to form hydrogen bonds with all of the
606     surface molecules in the basal and pyramidal facets. This means that
607     for similar surface areas, the two prismatic facets have an effective
608     hydrogen bonding surface area of half of the basal and pyramidal
609     facets. The reduction in the effective surface area would explain
610     much of the behavior observed in our simulations.
611 gezelter 4243
612     \begin{acknowledgments}
613     Support for this project was provided by the National
614     Science Foundation under grant CHE-1362211. Computational time was
615     provided by the Center for Research Computing (CRC) at the
616     University of Notre Dame.
617     \end{acknowledgments}
618    
619     \bibliography{iceWater}
620     % *****************************************
621     % There is significant interest in the properties of ice/ice and ice/water
622     % interfaces in the geophysics community. Most commonly, the results of shearing
623     % two ice blocks past one
624     % another\cite{Casassa91, Sukhorukov13, Pritchard12, Lishman13} or the shearing
625     % of ice through water\cite{Cuffey99, Bell08}. Using molecular dynamics
626     % simulations, Samadashvili has recently shown that when two smooth ice slabs
627     % slide past one another, a stable liquid-like layer develops between
628     % them\cite{Samadashvili13}. To fundamentally understand these processes, a
629     % molecular understanding of the ice/water interfaces is needed.
630    
631     % Investigation of the ice/water interface is also crucial in understanding
632     % processes such as nucleation, crystal
633     % growth,\cite{Han92, Granasy95, Vanfleet95} and crystal
634     % melting\cite{Weber83, Han92, Sakai96, Sakai96B}. Insight gained to these
635     % properties can also be applied to biological systems of interest, such as
636     % the behavior of the antifreeze protein found in winter
637     % flounder,\cite{Wierzbicki07,Chapsky97} and certain terrestial
638     % arthropods.\cite{Duman:2001qy,Meister29012013} Elucidating the properties which
639     % give rise to these processes through experimental techniques can be expensive,
640     % complicated, and sometimes infeasible. However, through the use of molecular
641     % dynamics simulations much of the problems of investigating these properties
642     % are alleviated.
643    
644     % Understanding ice/water interfaces inherently begins with the isolated
645     % systems. There has been extensive work parameterizing models for liquid water,
646     % such as the SPC\cite{Berendsen81}, SPC/E\cite{Berendsen87},
647     % TIP4P\cite{Jorgensen85}, TIP4P/2005\cite{Abascal05},
648     % ($\dots$), and more recently, models for simulating
649     % the solid phases of water, such as the TIP4P/Ice\cite{Abascal05b} model. The
650     % melting point of various crystal structures of ice have been calculated for
651     % many of these models
652     % (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}),
653     % and the partial or complete phase diagram for the model has been determined
654     % (SPC/E\cite{Baez95,Bryk04b,Sanz04b}, TIP4P\cite{Sanz04,Sanz04b,Koyama04}, TIP5P\cite{Sanz04,Koyama04}).
655     % Knowing the behavior and melting point for these models has enabled an initial
656     % investigation of ice/water interfaces.
657    
658     % The Ice-I$_\mathrm{h}$/water quiescent interface has been extensively studied
659     % over the past 30 years by theory and experiment. Haymet \emph{et al.} have
660     % done significant work characterizing and quantifying the width of these
661     % interfaces for the SPC,\cite{Karim90} SPC/E,\cite{Gay02,Bryk02},
662     % CF1,\cite{Hayward01,Hayward02} and TIP4P\cite{Karim88} models for water. In
663     % recent years, Haymet has focused on investigating the effects cations and
664     % anions have on crystal nucleaion and
665     % melting.\cite{Bryk04,Smith05,Wilson08,Wilson10} Nada and Furukawa have studied
666     % the the basal- and prismatic-water interface width\cite{Nada95}, crystal
667     % surface restructuring at temperatures approaching the melting
668     % point\cite{Nada00}, and the mechanism of ice growth inhibition by antifreeze
669     % proteins\cite{Nada08,Nada11,Nada12}. Nada has developed a six-site water model
670     % for ice/water interfaces near the melting point\cite{Nada03}, and studied the
671     % dependence of crystal growth shape on applied pressure\cite{Nada11b}. Using
672     % this model, Nada and Furukawa have established differential
673     % growth rates for the basal, prismatic, and secondary prismatic facets of
674     % Ice-I$_\mathrm{h}$ and found their origins due to a reordering of the hydrogen
675     % bond network in water near the interface\cite{Nada05}. While the work
676     % described so far has mainly focused on bulk water on ice, there is significant
677     % interest in thin films of water on ice surfaces as well.
678    
679     % It is well known that the surface of ice exhibits a premelting layer at
680     % temperatures near the melting point, often called a quasi-liquid layer (QLL).
681     % Molecular dynamics simulations of the facets of ice-I$_\mathrm{h}$ exposed
682     % to vacuum performed by Conde, Vega and Patrykiejew have found QLL widths of
683     % approximately 10 \AA\ at 3 K below the melting point\cite{Conde08}.
684     % Similarly, Limmer and Chandler have used course grain simulations and
685     % statistical field theory to estimated QLL widths at the same temperature to
686     % be about 3 nm\cite{Limmer14}.
687     % Recently, Sazaki and Furukawa have developed an experimental technique with
688     % sufficient spatial and temporal resolution to visulaize and quantitatively
689     % analyze QLLs on ice crystals at temperatures near melting\cite{Sazaki10}. They
690     % have found the width of the QLLs perpindicular to the surface at -2.2$^{o}$C
691     % to be $\mathcal{O}$(\AA). They have also seen the formation of two immiscible
692     % QLLs, which displayed different stabilities and dynamics on the crystal
693     % surface\cite{Sazaki12}. Knowledge of the hydrophilicities of each
694     % of the crystal facets would help further our understanding of the properties
695     % and dynamics of the QLLs.
696    
697     % Presented here is the follow up to our previous paper\cite{Louden13}, in which
698     % the basal and prismatic facets of an ice-I$_\mathrm{h}$/water interface were
699     % investigated where the ice was sheared relative to the liquid. By using a
700     % recently developed velocity shearing and scaling approach to reverse
701     % non-equilibrium molecular dynamics (VSS-RNEMD), simultaneous temperature and
702     % velocity gradients can be applied to the system, which allows for measurment
703     % of friction and thermal transport properties while maintaining a stable
704     % interfacial temperature\cite{Kuang12}. Structural analysis and dynamic
705     % correlation functions were used to probe the interfacial response to a shear,
706     % and the resulting solid/liquid kinetic friction coefficients were reported.
707     % In this paper we present the same analysis for the pyramidal and secondary
708     % prismatic facets, and show that the differential interfacial friction
709     % coefficients for the four facets of ice-I$_\mathrm{h}$ are determined by their
710     % relative hydrophilicity by means of dynamics water contact angle
711     % simulations.
712    
713     % The local tetrahedral order parameter, $q(z)$, is given by
714     % \begin{equation}
715     % q(z) = \int_0^L \sum_{k=1}^{N} \Bigg(1 -\frac{3}{8}\sum_{i=1}^{3}
716     % \sum_{j=i+1}^{4} \bigg(\cos\psi_{ikj}+\frac{1}{3}\bigg)^2\Bigg)
717     % \delta(z_{k}-z)\mathrm{d}z \Bigg/ N_z
718     % \label{eq:qz}
719     % \end{equation}
720     % where $\psi_{ikj}$ is the angle formed between the oxygen sites of molecules
721     % $i$,$k$, and $j$, where the centeral oxygen is located within molecule $k$ and
722     % molecules $i$ and $j$ are two of the closest four water molecules
723     % around molecule $k$. All four closest neighbors of molecule $k$ are also
724     % required to reside within the first peak of the pair distribution function
725     % for molecule $k$ (typically $<$ 3.41 \AA\ for water).
726     % $N_z = \int\delta(z_k - z) \mathrm{d}z$ is a normalization factor to account
727     % for the varying population of molecules within each finite-width bin.
728    
729    
730     % The hydrophobicity or hydrophilicity of a surface can be described by the
731     % extent a droplet of water wets the surface. The contact angle formed between
732     % the solid and the liquid, $\theta$, which relates the free energies of the
733     % three interfaces involved, is given by Young's equation.
734     % \begin{equation}\label{young}
735     % \cos\theta = (\gamma_{sv} - \gamma_{sl})/\gamma_{lv}
736     % \end{equation}
737     % Here $\gamma_{sv}$, $\gamma_{sl}$, and $\gamma_{lv}$ are the free energies
738     % of the solid/vapor, solid/liquid, and liquid/vapor interfaces respectively.
739     % Large contact angles ($\theta$ $\gg$ 90\textsuperscript{o}) correspond to low
740     % wettability and hydrophobic surfaces, while small contact angles
741     % ($\theta$ $\ll$ 90\textsuperscript{o}) correspond to high wettability and
742     % hydrophilic surfaces. Experimentally, measurements of the contact angle
743     % of sessile drops has been used to quantify the extent of wetting on surfaces
744     % with thermally selective wetting charactaristics\cite{Tadanaga00,Liu04,Sun04},
745     % as well as nano-pillared surfaces with electrically tunable Cassie-Baxter and
746     % Wenzel states\cite{Herbertson06,Dhindsa06,Verplanck07,Ahuja08,Manukyan11}.
747     % Luzar and coworkers have done significant work modeling these transitions on
748     % nano-patterned surfaces\cite{Daub07,Daub10,Daub11,Ritchie12}, and have found
749     % the change in contact angle to be due to the external field perturbing the
750     % hydrogen bonding of the liquid/vapor interface\cite{Daub07}.
751    
752 gezelter 4257 % SI stuff:
753 gezelter 4243
754 gezelter 4257 % Correlation functions:
755     % To compute eq. \eqref{C(t)1}, each 0.5 ns simulation was
756     % followed by an additional 200 ps NVE simulation during which the
757     % position and orientations of each molecule were recorded every 0.1 ps.
758 gezelter 4243
759 gezelter 4257
760    
761    
762 gezelter 4243 \end{article}
763    
764     \begin{figure}
765 gezelter 4247 \includegraphics[width=\linewidth]{Droplet}
766     \caption{\label{fig:Droplet} Computational model of a droplet of
767     liquid water spreading over the prismatic $\{1~0~\bar{1}~0\}$ facet
768 gezelter 4250 of ice, before (left) and 2.6 ns after (right) being introduced to the
769 gezelter 4247 surface. The contact angle ($\theta$) shrinks as the simulation
770     proceeds, and the long-time behavior of this angle is used to
771     estimate the hydrophilicity of the facet.}
772     \end{figure}
773    
774     \begin{figure}
775 gezelter 4250 \includegraphics[width=2in]{Shearing}
776 gezelter 4247 \caption{\label{fig:Shearing} Computational model of a slab of ice
777 gezelter 4250 being sheared through liquid water. In this figure, the ice is
778     presenting two copies of the prismatic $\{1~0~\bar{1}~0\}$ facet
779     towards the liquid phase. The RNEMD simulation exchanges both
780     linear momentum (indicated with arrows) and kinetic energy between
781     the central box and the box that spans the cell boundary. The
782     system responds with weak thermal gradient and a velocity profile
783     that shears the ice relative to the surrounding liquid.}
784 gezelter 4247 \end{figure}
785    
786     \begin{figure}
787 gezelter 4250 \includegraphics[width=\linewidth]{ContactAngle}
788     \caption{\label{fig:ContactAngle} The dynamic contact angle of a
789     droplet after approaching each of the four ice facets. The decay to
790     an equilibrium contact angle displays similar dynamics. Although
791     all the surfaces are hydrophilic, the long-time behavior stabilizes
792     to significantly flatter droplets for the basal and pyramidal
793     facets. This suggests a difference in hydrophilicity for these
794     facets compared with the two prismatic facets.}
795 gezelter 4243 \end{figure}
796    
797 gezelter 4257 % \begin{figure}
798     % \includegraphics[width=\linewidth]{Pyr_comic_strip}
799     % \caption{\label{fig:pyrComic} Properties of the pyramidal interface
800     % being sheared through water at 3.8 ms\textsuperscript{-1}. Lower
801     % panel: the local tetrahedral order parameter, $q(z)$, (circles) and
802     % the hyperbolic tangent fit (turquoise line). Middle panel: the
803     % imposed thermal gradient required to maintain a fixed interfacial
804     % temperature of 225 K. Upper panel: the transverse velocity gradient
805     % that develops in response to an imposed momentum flux. The vertical
806     % dotted lines indicate the locations of the midpoints of the two
807     % interfaces.}
808     % \end{figure}
809 gezelter 4243
810 gezelter 4250 % \begin{figure}
811     % \includegraphics[width=\linewidth]{SP_comic_strip}
812     % \caption{\label{fig:spComic} The secondary prismatic interface with a shear
813     % rate of 3.5 \
814     % ms\textsuperscript{-1}. Panel descriptions match those in figure \ref{fig:pyrComic}.}
815     % \end{figure}
816    
817 gezelter 4257 % \begin{figure}
818     % \includegraphics[width=\linewidth]{Pyr-orient}
819     % \caption{\label{fig:PyrOrient} The three decay constants of the
820     % orientational time correlation function, $C_2(z,t)$, for water as a
821     % function of distance from the center of the ice slab. The vertical
822     % dashed line indicates the edge of the pyramidal ice slab determined
823     % by the local order tetrahedral parameter. The control (circles) and
824     % sheared (squares) simulations were fit using shifted-exponential
825     % decay (see Eq. 9 in Ref. \citealp{Louden13}).}
826     % \end{figure}
827 gezelter 4243
828 gezelter 4250 % \begin{figure}
829     % \includegraphics[width=\linewidth]{SP-orient-less}
830 gezelter 4257 % \caption{\label{fig:SPorient} Decay constants for $C_2(z,t)$ at the secondary
831 gezelter 4250 % prismatic face. Panel descriptions match those in \ref{fig:PyrOrient}.}
832     % \end{figure}
833 gezelter 4243
834    
835     \begin{table}[h]
836     \centering
837 gezelter 4247 \caption{Sizes of the droplet and shearing simulations. Cell
838     dimensions are measured in \AA. \label{tab:method}}
839     \begin{tabular}{r|cccc|ccccc}
840     \toprule
841     \multirow{2}{*}{Interface} & \multicolumn{4}{c|}{Droplet} & \multicolumn{5}{c}{Shearing} \\
842     & $N_\mathrm{ice}$ & $N_\mathrm{droplet}$ & $L_x$ & $L_y$ & $N_\mathrm{ice}$ & $N_\mathrm{liquid}$ & $L_x$ & $L_y$ & $L_z$ \\
843     \midrule
844     Basal $\{0001\}$ & 12960 & 2048 & 134.70 & 140.04 & 900 & 1846 & 23.87 & 35.83 & 98.64 \\
845     Pyramidal $\{2~0~\bar{2}~1\}$ & 11136 & 2048 & 143.75 & 121.41 & 1216 & 2203 & 37.47 & 29.50 & 93.02 \\
846     Prismatic $\{1~0~\bar{1}~0\}$ & 9900 & 2048 & 110.04 & 115.00 & 3000 & 5464 & 35.95 & 35.65 & 205.77 \\
847     Secondary Prism $\{1~1~\bar{2}~0\}$ & 11520 & 2048 & 146.72 & 124.48 & 3840 & 8176 & 71.87 & 31.66 & 161.55 \\
848     \bottomrule
849 gezelter 4243 \end{tabular}
850     \end{table}
851    
852    
853     \begin{table}[h]
854     \centering
855 gezelter 4247 \caption{Structural and dynamic properties of the interfaces of Ice-I$_\mathrm{h}$
856     with water. Liquid-solid friction coefficients ($\kappa_x$ and $\kappa_y$) are expressed in 10\textsuperscript{-4} amu
857     \AA\textsuperscript{-2} fs\textsuperscript{-1}. \label{tab:kappa}}
858     \begin{tabular}{r|cc|cccc}
859     \toprule
860     \multirow{2}{*}{Interface} & \multicolumn{2}{c|}{Droplet} & \multicolumn{4}{c}{Shearing}\\
861     & $\theta_{\infty}$ ($^\circ$) & $k_\mathrm{spread}$ (ns\textsuperscript{-1}) &
862 gezelter 4257 $\kappa_{x}$ & $\kappa_{y}$ & $d_\mathrm{struct}$ (\AA) & $d_\mathrm{dyn}$ (\AA) \\
863 gezelter 4247 \midrule
864 plouden 4251 Basal $\{0001\}$ & $34.1 \pm 0.9$ &$0.60 \pm 0.07$
865 plouden 4253 & $5.9 \pm 0.3$ & $6.5 \pm 0.8$ & $3.2 \pm 0.4$ & $2 \pm 1$ \\
866 gezelter 4250 Pyramidal $\{2~0~\bar{2}~1\}$ & $35 \pm 3$ & $0.7 \pm 0.1$ &
867 plouden 4253 $5.8 \pm 0.4$ & $6.1 \pm 0.5$ & $3.2 \pm 0.2$ & $2.5 \pm 0.3$\\
868 plouden 4251 Prismatic $\{1~0~\bar{1}~0\}$ & $45 \pm 3$ & $0.75 \pm 0.09$ &
869     $3.0 \pm 0.2$ & $3.0 \pm 0.1$ & $3.6 \pm 0.2$ & $4 \pm 2$ \\
870 plouden 4252 Secondary Prism $\{1~1~\bar{2}~0\}$ & $43 \pm 2$ & $0.69 \pm 0.03$ &
871 plouden 4253 $3.5 \pm 0.1$ & $3.3 \pm 0.2$ & $3.2 \pm 0.2$ & $5 \pm 3$ \\
872 gezelter 4247 \bottomrule
873 gezelter 4243 \end{tabular}
874     \end{table}
875    
876     \end{document}