ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/iceWater2/iceWater3.tex
Revision: 4258
Committed: Tue Dec 16 15:53:23 2014 UTC (10 years, 7 months ago) by gezelter
Content type: application/x-tex
File size: 48732 byte(s)
Log Message:
New table

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 4258 \usepackage{tablefootnote}
19    
20 gezelter 4245 \bibpunct{(}{)}{,}{n}{,}{,}
21     \bibliographystyle{pnas2011}
22 gezelter 4243
23     %% OPTIONAL MACRO DEFINITIONS
24     \def\s{\sigma}
25     %%%%%%%%%%%%
26     %% For PNAS Only:
27     %\url{www.pnas.org/cgi/doi/10.1073/pnas.0709640104}
28     \copyrightyear{2014}
29     \issuedate{Issue Date}
30     \volume{Volume}
31     \issuenumber{Issue Number}
32     %\setcounter{page}{2687} %Set page number here if desired
33     %%%%%%%%%%%%
34    
35     \begin{document}
36    
37 gezelter 4254 \title{The different facets of ice have different hydrophilicities:
38     Friction at water / ice-I\textsubscript{h} interfaces}
39 gezelter 4243
40     \author{Patrick B. Louden\affil{1}{Department of Chemistry and Biochemistry, University of Notre Dame, Notre Dame,
41     IN 46556}
42     \and
43     J. Daniel Gezelter\affil{1}{}}
44    
45     \contributor{Submitted to Proceedings of the National Academy of Sciences
46     of the United States of America}
47    
48     %%%Newly updated.
49     %%% If significance statement need, then can use the below command otherwise just delete it.
50 gezelter 4257 \significancetext{Surface hydrophilicity is a measure of the
51     interaction strength between a solid surface and liquid water. Our
52     simulations show that the solid that is thought to be extremely
53     hydrophilic (ice) displays different behavior depending on which
54     crystal facet is presented to the liquid. This behavior is
55     potentially important in geophysics, in recognition of ice surfaces
56     by anti-freeze proteins, and in understanding how the friction
57     between ice and other solids may be mediated by a quasi-liquid layer
58     of water.}
59 gezelter 4243
60     \maketitle
61    
62     \begin{article}
63 gezelter 4245 \begin{abstract}
64 gezelter 4257 We present evidence that the prismatic and secondary prism facets
65     of ice-I$_\mathrm{h}$ crystals posess structural features that can
66     reduce the effective hydrophilicity of the ice/water
67     interface. The spreading dynamics of liquid water droplets on ice
68     facets exhibits long-time behavior that differs substantially for
69     the prismatic $\{1~0~\bar{1}~0\}$ and secondary prism
70     $\{1~1~\bar{2}~0\}$ facets when compared with the basal $\{0001\}$
71     and pyramidal $\{2~0~\bar{2}~1\}$ facets. We also present the
72     results of simulations of solid-liquid friction of the same four
73     crystal facets being drawn through liquid water. These simulation
74 gezelter 4245 techniques provide evidence that the two prismatic faces have an
75     effective surface area in contact with the liquid water of
76 gezelter 4257 approximately half of the total surface area of the crystal. The
77 gezelter 4245 ice / water interfacial widths for all four crystal facets are
78     similar (using both structural and dynamic measures), and were
79     found to be independent of the shear rate. Additionally,
80     decomposition of orientational time correlation functions show
81     position-dependence for the short- and longer-time decay
82     components close to the interface.
83 gezelter 4243 \end{abstract}
84    
85     \keywords{ice | water | interfaces | hydrophobicity}
86     \abbreviations{QLL, quasi liquid layer; MD, molecular dynamics; RNEMD,
87     reverse non-equilibrium molecular dynamics}
88    
89 gezelter 4245 \dropcap{S}urfaces can be characterized as hydrophobic or hydrophilic
90     based on the strength of the interactions with water. Hydrophobic
91     surfaces do not have strong enough interactions with water to overcome
92     the internal attraction between molecules in the liquid phase, and the
93     degree of hydrophilicity of a surface can be described by the extent a
94 gezelter 4254 droplet can spread out over the surface. The contact angle, $\theta$,
95     formed between the solid and the liquid depends on the free energies
96     of the three interfaces involved, and is given by Young's
97     equation~\cite{Young05},
98 gezelter 4245 \begin{equation}\label{young}
99     \cos\theta = (\gamma_{sv} - \gamma_{sl})/\gamma_{lv} .
100     \end{equation}
101     Here $\gamma_{sv}$, $\gamma_{sl}$, and $\gamma_{lv}$ are the free
102 gezelter 4254 energies of the solid/vapor, solid/liquid, and liquid/vapor interfaces,
103 gezelter 4245 respectively. Large contact angles, $\theta > 90^{\circ}$, correspond
104     to hydrophobic surfaces with low wettability, while small contact
105     angles, $\theta < 90^{\circ}$, correspond to hydrophilic surfaces.
106     Experimentally, measurements of the contact angle of sessile drops is
107     often used to quantify the extent of wetting on surfaces with
108     thermally selective wetting
109 gezelter 4254 characteristics~\cite{Tadanaga00,Liu04,Sun04}.
110 gezelter 4243
111 gezelter 4245 Nanometer-scale structural features of a solid surface can influence
112     the hydrophilicity to a surprising degree. Small changes in the
113     heights and widths of nano-pillars can change a surface from
114     superhydrophobic, $\theta \ge 150^{\circ}$, to hydrophilic, $\theta
115 plouden 4246 \sim 0^{\circ}$.\cite{Koishi09} This is often referred to as the
116 gezelter 4245 Cassie-Baxter to Wenzel transition. Nano-pillared surfaces with
117     electrically tunable Cassie-Baxter and Wenzel states have also been
118 gezelter 4254 observed~\cite{Herbertson06,Dhindsa06,Verplanck07,Ahuja08,Manukyan11}.
119 gezelter 4245 Luzar and coworkers have modeled these transitions on nano-patterned
120 gezelter 4254 surfaces~\cite{Daub07,Daub10,Daub11,Ritchie12}, and have found the
121 gezelter 4245 change in contact angle is due to the field-induced perturbation of
122 gezelter 4254 hydrogen bonding at the liquid/vapor interface~\cite{Daub07}.
123 gezelter 4245
124     One would expect the interfaces of ice to be highly hydrophilic (and
125     possibly the most hydrophilic of all solid surfaces). In this paper we
126     present evidence that some of the crystal facets of ice-I$_\mathrm{h}$
127 gezelter 4254 have structural features that can reduce the effective hydrophilicity.
128 gezelter 4245 Our evidence for this comes from molecular dynamics (MD) simulations
129     of the spreading dynamics of liquid droplets on these facets, as well
130     as reverse non-equilibrium molecular dynamics (RNEMD) simulations of
131     solid-liquid friction.
132    
133     Quiescent ice-I$_\mathrm{h}$/water interfaces have been studied
134     extensively using computer simulations. Haymet \textit{et al.}
135     characterized and measured the width of these interfaces for the
136     SPC~\cite{Karim90}, SPC/E~\cite{Gay02,Bryk02},
137     CF1~\cite{Hayward01,Hayward02} and TIP4P~\cite{Karim88} models, in
138     both neat water and with solvated
139     ions~\cite{Bryk04,Smith05,Wilson08,Wilson10}. Nada and Furukawa have
140     studied the width of basal/water and prismatic/water
141     interfaces~\cite{Nada95} as well as crystal restructuring at
142     temperatures approaching the melting point~\cite{Nada00}.
143    
144 gezelter 4243 The surface of ice exhibits a premelting layer, often called a
145 gezelter 4245 quasi-liquid layer (QLL), at temperatures near the melting point. MD
146     simulations of the facets of ice-I$_\mathrm{h}$ exposed to vacuum have
147     found QLL widths of approximately 10 \AA\ at 3 K below the melting
148 gezelter 4254 point~\cite{Conde08}. Similarly, Limmer and Chandler have used the mW
149 gezelter 4245 water model~\cite{Molinero09} and statistical field theory to estimate
150 gezelter 4254 QLL widths at similar temperatures to be about 3 nm~\cite{Limmer14}.
151 gezelter 4243
152 gezelter 4245 Recently, Sazaki and Furukawa have developed a technique using laser
153     confocal microscopy combined with differential interference contrast
154     microscopy that has sufficient spatial and temporal resolution to
155     visulaize and quantitatively analyze QLLs on ice crystals at
156 gezelter 4254 temperatures near melting~\cite{Sazaki10}. They have found the width of
157 gezelter 4245 the QLLs perpindicular to the surface at -2.2$^{o}$C to be 3-4 \AA\
158     wide. They have also seen the formation of two immiscible QLLs, which
159 gezelter 4254 displayed different dynamics on the crystal surface~\cite{Sazaki12}.
160 gezelter 4243
161 gezelter 4257 % There is now significant interest in the \textit{tribological}
162     % properties of ice/ice and ice/water interfaces in the geophysics
163     % community. Understanding the dynamics of solid-solid shearing that is
164     % mediated by a liquid layer~\cite{Cuffey99, Bell08} will aid in
165     % understanding the macroscopic motion of large ice
166     % masses~\cite{Casassa91, Sukhorukov13, Pritchard12, Lishman13}.
167 gezelter 4243
168     Using molecular dynamics simulations, Samadashvili has recently shown
169     that when two smooth ice slabs slide past one another, a stable
170 gezelter 4254 liquid-like layer develops between them~\cite{Samadashvili13}. In a
171 gezelter 4245 previous study, our RNEMD simulations of ice-I$_\mathrm{h}$ shearing
172     through liquid water have provided quantitative estimates of the
173 gezelter 4254 solid-liquid kinetic friction coefficients~\cite{Louden13}. These
174 gezelter 4245 displayed a factor of two difference between the basal and prismatic
175     facets. The friction was found to be independent of shear direction
176     relative to the surface orientation. We attributed facet-based
177     difference in liquid-solid friction to the 6.5 \AA\ corrugation of the
178     prismatic face which reduces the effective surface area of the ice
179     that is in direct contact with liquid water.
180 gezelter 4243
181 gezelter 4245 In the sections that follow, we outline the methodology used to
182     simulate droplet-spreading dynamics using standard MD and tribological
183     properties using RNEMD simulations. These simulation methods give
184     complementary results that point to the prismatic and secondary prism
185     facets having roughly half of their surface area in direct contact
186     with the liquid.
187 gezelter 4243
188 gezelter 4245 \section{Methodology}
189     \subsection{Construction of the Ice / Water Interfaces}
190     To construct the four interfacial ice/water systems, a proton-ordered,
191     zero-dipole crystal of ice-I$_\mathrm{h}$ with exposed strips of
192 gezelter 4247 H-atoms was created using Structure 6 of Hirsch and Ojam\"{a}e's set
193     of orthorhombic representations for ice-I$_{h}$~\cite{Hirsch04}. This
194     crystal structure was cleaved along the four different facets. The
195     exposed face was reoriented normal to the $z$-axis of the simulation
196     cell, and the structures were and extended to form large exposed
197     facets in rectangular box geometries. Liquid water boxes were created
198     with identical dimensions (in $x$ and $y$) as the ice, with a $z$
199     dimension of three times that of the ice block, and a density
200     corresponding to 1 g / cm$^3$. Each of the ice slabs and water boxes
201     were independently equilibrated at a pressure of 1 atm, and the
202     resulting systems were merged by carving out any liquid water
203     molecules within 3 \AA\ of any atoms in the ice slabs. Each of the
204     combined ice/water systems were then equilibrated at 225K, which is
205     the liquid-ice coexistence temperature for SPC/E
206 gezelter 4254 water~\cite{Bryk02}. Reference \citealp{Louden13} contains a more
207     detailed explanation of the construction of similar ice/water
208     interfaces. The resulting dimensions as well as the number of ice and
209     liquid water molecules contained in each of these systems are shown in
210     Table \ref{tab:method}.
211 gezelter 4243
212 gezelter 4247 The SPC/E water model~\cite{Berendsen87} has been extensively
213 plouden 4246 characterized over a wide range of liquid
214 gezelter 4254 conditions~\cite{Arbuckle02,Kuang12}, and its phase diagram has been
215     well studied~\cite{Baez95,Bryk04b,Sanz04b,Fennell:2005fk}, With longer
216 gezelter 4247 cutoff radii and careful treatment of electrostatics, SPC/E mostly
217     avoids metastable crystalline morphologies like
218     ice-\textit{i}~\cite{Fennell:2005fk} and ice-B~\cite{Baez95}. The
219 gezelter 4254 free energies and melting
220     points~\cite{Baez95,Arbuckle02,Gay02,Bryk02,Bryk04b,Sanz04b,Fennell:2005fk,Fernandez06,Abascal07,Vrbka07}
221 gezelter 4247 of various other crystalline polymorphs have also been calculated.
222     Haymet \textit{et al.} have studied quiescent Ice-I$_\mathrm{h}$/water
223     interfaces using the SPC/E water model, and have seen structural and
224     dynamic measurements of the interfacial width that agree well with
225     more expensive water models, although the coexistence temperature for
226     SPC/E is still well below the experimental melting point of real
227     water~\cite{Bryk02}. Given the extensive data and speed of this model,
228     it is a reasonable choice even though the temperatures required are
229     somewhat lower than real ice / water interfaces.
230 gezelter 4245
231 gezelter 4247 \subsection{Droplet Simulations}
232     Ice interfaces with a thickness of $\sim$~20~\AA\ were created as
233 gezelter 4245 described above, but were not solvated in a liquid box. The crystals
234     were then replicated along the $x$ and $y$ axes (parallel to the
235 gezelter 4247 surface) until a large surface ($>$ 126 nm\textsuperscript{2}) had
236     been created. The sizes and numbers of molecules in each of the
237     surfaces is given in Table \ref{tab:method}. Weak translational
238 gezelter 4254 restraining potentials with spring constants of 1.5~$\mathrm{kcal\
239     mol}^{-1}\mathrm{~\AA}^{-2}$ (prismatic and pyramidal facets) or
240     4.0~$\mathrm{kcal\ mol}^{-1}\mathrm{~\AA}^{-2}$ (basal facet) were
241     applied to the centers of mass of each molecule in order to prevent
242     surface melting, although the molecules were allowed to reorient
243     freely. A water doplet containing 2048 SPC/E molecules was created
244     separately. Droplets of this size can produce agreement with the Young
245     contact angle extrapolated to an infinite drop size~\cite{Daub10}. The
246     surfaces and droplet were independently equilibrated to 225 K, at
247     which time the droplet was placed 3-5~\AA\ above the surface. Five
248     statistically independent simulations were carried out for each facet,
249     and the droplet was placed at unique $x$ and $y$ locations for each of
250     these simulations. Each simulation was 5~ns in length and was
251     conducted in the microcanonical (NVE) ensemble. Representative
252     configurations for the droplet on the prismatic facet are shown in
253     figure \ref{fig:Droplet}.
254 gezelter 4243
255 gezelter 4247 \subsection{Shearing Simulations (Interfaces in Bulk Water)}
256    
257     To perform the shearing simulations, the velocity shearing and scaling
258     variant of reverse non-equilibrium molecular dynamics (VSS-RNEMD) was
259     employed \cite{Kuang12}. This method performs a series of simultaneous
260     non-equilibrium exchanges of linear momentum and kinetic energy
261     between two physically-separated regions of the simulation cell. The
262     system responds to this unphysical flux with velocity and temperature
263     gradients. When VSS-RNEMD is applied to bulk liquids, transport
264     properties like the thermal conductivity and the shear viscosity are
265     easily extracted assuming a linear response between the flux and the
266     gradient. At the interfaces between dissimilar materials, the same
267     method can be used to extract \textit{interfacial} transport
268     properties (e.g. the interfacial thermal conductance and the
269     hydrodynamic slip length).
270    
271     The kinetic energy flux (producing a thermal gradient) is necessary
272     when performing shearing simulations at the ice-water interface in
273     order to prevent the frictional heating due to the shear from melting
274 gezelter 4254 the crystal. Reference \citealp{Louden13} provides more details on the
275     VSS-RNEMD method as applied to ice-water interfaces. A representative
276     configuration of the solvated prismatic facet being sheared through
277     liquid water is shown in figure \ref{fig:Shearing}.
278 gezelter 4247
279 gezelter 4254 The exchanges between the two regions were carried out every 2 fs
280 gezelter 4257 (i.e. every time step). Although computationally expensive, this was
281     done to minimize the magnitude of each individual momentum exchange.
282     Because individual VSS-RNEMD exchanges conserve both total energy and
283     linear momentum, the method can be ``bolted-on'' to simulations in any
284     ensemble. The simulations of the pyramidal interface were performed
285     under the canonical (NVT) ensemble. When time correlation functions
286     were computed, the RNEMD simulations were done in the microcanonical
287     (NVE) ensemble. All simulations of the other interfaces were carried
288     out in the microcanonical ensemble.
289 gezelter 4247
290     \section{Results}
291     \subsection{Ice - Water Contact Angles}
292 plouden 4246
293     To determine the extent of wetting for each of the four crystal
294 gezelter 4247 facets, contact angles for liquid droplets on the ice surfaces were
295     computed using two methods. In the first method, the droplet is
296     assumed to form a spherical cap, and the contact angle is estimated
297     from the $z$-axis location of the droplet's center of mass
298     ($z_\mathrm{cm}$). This procedure was first described by Hautman and
299     Klein~\cite{Hautman91}, and was utilized by Hirvi and Pakkanen in
300     their investigation of water droplets on polyethylene and poly(vinyl
301     chloride) surfaces~\cite{Hirvi06}. For each stored configuration, the
302     contact angle, $\theta$, was found by inverting the expression for the
303     location of the droplet center of mass,
304 plouden 4246 \begin{equation}\label{contact_1}
305 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} ,
306 plouden 4246 \end{equation}
307 gezelter 4247 where $R_{0}$ is the radius of the free water droplet.
308 plouden 4246
309 gezelter 4247 The second method for obtaining the contact angle was described by
310     Ruijter, Blake, and Coninck~\cite{Ruijter99}. This method uses a
311     cylindrical averaging of the droplet's density profile. A threshold
312     density of 0.5 g cm\textsuperscript{-3} is used to estimate the
313     location of the edge of the droplet. The $r$ and $z$-dependence of
314     the droplet's edge is then fit to a circle, and the contact angle is
315     computed from the intersection of the fit circle with the $z$-axis
316     location of the solid surface. Again, for each stored configuration,
317     the density profile in a set of annular shells was computed. Due to
318     large density fluctuations close to the ice, all shells located within
319     2 \AA\ of the ice surface were left out of the circular fits. The
320     height of the solid surface ($z_\mathrm{suface}$) along with the best
321 gezelter 4254 fitting origin ($z_\mathrm{droplet}$) and radius
322 gezelter 4247 ($r_\mathrm{droplet}$) of the droplet can then be used to compute the
323     contact angle,
324     \begin{equation}
325 gezelter 4254 \theta = 90 + \frac{180}{\pi} \sin^{-1}\left(\frac{z_\mathrm{droplet} -
326 gezelter 4247 z_\mathrm{surface}}{r_\mathrm{droplet}} \right).
327     \end{equation}
328     Both methods provided similar estimates of the dynamic contact angle,
329     although the first method is significantly less prone to noise, and
330     is the method used to report contact angles below.
331    
332     Because the initial droplet was placed above the surface, the initial
333 gezelter 4250 value of 180$^{\circ}$ decayed over time (See figure
334     \ref{fig:ContactAngle}). Each of these profiles were fit to a
335     biexponential decay, with a short-time contribution ($\tau_c$) that
336     describes the initial contact with the surface, a long time
337     contribution ($\tau_s$) that describes the spread of the droplet over
338     the surface, and a constant ($\theta_\infty$) to capture the
339     infinite-time estimate of the equilibrium contact angle,
340 gezelter 4247 \begin{equation}
341 gezelter 4250 \theta(t) = \theta_\infty + (180-\theta_\infty) \left[ a e^{-t/\tau_c} +
342     (1-a) e^{-t/\tau_s} \right]
343 gezelter 4247 \end{equation}
344 gezelter 4250 We have found that the rate for water droplet spreading across all
345     four crystal facets, $k_\mathrm{spread} = 1/\tau_s \approx$ 0.7
346     ns$^{-1}$. However, the basal and pyramidal facets produced estimated
347     equilibrium contact angles, $\theta_\infty \approx$ 35$^{o}$, while
348 gezelter 4247 prismatic and secondary prismatic had values for $\theta_\infty$ near
349 gezelter 4250 43$^{o}$ as seen in Table \ref{tab:kappa}.
350 gezelter 4247
351 gezelter 4254 These results indicate that the basal and pyramidal facets are more
352 gezelter 4257 hydrophilic by traditional measures than the prismatic and secondary
353     prism facets, and surprisingly, that the differential hydrophilicities
354     of the crystal facets is not reflected in the spreading rate of the
355     droplet.
356 gezelter 4247
357 plouden 4246 % This is in good agreement with our calculations of friction
358     % coefficients, in which the basal
359     % and pyramidal had a higher coefficient of kinetic friction than the
360     % prismatic and secondary prismatic. Due to this, we beleive that the
361     % differences in friction coefficients can be attributed to the varying
362     % hydrophilicities of the facets.
363    
364 gezelter 4257 \subsection{Solid-liquid friction of the interfaces}
365 gezelter 4254 In a bulk fluid, the shear viscosity, $\eta$, can be determined
366     assuming a linear response to a shear stress,
367 plouden 4246 \begin{equation}\label{Shenyu-11}
368 gezelter 4254 j_{z}(p_{x}) = \eta \frac{\partial v_{x}}{\partial z}.
369 plouden 4246 \end{equation}
370 gezelter 4254 Here $j_{z}(p_{x})$ is the flux (in $x$-momentum) that is transferred
371     in the $z$ direction (i.e. the shear stress). The RNEMD simulations
372     impose an artificial momentum flux between two regions of the
373     simulation, and the velocity gradient is the fluid's response. This
374     technique has now been applied quite widely to determine the
375     viscosities of a number of bulk fluids~\cite{}.
376    
377     At the interface between two phases (e.g. liquid / solid) the same
378     momentum flux creates a velocity difference between the two materials,
379     and this can be used to define an interfacial friction coefficient
380     ($\kappa$),
381     \begin{equation}\label{Shenyu-13}
382     j_{z}(p_{x}) = \kappa \left[ v_{x}(liquid) - v_{x}(solid) \right]
383 plouden 4246 \end{equation}
384 gezelter 4254 where $v_{x}(liquid)$ and $v_{x}(solid)$ are the velocities measured
385     directly adjacent to the interface.
386    
387     The simulations described here contain significant quantities of both
388     liquid and solid phases, and the momentum flux must traverse a region
389     of the liquid that is simultaneously under a thermal gradient. Since
390     the liquid has a temperature-dependent shear viscosity, $\eta(T)$,
391     estimates of the solid-liquid friction coefficient can be obtained if
392     one knows the viscosity of the liquid at the interface (i.e. at the
393     melting temperature, $T_m$),
394 plouden 4246 \begin{equation}\label{kappa-2}
395 gezelter 4254 \kappa = \frac{\eta(T_{m})}{\left[v_{x}(fluid)-v_{x}(solid)\right]}\left(\frac{\partial v_{x}}{\partial z}\right).
396 plouden 4246 \end{equation}
397 gezelter 4254 For SPC/E, the melting temperature of Ice-I$_\mathrm{h}$ is estimated
398     to be 225~K~\cite{Bryk02}. To obtain the value of
399     $\eta(225\mathrm{~K})$ for the SPC/E model, a $31.09 \times 29.38
400     \times 124.39$ \AA\ box with 3744 water molecules in a disordered
401     configuration was equilibrated to 225~K, and five
402     statistically-independent shearing simulations were performed (with
403 gezelter 4257 imposed fluxes that spanned a range of $3 \rightarrow 13
404     \mathrm{~m~s}^{-1}$ ). Each simulation was conducted in the
405     microcanonical ensemble with total simulation times of 5 ns. The
406     VSS-RNEMD exchanges were carried out every 2 fs. We estimate
407     $\eta(225\mathrm{~K})$ to be 0.0148 $\pm$ 0.0007 Pa s for SPC/E,
408     roughly ten times larger than the shear viscosity previously computed
409     at 280~K~\cite{Kuang12}.
410 plouden 4246
411 gezelter 4257 The interfacial friction coefficient can equivalently be expressed as
412     the ratio of the viscosity of the fluid to the hydrodynamic slip
413     length, $\kappa = \eta / \delta$. The slip length is an indication of
414     strength of the interactions between the solid and liquid phases,
415     although the connection between slip length and surface hydrophobicity
416     is not yet clear. In some simulations, the slip length has been found
417     to have a link to the effective surface
418     hydrophobicity~\cite{Sendner:2009uq}, although Ho \textit{et al.} have
419     found that liquid water can also slip on hydrophilic
420     surfaces~\cite{Ho:2011zr}. Experimental evidence for a direct tie
421     between slip length and hydrophobicity is also not
422 gezelter 4254 definitive. Total-internal reflection particle image velocimetry
423     (TIR-PIV) studies have suggested that there is a link between slip
424     length and effective
425     hydrophobicity~\cite{Lasne:2008vn,Bouzigues:2008ys}. However, recent
426     surface sensitive cross-correlation spectroscopy (TIR-FCCS)
427     measurements have seen similar slip behavior for both hydrophobic and
428     hydrophilic surfaces~\cite{Schaeffel:2013kx}.
429 plouden 4246
430 gezelter 4254 In each of the systems studied here, the interfacial temperature was
431     kept fixed to 225K, which ensured the viscosity of the fluid at the
432     interace was identical. Thus, any significant variation in $\kappa$
433     between the systems is a direct indicator of the slip length and the
434     effective interaction strength between the solid and liquid phases.
435 gezelter 4243
436 gezelter 4254 The calculated $\kappa$ values found for the four crystal facets of
437     Ice-I$_\mathrm{h}$ are shown in Table \ref{tab:kappa}. The basal and
438     pyramidal facets were found to have similar values of $\kappa \approx
439     6$ ($\times 10^{-4} \mathrm{amu~\AA}^{-2}\mathrm{fs}^{-1}$), while the
440     prismatic and secondary prism facets exhibited $\kappa \approx 3$
441     ($\times 10^{-4} \mathrm{amu~\AA}^{-2}\mathrm{fs}^{-1}$). These
442     results are also essentially independent of shearing direction
443     relative to features on the surface of the facets. The friction
444     coefficients indicate that the basal and pyramidal facets have
445     significantly stronger interactions with liquid water than either of
446     the two prismatic facets. This is in agreement with the contact angle
447     results above - both of the high-friction facets exhbited smaller
448     contact angles, suggesting that the solid-liquid friction is
449     correlated with the hydrophilicity of these facets.
450 gezelter 4243
451 gezelter 4254 \subsection{Structural measures of interfacial width under shear}
452     One of the open questions about ice/water interfaces is whether the
453     thickness of the 'slush-like' quasi-liquid layer (QLL) depends on the
454     facet of ice presented to the water. In the QLL region, the water
455     molecules are ordered differently than in either the solid or liquid
456     phases, and also exhibit distinct dynamical behavior. The width of
457     this quasi-liquid layer has been estimated by finding the distance
458     over which structural order parameters or dynamic properties change
459     from their bulk liquid values to those of the solid ice. The
460     properties used to find interfacial widths have included the local
461     density, the diffusion constant, and the translational and
462     orientational order
463     parameters~\cite{Karim88,Karim90,Hayward01,Hayward02,Bryk02,Gay02,Louden13}.
464    
465     The VSS-RNEMD simulations impose thermal and velocity gradients.
466     These gradients perturb the momenta of the water molecules, so
467     parameters that depend on translational motion are often measuring the
468     momentum exchange, and not physical properties of the interface. As a
469     structural measure of the interface, we have used the local
470     tetrahedral order parameter to estimate the width of the interface.
471     This quantity was originally described by Errington and
472     Debenedetti~\cite{Errington01} and has been used in bulk simulations
473     by Kumar \textit{et al.}~\cite{Kumar09}. It has previously been used
474     in ice/water interfaces by by Bryk and Haymet~\cite{Bryk04b}.
475    
476     To determine the structural widths of the interfaces under shear, each
477     of the systems was divided into 100 bins along the $z$-dimension, and
478     the local tetrahedral order parameter (Eq. 5 in Reference
479     \citealp{Louden13}) was time-averaged in each bin for the duration of
480     the shearing simulation. The spatial dependence of this order
481     parameter, $q(z)$, is the tetrahedrality profile of the interface. A
482     representative profile for the pyramidal facet is shown in circles in
483     panel $a$ of figure \ref{fig:pyrComic}. The $q(z)$ function has a
484     range of $(0,1)$, where a value of unity indicates a perfectly
485     tetrahedral environment. The $q(z)$ for the bulk liquid was found to
486     be $\approx~0.77$, while values of $\approx~0.92$ were more common in
487     the ice. The tetrahedrality profiles were fit using a hyperbolic
488     tangent function (see Eq. 6 in Reference \citealp{Louden13}) designed
489     to smoothly fit the bulk to ice transition while accounting for the
490     weak thermal gradient. In panels $b$ and $c$, the resulting thermal
491     and velocity gradients from an imposed kinetic energy and momentum
492     fluxes can be seen. The vertical dotted lines traversing all three
493     panels indicate the midpoints of the interface as determined by the
494     tetrahedrality profiles.
495 gezelter 4243
496 gezelter 4254 We find the interfacial width to be $3.2 \pm 0.2$ \AA\ (pyramidal) and
497     $3.2 \pm 0.2$ \AA\ (secondary prism) for the control systems with no
498     applied momentum flux. This is similar to our previous results for the
499     interfacial widths of the quiescent basal ($3.2 \pm 0.4$ \AA) and
500     prismatic systems ($3.6 \pm 0.2$ \AA).
501 gezelter 4243
502 gezelter 4254 Over the range of shear rates investigated, $0.4 \rightarrow
503     6.0~\mathrm{~m~s}^{-1}$ for the pyramidal system and $0.6 \rightarrow
504     5.2~\mathrm{~m~s}^{-1}$ for the secondary prism, we found no
505     significant change in the interfacial width. The mean interfacial
506     widths are collected in table \ref{tab:kappa}. This follows our
507     previous findings of the basal and prismatic systems, in which the
508     interfacial widths of the basal and prismatic facets were also found
509     to be insensitive to the shear rate~\cite{Louden13}.
510 gezelter 4243
511 gezelter 4254 The similarity of these interfacial width estimates indicate that the
512     particular facet of the exposed ice crystal has little to no effect on
513     how far into the bulk the ice-like structural ordering persists. Also,
514     it appears that for the shearing rates imposed in this study, the
515     interfacial width is not structurally modified by the movement of
516     water over the ice.
517 gezelter 4243
518 gezelter 4254 \subsection{Dynamic measures of interfacial width under shear}
519 gezelter 4257 The spatially-resolved orientational time correlation function,
520 gezelter 4243 \begin{equation}\label{C(t)1}
521 gezelter 4257 C_{2}(z,t)=\langle P_{2}(\mathbf{u}_i(0)\cdot \mathbf{u}_i(t))
522     \delta(z_i(0) - z) \rangle,
523 gezelter 4243 \end{equation}
524 gezelter 4257 provides local information about the decorrelation of molecular
525     orientations in time. Here, $P_{2}$ is the second-order Legendre
526     polynomial, and $\mathbf{u}_i$ is the molecular vector that bisects
527     the HOH angle of molecule $i$. The angle brackets indicate an average
528     over all the water molecules, and the delta function restricts the
529     average to specific regions. In the crystal, decay of $C_2(z,t)$ is
530     incomplete, while liquid water correlation times are typically
531     measured in ps. Observing the spatial-transition between the decay
532     regimes can define a dynamic measure of the interfacial width.
533 gezelter 4243
534 gezelter 4257 Each of the systems was divided into bins along the $z$-dimension
535     ($\approx$ 3 \AA\ wide) and $C_2(z,t)$ was computed using only those
536     molecules that were in the bin at the initial time. The
537     time-dependence was fit to a triexponential decay, with three time
538     constants: $\tau_{short}$, measuring the librational motion of the
539     water molecules, $\tau_{middle}$, measuring the timescale for breaking
540     and making of hydrogen bonds, and $\tau_{long}$, corresponding to the
541     translational motion of the water molecules. An additional constant
542     was introduced in the fits to describe molecules in the crystal which
543     do not experience long-time orientational decay.
544 gezelter 4243
545 gezelter 4257 In Figures S5-S8 in the supporting information, the $z$-coordinate
546     profiles for the three decay constants, $\tau_{short}$,
547     $\tau_{middle}$, and $\tau_{long}$ for the different interfaces are
548     shown. Figures S5 \& S6 are new results, and Figures S7 \& S8 are
549     updated plots from Ref \citealp{Louden13}. In the liquid regions of
550     all four interfaces, we observe $\tau_{middle}$ and $\tau_{long}$ to
551     have approximately consistent values of $3-6$ ps and $30-40$ ps,
552     respectively. Both of these times increase in value approaching the
553     interface. Approaching the interface, we also observe that
554     $\tau_{short}$ decreases from its liquid-state value of $72-76$ fs.
555     The approximate values for the decay constants and the trends
556     approaching the interface match those reported previously for the
557     basal and prismatic interfaces.
558 gezelter 4243
559 gezelter 4257 We have estimated the dynamic interfacial width $d_\mathrm{dyn}$ by
560     fitting the profiles of all the three orientational time constants
561     with an exponential decay to the bulk-liquid behavior,
562 gezelter 4243 \begin{equation}\label{tauFit}
563 gezelter 4257 \tau(z)\approx\tau_{liquid}+(\tau_{wall}-\tau_{liquid})e^{-(z-z_{wall})/d_\mathrm{dyn}}
564 gezelter 4243 \end{equation}
565 gezelter 4257 where $\tau_{liquid}$ and $\tau_{wall}$ are the liquid and projected
566     wall values of the decay constants, $z_{wall}$ is the location of the
567     interface, as measured by the structural order parameter. These
568     values are shown in table \ref{tab:kappa}. Because the bins must be
569     quite wide to obtain reasonable profiles of $C_2(z,t)$, the error
570     estimates for the dynamic widths of the interface are significantly
571     larger than for the structural widths. However, all four interfaces
572     exhibit dynamic widths that are significantly below 1~nm, and are in
573     reasonable agreement with the structural width above.
574 gezelter 4243
575 gezelter 4257 \section{Conclusions}
576     In this work, we used MD simulations to measure the advancing contact
577     angles of water droplets on the basal, prismatic, pyramidal, and
578     secondary prism facets of Ice-I$_\mathrm{h}$. Although there was no
579     significant change in the \textit{rate} at which the droplets spread
580     over the surface, the long-time behavior indicates that we should
581     expect to see larger equilibrium contact angles for the two prismatic
582     facets.
583 gezelter 4243
584 gezelter 4257 We have also used RNEMD simulations of water interfaces with the same
585     four crystal facets to compute solid-liquid friction coefficients. We
586     have observed coefficients of friction that differ by a factor of two
587     between the two prismatic facets and the basal and pyramidal facets.
588     Because the solid-liquid friction coefficient is directly tied to the
589     hydrodynamic slip length, this suggests that there are significant
590     differences in the overall interaction strengths between these facets
591     and the liquid layers immediately in contact with them.
592 gezelter 4243
593 gezelter 4257 The agreement between these two measures have lead us to conclude that
594     the two prismatic facets have a lower hydrophilicity than either the
595     basal or pyramidal facets. One possible explanation of this behavior
596     is that the face presented by both prismatic facets consists of deep,
597     narrow channels (i.e. stripes of adjacent rows of pairs of
598     hydrodgen-bound water molecules). At the surfaces of these facets,
599     the channels are 6.35 \AA\ wide and the sub-surface ice layer is 2.25
600     \AA\ below (and therefore blocked from hydrogen bonding with the
601     liquid). This means that only 1/2 of the surface molecules can form
602     hydrogen bonds with liquid-phase molecules.
603 gezelter 4243
604 gezelter 4257 In the basal plane, the surface features are narrower (4.49 \AA) and
605     shallower (1.3 \AA), while the pyramidal face has much wider channels
606     (8.65 \AA) which are also quite shallow (1.37 \AA). These features
607     allow liquid phase molecules to form hydrogen bonds with all of the
608     surface molecules in the basal and pyramidal facets. This means that
609     for similar surface areas, the two prismatic facets have an effective
610     hydrogen bonding surface area of half of the basal and pyramidal
611     facets. The reduction in the effective surface area would explain
612     much of the behavior observed in our simulations.
613 gezelter 4243
614     \begin{acknowledgments}
615     Support for this project was provided by the National
616     Science Foundation under grant CHE-1362211. Computational time was
617     provided by the Center for Research Computing (CRC) at the
618     University of Notre Dame.
619     \end{acknowledgments}
620    
621     \bibliography{iceWater}
622     % *****************************************
623     % There is significant interest in the properties of ice/ice and ice/water
624     % interfaces in the geophysics community. Most commonly, the results of shearing
625     % two ice blocks past one
626     % another\cite{Casassa91, Sukhorukov13, Pritchard12, Lishman13} or the shearing
627     % of ice through water\cite{Cuffey99, Bell08}. Using molecular dynamics
628     % simulations, Samadashvili has recently shown that when two smooth ice slabs
629     % slide past one another, a stable liquid-like layer develops between
630     % them\cite{Samadashvili13}. To fundamentally understand these processes, a
631     % molecular understanding of the ice/water interfaces is needed.
632    
633     % Investigation of the ice/water interface is also crucial in understanding
634     % processes such as nucleation, crystal
635     % growth,\cite{Han92, Granasy95, Vanfleet95} and crystal
636     % melting\cite{Weber83, Han92, Sakai96, Sakai96B}. Insight gained to these
637     % properties can also be applied to biological systems of interest, such as
638     % the behavior of the antifreeze protein found in winter
639     % flounder,\cite{Wierzbicki07,Chapsky97} and certain terrestial
640     % arthropods.\cite{Duman:2001qy,Meister29012013} Elucidating the properties which
641     % give rise to these processes through experimental techniques can be expensive,
642     % complicated, and sometimes infeasible. However, through the use of molecular
643     % dynamics simulations much of the problems of investigating these properties
644     % are alleviated.
645    
646     % Understanding ice/water interfaces inherently begins with the isolated
647     % systems. There has been extensive work parameterizing models for liquid water,
648     % such as the SPC\cite{Berendsen81}, SPC/E\cite{Berendsen87},
649     % TIP4P\cite{Jorgensen85}, TIP4P/2005\cite{Abascal05},
650     % ($\dots$), and more recently, models for simulating
651     % the solid phases of water, such as the TIP4P/Ice\cite{Abascal05b} model. The
652     % melting point of various crystal structures of ice have been calculated for
653     % many of these models
654     % (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}),
655     % and the partial or complete phase diagram for the model has been determined
656     % (SPC/E\cite{Baez95,Bryk04b,Sanz04b}, TIP4P\cite{Sanz04,Sanz04b,Koyama04}, TIP5P\cite{Sanz04,Koyama04}).
657     % Knowing the behavior and melting point for these models has enabled an initial
658     % investigation of ice/water interfaces.
659    
660     % The Ice-I$_\mathrm{h}$/water quiescent interface has been extensively studied
661     % over the past 30 years by theory and experiment. Haymet \emph{et al.} have
662     % done significant work characterizing and quantifying the width of these
663     % interfaces for the SPC,\cite{Karim90} SPC/E,\cite{Gay02,Bryk02},
664     % CF1,\cite{Hayward01,Hayward02} and TIP4P\cite{Karim88} models for water. In
665     % recent years, Haymet has focused on investigating the effects cations and
666     % anions have on crystal nucleaion and
667     % melting.\cite{Bryk04,Smith05,Wilson08,Wilson10} Nada and Furukawa have studied
668     % the the basal- and prismatic-water interface width\cite{Nada95}, crystal
669     % surface restructuring at temperatures approaching the melting
670     % point\cite{Nada00}, and the mechanism of ice growth inhibition by antifreeze
671     % proteins\cite{Nada08,Nada11,Nada12}. Nada has developed a six-site water model
672     % for ice/water interfaces near the melting point\cite{Nada03}, and studied the
673     % dependence of crystal growth shape on applied pressure\cite{Nada11b}. Using
674     % this model, Nada and Furukawa have established differential
675     % growth rates for the basal, prismatic, and secondary prismatic facets of
676     % Ice-I$_\mathrm{h}$ and found their origins due to a reordering of the hydrogen
677     % bond network in water near the interface\cite{Nada05}. While the work
678     % described so far has mainly focused on bulk water on ice, there is significant
679     % interest in thin films of water on ice surfaces as well.
680    
681     % It is well known that the surface of ice exhibits a premelting layer at
682     % temperatures near the melting point, often called a quasi-liquid layer (QLL).
683     % Molecular dynamics simulations of the facets of ice-I$_\mathrm{h}$ exposed
684     % to vacuum performed by Conde, Vega and Patrykiejew have found QLL widths of
685     % approximately 10 \AA\ at 3 K below the melting point\cite{Conde08}.
686     % Similarly, Limmer and Chandler have used course grain simulations and
687     % statistical field theory to estimated QLL widths at the same temperature to
688     % be about 3 nm\cite{Limmer14}.
689     % Recently, Sazaki and Furukawa have developed an experimental technique with
690     % sufficient spatial and temporal resolution to visulaize and quantitatively
691     % analyze QLLs on ice crystals at temperatures near melting\cite{Sazaki10}. They
692     % have found the width of the QLLs perpindicular to the surface at -2.2$^{o}$C
693     % to be $\mathcal{O}$(\AA). They have also seen the formation of two immiscible
694     % QLLs, which displayed different stabilities and dynamics on the crystal
695     % surface\cite{Sazaki12}. Knowledge of the hydrophilicities of each
696     % of the crystal facets would help further our understanding of the properties
697     % and dynamics of the QLLs.
698    
699     % Presented here is the follow up to our previous paper\cite{Louden13}, in which
700     % the basal and prismatic facets of an ice-I$_\mathrm{h}$/water interface were
701     % investigated where the ice was sheared relative to the liquid. By using a
702     % recently developed velocity shearing and scaling approach to reverse
703     % non-equilibrium molecular dynamics (VSS-RNEMD), simultaneous temperature and
704     % velocity gradients can be applied to the system, which allows for measurment
705     % of friction and thermal transport properties while maintaining a stable
706     % interfacial temperature\cite{Kuang12}. Structural analysis and dynamic
707     % correlation functions were used to probe the interfacial response to a shear,
708     % and the resulting solid/liquid kinetic friction coefficients were reported.
709     % In this paper we present the same analysis for the pyramidal and secondary
710     % prismatic facets, and show that the differential interfacial friction
711     % coefficients for the four facets of ice-I$_\mathrm{h}$ are determined by their
712     % relative hydrophilicity by means of dynamics water contact angle
713     % simulations.
714    
715     % The local tetrahedral order parameter, $q(z)$, is given by
716     % \begin{equation}
717     % q(z) = \int_0^L \sum_{k=1}^{N} \Bigg(1 -\frac{3}{8}\sum_{i=1}^{3}
718     % \sum_{j=i+1}^{4} \bigg(\cos\psi_{ikj}+\frac{1}{3}\bigg)^2\Bigg)
719     % \delta(z_{k}-z)\mathrm{d}z \Bigg/ N_z
720     % \label{eq:qz}
721     % \end{equation}
722     % where $\psi_{ikj}$ is the angle formed between the oxygen sites of molecules
723     % $i$,$k$, and $j$, where the centeral oxygen is located within molecule $k$ and
724     % molecules $i$ and $j$ are two of the closest four water molecules
725     % around molecule $k$. All four closest neighbors of molecule $k$ are also
726     % required to reside within the first peak of the pair distribution function
727     % for molecule $k$ (typically $<$ 3.41 \AA\ for water).
728     % $N_z = \int\delta(z_k - z) \mathrm{d}z$ is a normalization factor to account
729     % for the varying population of molecules within each finite-width bin.
730    
731    
732     % The hydrophobicity or hydrophilicity of a surface can be described by the
733     % extent a droplet of water wets the surface. The contact angle formed between
734     % the solid and the liquid, $\theta$, which relates the free energies of the
735     % three interfaces involved, is given by Young's equation.
736     % \begin{equation}\label{young}
737     % \cos\theta = (\gamma_{sv} - \gamma_{sl})/\gamma_{lv}
738     % \end{equation}
739     % Here $\gamma_{sv}$, $\gamma_{sl}$, and $\gamma_{lv}$ are the free energies
740     % of the solid/vapor, solid/liquid, and liquid/vapor interfaces respectively.
741     % Large contact angles ($\theta$ $\gg$ 90\textsuperscript{o}) correspond to low
742     % wettability and hydrophobic surfaces, while small contact angles
743     % ($\theta$ $\ll$ 90\textsuperscript{o}) correspond to high wettability and
744     % hydrophilic surfaces. Experimentally, measurements of the contact angle
745     % of sessile drops has been used to quantify the extent of wetting on surfaces
746     % with thermally selective wetting charactaristics\cite{Tadanaga00,Liu04,Sun04},
747     % as well as nano-pillared surfaces with electrically tunable Cassie-Baxter and
748     % Wenzel states\cite{Herbertson06,Dhindsa06,Verplanck07,Ahuja08,Manukyan11}.
749     % Luzar and coworkers have done significant work modeling these transitions on
750     % nano-patterned surfaces\cite{Daub07,Daub10,Daub11,Ritchie12}, and have found
751     % the change in contact angle to be due to the external field perturbing the
752     % hydrogen bonding of the liquid/vapor interface\cite{Daub07}.
753    
754 gezelter 4257 % SI stuff:
755 gezelter 4243
756 gezelter 4257 % Correlation functions:
757     % To compute eq. \eqref{C(t)1}, each 0.5 ns simulation was
758     % followed by an additional 200 ps NVE simulation during which the
759     % position and orientations of each molecule were recorded every 0.1 ps.
760 gezelter 4243
761 gezelter 4257
762    
763    
764 gezelter 4243 \end{article}
765    
766     \begin{figure}
767 gezelter 4247 \includegraphics[width=\linewidth]{Droplet}
768     \caption{\label{fig:Droplet} Computational model of a droplet of
769     liquid water spreading over the prismatic $\{1~0~\bar{1}~0\}$ facet
770 gezelter 4250 of ice, before (left) and 2.6 ns after (right) being introduced to the
771 gezelter 4247 surface. The contact angle ($\theta$) shrinks as the simulation
772     proceeds, and the long-time behavior of this angle is used to
773     estimate the hydrophilicity of the facet.}
774     \end{figure}
775    
776     \begin{figure}
777 gezelter 4250 \includegraphics[width=2in]{Shearing}
778 gezelter 4247 \caption{\label{fig:Shearing} Computational model of a slab of ice
779 gezelter 4250 being sheared through liquid water. In this figure, the ice is
780     presenting two copies of the prismatic $\{1~0~\bar{1}~0\}$ facet
781     towards the liquid phase. The RNEMD simulation exchanges both
782     linear momentum (indicated with arrows) and kinetic energy between
783     the central box and the box that spans the cell boundary. The
784     system responds with weak thermal gradient and a velocity profile
785     that shears the ice relative to the surrounding liquid.}
786 gezelter 4247 \end{figure}
787    
788     \begin{figure}
789 gezelter 4250 \includegraphics[width=\linewidth]{ContactAngle}
790     \caption{\label{fig:ContactAngle} The dynamic contact angle of a
791     droplet after approaching each of the four ice facets. The decay to
792     an equilibrium contact angle displays similar dynamics. Although
793     all the surfaces are hydrophilic, the long-time behavior stabilizes
794     to significantly flatter droplets for the basal and pyramidal
795     facets. This suggests a difference in hydrophilicity for these
796     facets compared with the two prismatic facets.}
797 gezelter 4243 \end{figure}
798    
799 gezelter 4257 % \begin{figure}
800     % \includegraphics[width=\linewidth]{Pyr_comic_strip}
801     % \caption{\label{fig:pyrComic} Properties of the pyramidal interface
802     % being sheared through water at 3.8 ms\textsuperscript{-1}. Lower
803     % panel: the local tetrahedral order parameter, $q(z)$, (circles) and
804     % the hyperbolic tangent fit (turquoise line). Middle panel: the
805     % imposed thermal gradient required to maintain a fixed interfacial
806     % temperature of 225 K. Upper panel: the transverse velocity gradient
807     % that develops in response to an imposed momentum flux. The vertical
808     % dotted lines indicate the locations of the midpoints of the two
809     % interfaces.}
810     % \end{figure}
811 gezelter 4243
812 gezelter 4250 % \begin{figure}
813     % \includegraphics[width=\linewidth]{SP_comic_strip}
814     % \caption{\label{fig:spComic} The secondary prismatic interface with a shear
815     % rate of 3.5 \
816     % ms\textsuperscript{-1}. Panel descriptions match those in figure \ref{fig:pyrComic}.}
817     % \end{figure}
818    
819 gezelter 4257 % \begin{figure}
820     % \includegraphics[width=\linewidth]{Pyr-orient}
821     % \caption{\label{fig:PyrOrient} The three decay constants of the
822     % orientational time correlation function, $C_2(z,t)$, for water as a
823     % function of distance from the center of the ice slab. The vertical
824     % dashed line indicates the edge of the pyramidal ice slab determined
825     % by the local order tetrahedral parameter. The control (circles) and
826     % sheared (squares) simulations were fit using shifted-exponential
827     % decay (see Eq. 9 in Ref. \citealp{Louden13}).}
828     % \end{figure}
829 gezelter 4243
830 gezelter 4250 % \begin{figure}
831     % \includegraphics[width=\linewidth]{SP-orient-less}
832 gezelter 4257 % \caption{\label{fig:SPorient} Decay constants for $C_2(z,t)$ at the secondary
833 gezelter 4250 % prismatic face. Panel descriptions match those in \ref{fig:PyrOrient}.}
834     % \end{figure}
835 gezelter 4243
836    
837     \begin{table}[h]
838     \centering
839 gezelter 4247 \caption{Sizes of the droplet and shearing simulations. Cell
840     dimensions are measured in \AA. \label{tab:method}}
841     \begin{tabular}{r|cccc|ccccc}
842     \toprule
843     \multirow{2}{*}{Interface} & \multicolumn{4}{c|}{Droplet} & \multicolumn{5}{c}{Shearing} \\
844     & $N_\mathrm{ice}$ & $N_\mathrm{droplet}$ & $L_x$ & $L_y$ & $N_\mathrm{ice}$ & $N_\mathrm{liquid}$ & $L_x$ & $L_y$ & $L_z$ \\
845     \midrule
846     Basal $\{0001\}$ & 12960 & 2048 & 134.70 & 140.04 & 900 & 1846 & 23.87 & 35.83 & 98.64 \\
847     Pyramidal $\{2~0~\bar{2}~1\}$ & 11136 & 2048 & 143.75 & 121.41 & 1216 & 2203 & 37.47 & 29.50 & 93.02 \\
848     Prismatic $\{1~0~\bar{1}~0\}$ & 9900 & 2048 & 110.04 & 115.00 & 3000 & 5464 & 35.95 & 35.65 & 205.77 \\
849     Secondary Prism $\{1~1~\bar{2}~0\}$ & 11520 & 2048 & 146.72 & 124.48 & 3840 & 8176 & 71.87 & 31.66 & 161.55 \\
850     \bottomrule
851 gezelter 4243 \end{tabular}
852     \end{table}
853    
854    
855     \begin{table}[h]
856     \centering
857 gezelter 4258 \caption{Structural and dynamic properties of the interfaces of
858     Ice-I$_\mathrm{h}$ with water.\label{tab:kappa}}
859     \begin{tabular}{r|cc|cc|cccc}
860 gezelter 4247 \toprule
861 gezelter 4258 \multirow{2}{*}{Interface} & \multicolumn{2}{c|}{Channel Size} &\multicolumn{2}{c|}{Droplet} & \multicolumn{4}{c}{Shearing\footnotemark[1]}\\
862     & Width (\AA) & Depth (\AA) & $\theta_{\infty}$ ($^\circ$) & $k_\mathrm{spread}$ (ns\textsuperscript{-1}) &
863 gezelter 4257 $\kappa_{x}$ & $\kappa_{y}$ & $d_\mathrm{struct}$ (\AA) & $d_\mathrm{dyn}$ (\AA) \\
864 gezelter 4247 \midrule
865 gezelter 4258 Basal $\{0001\}$ & 4.49 & 1.30 & $34.1(9)$ &$0.60(7)$
866     & $5.9(3)$ & $6.5(8)$ & $3.2(4)$ & $2(1)$ \\
867     Pyramidal $\{2~0~\bar{2}~1\}$ & 8.65 & 1.37 & $35(3)$ & $0.7(1)$ &
868     $5.8(4)$ & $6.1(5)$ & $3.2(2)$ & $2.5(3)$\\
869     Prismatic $\{1~0~\bar{1}~0\}$ & 6.35 & 2.25 & $45(3)$ & $0.75(9)$ &
870     $3.0(2)$ & $3.0(1)$ & $3.6(2)$ & $4(2)$ \\
871     Secondary Prism $\{1~1~\bar{2}~0\}$ & 6.35 & 2.25 & $43(2)$ & $0.69(3)$ &
872     $3.5(1)$ & $3.3(2)$ & $3.2(2)$ & $5(3)$ \\
873 gezelter 4247 \bottomrule
874 gezelter 4243 \end{tabular}
875 gezelter 4258 \begin{flushleft}
876     \footnotemark[1]\footnotesize{Liquid-solid friction coefficients ($\kappa_x$ and
877     $\kappa_y$) are expressed in 10\textsuperscript{-4} amu
878     \AA\textsuperscript{-2} fs\textsuperscript{-1}.} \\
879     \footnotemark[2]\footnotesize{Uncertainties in
880     the last digits are given in parentheses.}
881     \end{flushleft}
882 gezelter 4243 \end{table}
883    
884 gezelter 4258 % Basal $\{0001\}$ & 4.49 & 1.30 & $34.1 \pm 0.9$ &$0.60 \pm 0.07$
885     % & $5.9 \pm 0.3$ & $6.5 \pm 0.8$ & $3.2 \pm 0.4$ & $2 \pm 1$ \\
886     % Pyramidal $\{2~0~\bar{2}~1\}$ & 8.65 & 1.37 & $35 \pm 3$ & $0.7 \pm 0.1$ &
887     % $5.8 \pm 0.4$ & $6.1 \pm 0.5$ & $3.2 \pm 0.2$ & $2.5 \pm 0.3$\\
888     % Prismatic $\{1~0~\bar{1}~0\}$ & 6.35 & 2.25 & $45 \pm 3$ & $0.75 \pm 0.09$ &
889     % $3.0 \pm 0.2$ & $3.0 \pm 0.1$ & $3.6 \pm 0.2$ & $4 \pm 2$ \\
890     % Secondary Prism $\{1~1~\bar{2}~0\}$ & 6.35 & 2.25 & $43 \pm 2$ & $0.69 \pm 0.03$ &
891     % $3.5 \pm 0.1$ & $3.3 \pm 0.2$ & $3.2 \pm 0.2$ & $5 \pm 3$ \\
892    
893    
894 gezelter 4243 \end{document}