ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/iceWater2/iceWater3.tex
Revision: 4252
Committed: Mon Dec 15 13:59:35 2014 UTC (9 years, 9 months ago) by plouden
Content type: application/x-tex
File size: 48919 byte(s)
Log Message:
Updated contact angle data.

File Contents

# Content
1 %% 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 \usepackage[round,numbers,sort&compress]{natbib}
15 \usepackage{fixltx2e}
16 \usepackage{booktabs}
17 \usepackage{multirow}
18 \bibpunct{(}{)}{,}{n}{,}{,}
19 \bibliographystyle{pnas2011}
20
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 \title{Friction at water / ice-I\textsubscript{h} interfaces: Do the
36 different facets of ice have different hydrophilicities?}
37
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 %\significancetext{RJSM and ACAC developed the concept of the study. RJSM conducted the analysis, data interpretation and drafted the manuscript. AGB contributed to the development of the statistical methods, data interpretation and drafting of the manuscript.}
49
50 \maketitle
51
52 \begin{article}
53 \begin{abstract}
54 In this paper we present evidence that some of the crystal facets
55 of ice-I$_\mathrm{h}$ posess structural features that can halve
56 the effective hydrophilicity of the ice/water interface. The
57 spreading dynamics of liquid water droplets on ice facets exhibits
58 long-time behavior that differs substantially for the prismatic
59 $\{1~0~\bar{1}~0\}$ and secondary prism $\{1~1~\bar{2}~0\}$ facets
60 when compared with the basal $\{0001\}$ and pyramidal
61 $\{2~0~\bar{2}~1\}$ facets. We also present the results of
62 simulations of solid-liquid friction of the same four crystal
63 facets being drawn through liquid water. Both simulation
64 techniques provide evidence that the two prismatic faces have an
65 effective surface area in contact with the liquid water of
66 approximately half of the total surface area of the crystal. The
67 ice / water interfacial widths for all four crystal facets are
68 similar (using both structural and dynamic measures), and were
69 found to be independent of the shear rate. Additionally,
70 decomposition of orientational time correlation functions show
71 position-dependence for the short- and longer-time decay
72 components close to the interface.
73 \end{abstract}
74
75 \keywords{ice | water | interfaces | hydrophobicity}
76 \abbreviations{QLL, quasi liquid layer; MD, molecular dynamics; RNEMD,
77 reverse non-equilibrium molecular dynamics}
78
79 \dropcap{S}urfaces can be characterized as hydrophobic or hydrophilic
80 based on the strength of the interactions with water. Hydrophobic
81 surfaces do not have strong enough interactions with water to overcome
82 the internal attraction between molecules in the liquid phase, and the
83 degree of hydrophilicity of a surface can be described by the extent a
84 droplet can spread out over the surface. The contact angle formed
85 between the solid and the liquid depends on the free energies of the
86 three interfaces involved, and is given by Young's
87 equation.\cite{Young05}
88 \begin{equation}\label{young}
89 \cos\theta = (\gamma_{sv} - \gamma_{sl})/\gamma_{lv} .
90 \end{equation}
91 Here $\gamma_{sv}$, $\gamma_{sl}$, and $\gamma_{lv}$ are the free
92 energies of the solid/vapor, solid/liquid, and liquid/vapor interfaces
93 respectively. Large contact angles, $\theta > 90^{\circ}$, correspond
94 to hydrophobic surfaces with low wettability, while small contact
95 angles, $\theta < 90^{\circ}$, correspond to hydrophilic surfaces.
96 Experimentally, measurements of the contact angle of sessile drops is
97 often used to quantify the extent of wetting on surfaces with
98 thermally selective wetting
99 characteristics.\cite{Tadanaga00,Liu04,Sun04}
100
101 Nanometer-scale structural features of a solid surface can influence
102 the hydrophilicity to a surprising degree. Small changes in the
103 heights and widths of nano-pillars can change a surface from
104 superhydrophobic, $\theta \ge 150^{\circ}$, to hydrophilic, $\theta
105 \sim 0^{\circ}$.\cite{Koishi09} This is often referred to as the
106 Cassie-Baxter to Wenzel transition. Nano-pillared surfaces with
107 electrically tunable Cassie-Baxter and Wenzel states have also been
108 observed.\cite{Herbertson06,Dhindsa06,Verplanck07,Ahuja08,Manukyan11}
109 Luzar and coworkers have modeled these transitions on nano-patterned
110 surfaces\cite{Daub07,Daub10,Daub11,Ritchie12}, and have found the
111 change in contact angle is due to the field-induced perturbation of
112 hydrogen bonding at the liquid/vapor interface.\cite{Daub07}
113
114 One would expect the interfaces of ice to be highly hydrophilic (and
115 possibly the most hydrophilic of all solid surfaces). In this paper we
116 present evidence that some of the crystal facets of ice-I$_\mathrm{h}$
117 have structural features that can halve the effective hydrophilicity.
118 Our evidence for this comes from molecular dynamics (MD) simulations
119 of the spreading dynamics of liquid droplets on these facets, as well
120 as reverse non-equilibrium molecular dynamics (RNEMD) simulations of
121 solid-liquid friction.
122
123 Quiescent ice-I$_\mathrm{h}$/water interfaces have been studied
124 extensively using computer simulations. Haymet \textit{et al.}
125 characterized and measured the width of these interfaces for the
126 SPC~\cite{Karim90}, SPC/E~\cite{Gay02,Bryk02},
127 CF1~\cite{Hayward01,Hayward02} and TIP4P~\cite{Karim88} models, in
128 both neat water and with solvated
129 ions~\cite{Bryk04,Smith05,Wilson08,Wilson10}. Nada and Furukawa have
130 studied the width of basal/water and prismatic/water
131 interfaces~\cite{Nada95} as well as crystal restructuring at
132 temperatures approaching the melting point~\cite{Nada00}.
133
134 The surface of ice exhibits a premelting layer, often called a
135 quasi-liquid layer (QLL), at temperatures near the melting point. MD
136 simulations of the facets of ice-I$_\mathrm{h}$ exposed to vacuum have
137 found QLL widths of approximately 10 \AA\ at 3 K below the melting
138 point.\cite{Conde08} Similarly, Limmer and Chandler have used the mW
139 water model~\cite{Molinero09} and statistical field theory to estimate
140 QLL widths at similar temperatures to be about 3 nm.\cite{Limmer14}
141
142 Recently, Sazaki and Furukawa have developed a technique using laser
143 confocal microscopy combined with differential interference contrast
144 microscopy that has sufficient spatial and temporal resolution to
145 visulaize and quantitatively analyze QLLs on ice crystals at
146 temperatures near melting.\cite{Sazaki10} They have found the width of
147 the QLLs perpindicular to the surface at -2.2$^{o}$C to be 3-4 \AA\
148 wide. They have also seen the formation of two immiscible QLLs, which
149 displayed different dynamics on the crystal surface.\cite{Sazaki12}
150
151 There is now significant interest in the \textit{tribological}
152 properties of ice/ice and ice/water interfaces in the geophysics
153 community. Understanding the dynamics of solid-solid shearing that is
154 mediated by a liquid layer\cite{Cuffey99, Bell08} will aid in
155 understanding the macroscopic motion of large ice
156 masses.\cite{Casassa91, Sukhorukov13, Pritchard12, Lishman13}
157
158 Using molecular dynamics simulations, Samadashvili has recently shown
159 that when two smooth ice slabs slide past one another, a stable
160 liquid-like layer develops between them.\cite{Samadashvili13} In a
161 previous study, our RNEMD simulations of ice-I$_\mathrm{h}$ shearing
162 through liquid water have provided quantitative estimates of the
163 solid-liquid kinetic friction coefficients.\cite{Louden13} These
164 displayed a factor of two difference between the basal and prismatic
165 facets. The friction was found to be independent of shear direction
166 relative to the surface orientation. We attributed facet-based
167 difference in liquid-solid friction to the 6.5 \AA\ corrugation of the
168 prismatic face which reduces the effective surface area of the ice
169 that is in direct contact with liquid water.
170
171 In the sections that follow, we outline the methodology used to
172 simulate droplet-spreading dynamics using standard MD and tribological
173 properties using RNEMD simulations. These simulation methods give
174 complementary results that point to the prismatic and secondary prism
175 facets having roughly half of their surface area in direct contact
176 with the liquid.
177
178 \section{Methodology}
179 \subsection{Construction of the Ice / Water Interfaces}
180 To construct the four interfacial ice/water systems, a proton-ordered,
181 zero-dipole crystal of ice-I$_\mathrm{h}$ with exposed strips of
182 H-atoms was created using Structure 6 of Hirsch and Ojam\"{a}e's set
183 of orthorhombic representations for ice-I$_{h}$~\cite{Hirsch04}. This
184 crystal structure was cleaved along the four different facets. The
185 exposed face was reoriented normal to the $z$-axis of the simulation
186 cell, and the structures were and extended to form large exposed
187 facets in rectangular box geometries. Liquid water boxes were created
188 with identical dimensions (in $x$ and $y$) as the ice, with a $z$
189 dimension of three times that of the ice block, and a density
190 corresponding to 1 g / cm$^3$. Each of the ice slabs and water boxes
191 were independently equilibrated at a pressure of 1 atm, and the
192 resulting systems were merged by carving out any liquid water
193 molecules within 3 \AA\ of any atoms in the ice slabs. Each of the
194 combined ice/water systems were then equilibrated at 225K, which is
195 the liquid-ice coexistence temperature for SPC/E
196 water~\cite{Bryk02}. Ref. \citealp{Louden13} contains a more detailed
197 explanation of the construction of similar ice/water interfaces. The
198 resulting dimensions as well as the number of ice and liquid water
199 molecules contained in each of these systems are shown in Table
200 \ref{tab:method}.
201
202 The SPC/E water model~\cite{Berendsen87} has been extensively
203 characterized over a wide range of liquid
204 conditions,\cite{Arbuckle02,Kuang12} and its phase diagram has been
205 well studied.\cite{Baez95,Bryk04b,Sanz04b,Fennell:2005fk} With longer
206 cutoff radii and careful treatment of electrostatics, SPC/E mostly
207 avoids metastable crystalline morphologies like
208 ice-\textit{i}~\cite{Fennell:2005fk} and ice-B~\cite{Baez95}. The
209 free energies and melting points
210 \cite{Baez95,Arbuckle02,Gay02,Bryk02,Bryk04b,Sanz04b,Fennell:2005fk,Fernandez06,Abascal07,Vrbka07}
211 of various other crystalline polymorphs have also been calculated.
212 Haymet \textit{et al.} have studied quiescent Ice-I$_\mathrm{h}$/water
213 interfaces using the SPC/E water model, and have seen structural and
214 dynamic measurements of the interfacial width that agree well with
215 more expensive water models, although the coexistence temperature for
216 SPC/E is still well below the experimental melting point of real
217 water~\cite{Bryk02}. Given the extensive data and speed of this model,
218 it is a reasonable choice even though the temperatures required are
219 somewhat lower than real ice / water interfaces.
220
221 \subsection{Droplet Simulations}
222 Ice interfaces with a thickness of $\sim$~20~\AA\ were created as
223 described above, but were not solvated in a liquid box. The crystals
224 were then replicated along the $x$ and $y$ axes (parallel to the
225 surface) until a large surface ($>$ 126 nm\textsuperscript{2}) had
226 been created. The sizes and numbers of molecules in each of the
227 surfaces is given in Table \ref{tab:method}. Weak translational
228 restraining potentials with spring constants of 1.5-4.0~$\mathrm{kcal\
229 mol}^{-1}\mathrm{~\AA}^{-2}$ were applied to the centers of mass of
230 each molecule in order to prevent surface melting, although the
231 molecules were allowed to reorient freely. A water doplet containing
232 2048 SPC/E molecules was created separately. Droplets of this size can
233 produce agreement with the Young contact angle extrapolated to an
234 infinite drop size~\cite{Daub10}. The surfaces and droplet were
235 independently equilibrated to 225 K, at which time the droplet was
236 placed 3-5~\AA\ above the surface. Five statistically independent
237 simulations were carried out for each facet, and the droplet was
238 placed at unique $x$ and $y$ locations for each of these simulations.
239 Each simulation was 5~ns in length and was conducted in the
240 microcanonical (NVE) ensemble. Representative configurations for the
241 droplet on the prismatic facet are shown in figure \ref{fig:Droplet}.
242
243
244 \subsection{Shearing Simulations (Interfaces in Bulk Water)}
245
246 To perform the shearing simulations, the velocity shearing and scaling
247 variant of reverse non-equilibrium molecular dynamics (VSS-RNEMD) was
248 employed \cite{Kuang12}. This method performs a series of simultaneous
249 non-equilibrium exchanges of linear momentum and kinetic energy
250 between two physically-separated regions of the simulation cell. The
251 system responds to this unphysical flux with velocity and temperature
252 gradients. When VSS-RNEMD is applied to bulk liquids, transport
253 properties like the thermal conductivity and the shear viscosity are
254 easily extracted assuming a linear response between the flux and the
255 gradient. At the interfaces between dissimilar materials, the same
256 method can be used to extract \textit{interfacial} transport
257 properties (e.g. the interfacial thermal conductance and the
258 hydrodynamic slip length).
259
260 The kinetic energy flux (producing a thermal gradient) is necessary
261 when performing shearing simulations at the ice-water interface in
262 order to prevent the frictional heating due to the shear from melting
263 the interface. Reference \citealp{Louden13} provides more details on
264 the VSS-RNEMD method as applied to ice-water interfaces. A
265 representative configuration of the solvated prismatic facet being
266 sheared through liquid water is shown in figure \ref{fig:Shearing}.
267
268 In the results discussed below, the exchanges between the two regions
269 were carried out every 2 fs (e.g. every time step). This was done to
270 minimize the magnitude of each individual momentum exchange. Because
271 individual VSS-RNEMD exchanges conserve both total energy and linear
272 momentum, the method can be ``bolted-on'' to simulations in any
273 ensemble. The simulations of the pyramidal interface were performed
274 under the canonical (NVT) ensemble. When time correlation functions
275 were computed (see section \ref{sec:orient}), these simulations were
276 done in the microcanonical (NVE) ensemble. All simulations of the
277 other interfaces were done in the microcanonical ensemble.
278
279 \section{Results}
280 \subsection{Ice - Water Contact Angles}
281
282 To determine the extent of wetting for each of the four crystal
283 facets, contact angles for liquid droplets on the ice surfaces were
284 computed using two methods. In the first method, the droplet is
285 assumed to form a spherical cap, and the contact angle is estimated
286 from the $z$-axis location of the droplet's center of mass
287 ($z_\mathrm{cm}$). This procedure was first described by Hautman and
288 Klein~\cite{Hautman91}, and was utilized by Hirvi and Pakkanen in
289 their investigation of water droplets on polyethylene and poly(vinyl
290 chloride) surfaces~\cite{Hirvi06}. For each stored configuration, the
291 contact angle, $\theta$, was found by inverting the expression for the
292 location of the droplet center of mass,
293 \begin{equation}\label{contact_1}
294 \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} ,
295 \end{equation}
296 where $R_{0}$ is the radius of the free water droplet.
297
298 The second method for obtaining the contact angle was described by
299 Ruijter, Blake, and Coninck~\cite{Ruijter99}. This method uses a
300 cylindrical averaging of the droplet's density profile. A threshold
301 density of 0.5 g cm\textsuperscript{-3} is used to estimate the
302 location of the edge of the droplet. The $r$ and $z$-dependence of
303 the droplet's edge is then fit to a circle, and the contact angle is
304 computed from the intersection of the fit circle with the $z$-axis
305 location of the solid surface. Again, for each stored configuration,
306 the density profile in a set of annular shells was computed. Due to
307 large density fluctuations close to the ice, all shells located within
308 2 \AA\ of the ice surface were left out of the circular fits. The
309 height of the solid surface ($z_\mathrm{suface}$) along with the best
310 fitting central height ($z_\mathrm{center}$) and radius
311 ($r_\mathrm{droplet}$) of the droplet can then be used to compute the
312 contact angle,
313 \begin{equation}
314 \theta = 90 + \frac{180}{\pi} \sin^{-1}\left(\frac{z_\mathrm{center} -
315 z_\mathrm{surface}}{r_\mathrm{droplet}} \right).
316 \end{equation}
317 Both methods provided similar estimates of the dynamic contact angle,
318 although the first method is significantly less prone to noise, and
319 is the method used to report contact angles below.
320
321 Because the initial droplet was placed above the surface, the initial
322 value of 180$^{\circ}$ decayed over time (See figure
323 \ref{fig:ContactAngle}). Each of these profiles were fit to a
324 biexponential decay, with a short-time contribution ($\tau_c$) that
325 describes the initial contact with the surface, a long time
326 contribution ($\tau_s$) that describes the spread of the droplet over
327 the surface, and a constant ($\theta_\infty$) to capture the
328 infinite-time estimate of the equilibrium contact angle,
329 \begin{equation}
330 \theta(t) = \theta_\infty + (180-\theta_\infty) \left[ a e^{-t/\tau_c} +
331 (1-a) e^{-t/\tau_s} \right]
332 \end{equation}
333 We have found that the rate for water droplet spreading across all
334 four crystal facets, $k_\mathrm{spread} = 1/\tau_s \approx$ 0.7
335 ns$^{-1}$. However, the basal and pyramidal facets produced estimated
336 equilibrium contact angles, $\theta_\infty \approx$ 35$^{o}$, while
337 prismatic and secondary prismatic had values for $\theta_\infty$ near
338 43$^{o}$ as seen in Table \ref{tab:kappa}.
339
340 These results indicate that the basal and pyramidal facets are
341 somewhat more hydrophilic than the prismatic and secondary prism
342 facets, and surprisingly, that the differential hydrophilicities of
343 the crystal facets is not reflected in the spreading rate of the
344 droplet.
345
346 % This is in good agreement with our calculations of friction
347 % coefficients, in which the basal
348 % and pyramidal had a higher coefficient of kinetic friction than the
349 % prismatic and secondary prismatic. Due to this, we beleive that the
350 % differences in friction coefficients can be attributed to the varying
351 % hydrophilicities of the facets.
352
353 \subsection{Coefficient of friction of the interfaces}
354 While investigating the kinetic coefficient of friction, there was found
355 to be a dependence for $\mu_k$
356 on the temperature of the liquid water in the system. We believe this
357 dependence
358 arrises from the sharp discontinuity of the viscosity for the SPC/E model
359 at temperatures approaching 200 K\cite{Kuang12}. Due to this, we propose
360 a weighting to the interfacial friction coefficient, $\kappa$ by the
361 shear viscosity of the fluid at 225 K. The interfacial friction coefficient
362 relates the shear stress with the relative velocity of the fluid normal to the
363 interface:
364 \begin{equation}\label{Shenyu-13}
365 j_{z}(p_{x}) = \kappa[v_{x}(fluid)-v_{x}(solid)]
366 \end{equation}
367 where $j_{z}(p_{x})$ is the applied momentum flux (shear stress) across $z$
368 in the
369 $x$-dimension, and $v_{x}$(fluid) and $v_{x}$(solid) are the velocities
370 directly adjacent to the interface. The shear viscosity, $\eta(T)$, of the
371 fluid can be determined under a linear response of the momentum
372 gradient to the applied shear stress by
373 \begin{equation}\label{Shenyu-11}
374 j_{z}(p_{x}) = \eta(T) \frac{\partial v_{x}}{\partial z}.
375 \end{equation}
376 Using eqs \eqref{Shenyu-13} and \eqref{Shenyu-11}, we can find the following
377 expression for $\kappa$,
378 \begin{equation}\label{kappa-1}
379 \kappa = \eta(T) \frac{\partial v_{x}}{\partial z}\frac{1}{[v_{x}(fluid)-v_{x}(solid)]}.
380 \end{equation}
381 Here is where we will introduce the weighting term of $\eta(225)/\eta(T)$
382 giving us
383 \begin{equation}\label{kappa-2}
384 \kappa = \frac{\eta(225)}{[v_{x}(fluid)-v_{x}(solid)]}\frac{\partial v_{x}}{\partial z}.
385 \end{equation}
386
387 To obtain the value of $\eta(225)$ for the SPC/E model, a $31.09 \times 29.38
388 \times 124.39$ \AA\ box with 3744 SPC/E liquid water molecules was
389 equilibrated to 225K,
390 and 5 unique shearing experiments were performed. Each experiment was
391 conducted in the NVE and were 5 ns in
392 length. The VSS were attempted every timestep, which was set to 2 fs.
393 For our SPC/E systems, we found $\eta(225)$ to be 0.0148 $\pm$ 0.0007 Pa s,
394 roughly ten times larger than the value found for 280 K SPC/E bulk water by
395 Kuang\cite{Kuang12}.
396
397 The interfacial friction coefficient, $\kappa$, can equivalently be expressed
398 as the ratio of the viscosity of the fluid to the slip length, $\delta$, which
399 is an indication of how 'slippery' the interface is.
400 \begin{equation}\label{kappa-3}
401 \kappa = \frac{\eta}{\delta}
402 \end{equation}
403 In each of the systems, the interfacial temperature was kept fixed to 225K,
404 which ensured the viscosity of the fluid at the
405 interace was approximately the same. Thus, any significant variation in
406 $\kappa$ between
407 the systems indicates differences in the 'slipperiness' of the interfaces.
408 As each of the ice systems are sheared relative to liquid water, the
409 'slipperiness' of the interface can be taken as an indication of how
410 hydrophobic or hydrophilic the interface is. The calculated $\kappa$ values
411 found for the four crystal facets of Ice-I$_\mathrm{h}$ investigated are shown
412 in Table \ref{tab:kappa}. The basal and pyramidal facets were found to have
413 similar values of $\kappa \approx$ 0.0006
414 (amu \AA\textsuperscript{-2} fs\textsuperscript{-1}), while values of
415 $\kappa \approx$ 0.0003 (amu \AA\textsuperscript{-2} fs\textsuperscript{-1})
416 were found for the prismatic and secondary prismatic systems.
417 These results indicate that the basal and pyramidal facets are
418 more hydrophilic than the prismatic and secondary prismatic facets.
419
420 \subsection{Interfacial width}
421 In the literature there is good agreement that between the solid ice and
422 the bulk water, there exists a region of 'slush-like' water molecules.
423 In this region, the water molecules are structurely distinguishable and
424 behave differently than those of the solid ice or the bulk water.
425 The characteristics of this region have been defined by both structural
426 and dynamic properties; and its width has been measured by the change of these
427 properties from their bulk liquid values to those of the solid ice.
428 Examples of these properties include the density, the diffusion constant, and
429 the translational order profile. \cite{Bryk02,Karim90,Gay02,Hayward01,Hayward02,Karim88}
430
431 Since the VSS-RNEMD moves used to impose the thermal and velocity gradients
432 perturb the momenta of the water molecules in
433 the systems, parameters that depend on translational motion may give
434 faulty results. A stuructural parameter will be less effected by the
435 VSS-RNEMD perturbations to the system. Due to this, we have used the
436 local tetrahedral order parameter (Eq 5\cite{Louden13}) to quantify the width of the interface,
437 which was originally described by Kumar\cite{Kumar09} and
438 Errington\cite{Errington01}, and used by Bryk and Haymet in a previous study
439 of ice/water interfaces.\cite{Bryk04b}
440
441 To determine the width of the interfaces, each of the systems were
442 divided into 100 artificial bins along the
443 $z$-dimension, and the local tetrahedral order parameter, $q(z)$, was
444 time-averaged for each of the bins, resulting in a tetrahedrality profile of
445 the system. These profiles are shown across the $z$-dimension of the systems
446 in panel $a$ of Figures \ref{fig:pyrComic}
447 and \ref{fig:spComic} (black circles). The $q(z)$ function has a range of
448 (0,1), where a larger value indicates a more tetrahedral environment.
449 The $q(z)$ for the bulk liquid was found to be $\approx $ 0.77, while values of
450 $\approx $ 0.92 were more common for the ice. The tetrahedrality profiles were
451 fit using a hyperbolic tangent (Eq. 6\cite{Louden13}) designed to smoothly fit the
452 bulk to ice
453 transition, while accounting for the thermal influence on the profile by the
454 kinetic energy exchanges of the VSS-RNEMD moves. In panels $b$ and $c$, the
455 resulting thermal and velocity gradients from an imposed kinetic energy and
456 momentum fluxes can be seen. The verticle dotted
457 lines traversing all three panels indicate the midpoints of the interface
458 as determined by the hyperbolic tangent fit of the tetrahedrality profiles.
459
460 From fitting the tetrahedrality profiles for each of the 0.5 nanosecond
461 simulations (panel c of Figures \ref{fig:pyrComic} and \ref{fig:spComic})
462 by eq. 6\cite{Louden13},we find the interfacial width to be
463 3.2 $\pm$ 0.2 and 3.2 $\pm$ 0.2 \AA\ for the control system with no applied
464 momentum flux for both the pyramidal and secondary prismatic systems.
465 Over the range of shear rates investigated,
466 0.6 $\pm$ 0.2 $\mathrm{ms}^{-1} \rightarrow$ 5.6 $\pm$ 0.4 $\mathrm{ms}^{-1}$
467 for the pyramidal system and 0.9 $\pm$ 0.3 $\mathrm{ms}^{-1} \rightarrow$ 5.4
468 $\pm$ 0.1 $\mathrm{ms}^{-1}$ for the secondary prismatic, we found no
469 significant change in the interfacial width. This follows our previous
470 findings of the basal and
471 prismatic systems, in which the interfacial width was invarient of the
472 shear rate of the ice. The interfacial width of the quiescent basal and
473 prismatic systems was found to be 3.2 $\pm$ 0.4 \AA\ and 3.6 $\pm$ 0.2 \AA\
474 respectively, over the range of shear rates investigated, 0.6 $\pm$ 0.3
475 $\mathrm{ms}^{-1} \rightarrow$ 5.3 $\pm$ 0.5 $\mathrm{ms}^{-1}$ for the basal
476 system and 0.9 $\pm$ 0.2 $\mathrm{ms}^{-1} \rightarrow$ 4.5 $\pm$ 0.1
477 $\mathrm{ms}^{-1}$ for the prismatic.
478
479 These results indicate that the surface structure of the exposed ice crystal
480 has little to no effect on how far into the bulk the ice-like structural
481 ordering is. Also, it appears that the interface is not structurally effected
482 by the movement of water over the ice.
483
484
485 \subsection{Orientational dynamics \label{sec:orient}}
486 %Should we include the math here?
487 The orientational time correlation function,
488 \begin{equation}\label{C(t)1}
489 C_{2}(t)=\langle P_{2}(\mathbf{u}(0)\cdot \mathbf{u}(t))\rangle,
490 \end{equation}
491 helps indicate the local environment around the water molecules. The function
492 begins with an initial value of unity, and decays to zero as the water molecule
493 loses memory of its former orientation. Observing the rate at which this decay
494 occurs can provide insight to the mechanism and timescales for the relaxation.
495 In eq. \eqref{C(t)1}, $P_{2}$ is the second-order Legendre polynomial, and
496 $\mathbf{u}$ is the bisecting HOH vector. The angle brackets indicate
497 an ensemble average over all the water molecules in a given spatial region.
498
499 To investigate the dynamics of the water molecules across the interface, the
500 systems were divided in the $z$-dimension into bins, each $\approx$ 3 \AA\
501 wide, and eq. \eqref{C(t)1} was computed for each of the bins. A water
502 molecule was allocated to a particular bin if it was initially in the bin
503 at time zero. To compute eq. \eqref{C(t)1}, each 0.5 ns simulation was
504 followed by an additional 200 ps NVE simulation during which the
505 position and orientations of each molecule were recorded every 0.1 ps.
506
507 The data obtained for each bin was then fit to a triexponential decay
508 with the three decay constants
509 $\tau_{short}$ corresponding to the librational motion of the water
510 molecules, $\tau_{middle}$ corresponding to jumps between the breaking and
511 making of hydrogen bonds, and $\tau_{long}$ corresponding to the translational
512 motion of the water molecules. An additive constant in the fit accounts
513 for the water molecules trapped in the ice which do not experience any
514 long-time orientational decay.
515
516 In Figure \ref{fig:PyrOrient} we see the $z$-coordinate profiles for
517 the three decay constants, $\tau_{short}$ (panel a), $\tau_{middle}$
518 (panel b), and $\tau_{long}$ (panel c) for the pyramidal and secondary
519 prismatic systems respectively. The control experiments (no shear) are
520 shown with circles, and an experiment with an imposed momentum flux is
521 shown with squares. The vertical dotted line traversing all three
522 panels denotes the midpoint of the interface determined using the
523 local tetrahedral order parameter. In the liquid regions of both
524 systems, we see that $\tau_{middle}$ and $\tau_{long}$ have
525 approximately consistent values of $3-6$ ps and $30-40$ ps,
526 resepctively, and increase in value as we approach the
527 interface. Conversely, in panel a, we see that $\tau_{short}$
528 decreases from the liquid value of $72-76$ fs as we approach the
529 interface. We believe this speed up is due to the constrained motion
530 of librations closer to the interface. Both the approximate values for
531 the decays and trends approaching the interface match those reported
532 previously for the basal and prismatic interfaces.
533
534 As done previously, we have attempted to quantify the distance, $d_{pyramidal}$
535 and $d_{secondary prismatic}$, from the
536 interface that the deviations from the bulk liquid values begin. This was done
537 by fitting the orientational decay constant $z$-profiles by
538 \begin{equation}\label{tauFit}
539 \tau(z)\approx\tau_{liquid}+(\tau_{wall}-\tau_{liquid})e^{-(z-z_{wall})/d}
540 \end{equation}
541 where $\tau_{liquid}$ and $\tau_{wall}$ are the liquid and projected wall
542 values of the decay constants, $z_{wall}$ is the location of the interface,
543 and $d$ is the displacement from the interface at which these deviations
544 occur. The values for $d_{pyramidal}$ and $d_{secondary prismatic}$ were
545 determined
546 for each of the decay constants, and then averaged for better statistics
547 ($\tau_{middle}$ was ommitted for secondary prismatic). For the pyramidal
548 system,
549 $d_{pyramidal}$ was found to be 2.7 \AA\ for both the control and the sheared
550 system. We found $d_{secondary prismatic}$ to be slightly larger than
551 $d_{pyramidal}$ for both the control and with an applied shear, with
552 displacements of $4$ \AA\ for the control system and $3$ \AA\ for the
553 experiment with the imposed momentum flux. These values are consistent with
554 those found for the basal ($d_{basal}\approx2.9$ \AA\ ) and prismatic
555 ($d_{prismatic}\approx3.5$ \AA\ ) systems.
556
557
558
559
560
561 \section{Conclusion}
562 We present the results of molecular dynamics simulations of the basal,
563 prismatic, pyrmaidal
564 and secondary prismatic facets of an SPC/E model of the
565 Ice-I$_\mathrm{h}$/water interface, and show that the differential
566 coefficients of friction among the four facets are due to their
567 relative hydrophilicities by means
568 of water contact angle calculations. To obtain the coefficients of
569 friction, the ice was sheared through the liquid
570 water while being exposed to a thermal gradient to maintain a stable
571 interface by using the minimally perturbing VSS RNEMD method. Water
572 contact angles are obtained by fitting the spreading of a liquid water
573 droplet over the crystal facets.
574
575 In agreement with our previous findings for the basal and prismatic facets, the interfacial
576 width of the prismatic and secondary prismatic crystal faces were
577 found to be independent of shear rate as measured by the local
578 tetrahedral order parameter. This width was found to be
579 3.2~$\pm$ 0.2~\AA\ for both the pyramidal and the secondary prismatic systems.
580 These values are in good agreement with our previously calculated interfacial
581 widths for the basal (3.2~$\pm$ 0.4~\AA\ ) and prismatic (3.6~$\pm$ 0.2~\AA\ )
582 systems.
583
584 Orientational dynamics of the Ice-I$_\mathrm{h}$/water interfaces were studied
585 by calculation of the orientational time correlation function at varying
586 displacements normal to the interface. The decays were fit
587 to a tri-exponential decay, where the three decay constants correspond to
588 the librational motion of the molecules driven by the restoring forces of
589 existing hydrogen bonds ($\tau_{short}$ $\mathcal{O}$(10 fs)), jumps between
590 two different hydrogen bonds ($\tau_{middle}$ $\mathcal{O}$(1 ps)), and
591 translational motion of the molecules ($\tau_{long}$ $\mathcal{O}$(100 ps)).
592 $\tau_{short}$ was found to decrease approaching the interface due to the
593 constrained motion of the molecules as the local environment becomes more
594 ice-like. Conversely, the two longer-time decay constants were found to
595 increase at small displacements from the interface. As seen in our previous
596 work on the basal and prismatic facets, there appears to be a dynamic
597 interface width at which deviations from the bulk liquid values occur.
598 We had previously found $d_{basal}$ and $d_{prismatic}$ to be approximately
599 2.8~\AA\ and 3.5~\AA. We found good agreement of these values for the
600 pyramidal and secondary prismatic systems with $d_{pyramidal}$ and
601 $d_{secondary prismatic}$ to be 2.7~\AA\ and 3~\AA. For all four of the
602 facets, no apparent dependence of the dynamic width on the shear rate was
603 found.
604
605 The interfacial friction coefficient, $\kappa$, was determined for each facet
606 interface. We were able to reach an expression for $\kappa$ as a function of
607 the velocity profile of the system which is scaled by the viscosity of the liquid
608 at 225 K. In doing so, we have obtained an expression for $\kappa$ which is
609 independent of temperature differences of the liquid water at far displacements
610 from the interface. We found the basal and pyramidal facets to have
611 similar $\kappa$ values, of $\kappa \approx$ 6
612 (x$10^{-4}$amu \AA\textsuperscript{-2} fs\textsuperscript{-1}). However, the
613 prismatic and secondary prismatic facets were found to have $\kappa$ values of
614 $\kappa \approx$ 3 (x$10^{-4}$amu \AA\textsuperscript{-2} fs\textsuperscript{-1}).
615 Believing this difference was due to the relative hydrophilicities of
616 the crystal faces, we have calculated the infinite decay of the water
617 contact angle, $\theta_{\infty}$, by watching the spreading of a water
618 droplet over the surface of the crystal facets. We have found
619 $\theta_{\infty}$ for the basal and pyramidal faces to be $\approx$ 34
620 degrees, while obtaining $\theta_{\infty}$ of $\approx$ 43 degrees for
621 the prismatic and secondary prismatic faces. This indicates that the
622 basal and pyramidal faces of ice-I$_\mathrm{h}$ are more hydrophilic
623 than the prismatic and secondary prismatic. These results also seem to
624 explain the differential friction coefficients obtained through the
625 shearing simulations, namely, that the coefficients of friction of the
626 ice-I$_\mathrm{h}$ crystal facets are governed by their inherent
627 hydrophilicities.
628
629
630 \begin{acknowledgments}
631 Support for this project was provided by the National
632 Science Foundation under grant CHE-1362211. Computational time was
633 provided by the Center for Research Computing (CRC) at the
634 University of Notre Dame.
635 \end{acknowledgments}
636
637 \bibliography{iceWater}
638 % *****************************************
639 % There is significant interest in the properties of ice/ice and ice/water
640 % interfaces in the geophysics community. Most commonly, the results of shearing
641 % two ice blocks past one
642 % another\cite{Casassa91, Sukhorukov13, Pritchard12, Lishman13} or the shearing
643 % of ice through water\cite{Cuffey99, Bell08}. Using molecular dynamics
644 % simulations, Samadashvili has recently shown that when two smooth ice slabs
645 % slide past one another, a stable liquid-like layer develops between
646 % them\cite{Samadashvili13}. To fundamentally understand these processes, a
647 % molecular understanding of the ice/water interfaces is needed.
648
649 % Investigation of the ice/water interface is also crucial in understanding
650 % processes such as nucleation, crystal
651 % growth,\cite{Han92, Granasy95, Vanfleet95} and crystal
652 % melting\cite{Weber83, Han92, Sakai96, Sakai96B}. Insight gained to these
653 % properties can also be applied to biological systems of interest, such as
654 % the behavior of the antifreeze protein found in winter
655 % flounder,\cite{Wierzbicki07,Chapsky97} and certain terrestial
656 % arthropods.\cite{Duman:2001qy,Meister29012013} Elucidating the properties which
657 % give rise to these processes through experimental techniques can be expensive,
658 % complicated, and sometimes infeasible. However, through the use of molecular
659 % dynamics simulations much of the problems of investigating these properties
660 % are alleviated.
661
662 % Understanding ice/water interfaces inherently begins with the isolated
663 % systems. There has been extensive work parameterizing models for liquid water,
664 % such as the SPC\cite{Berendsen81}, SPC/E\cite{Berendsen87},
665 % TIP4P\cite{Jorgensen85}, TIP4P/2005\cite{Abascal05},
666 % ($\dots$), and more recently, models for simulating
667 % the solid phases of water, such as the TIP4P/Ice\cite{Abascal05b} model. The
668 % melting point of various crystal structures of ice have been calculated for
669 % many of these models
670 % (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}),
671 % and the partial or complete phase diagram for the model has been determined
672 % (SPC/E\cite{Baez95,Bryk04b,Sanz04b}, TIP4P\cite{Sanz04,Sanz04b,Koyama04}, TIP5P\cite{Sanz04,Koyama04}).
673 % Knowing the behavior and melting point for these models has enabled an initial
674 % investigation of ice/water interfaces.
675
676 % The Ice-I$_\mathrm{h}$/water quiescent interface has been extensively studied
677 % over the past 30 years by theory and experiment. Haymet \emph{et al.} have
678 % done significant work characterizing and quantifying the width of these
679 % interfaces for the SPC,\cite{Karim90} SPC/E,\cite{Gay02,Bryk02},
680 % CF1,\cite{Hayward01,Hayward02} and TIP4P\cite{Karim88} models for water. In
681 % recent years, Haymet has focused on investigating the effects cations and
682 % anions have on crystal nucleaion and
683 % melting.\cite{Bryk04,Smith05,Wilson08,Wilson10} Nada and Furukawa have studied
684 % the the basal- and prismatic-water interface width\cite{Nada95}, crystal
685 % surface restructuring at temperatures approaching the melting
686 % point\cite{Nada00}, and the mechanism of ice growth inhibition by antifreeze
687 % proteins\cite{Nada08,Nada11,Nada12}. Nada has developed a six-site water model
688 % for ice/water interfaces near the melting point\cite{Nada03}, and studied the
689 % dependence of crystal growth shape on applied pressure\cite{Nada11b}. Using
690 % this model, Nada and Furukawa have established differential
691 % growth rates for the basal, prismatic, and secondary prismatic facets of
692 % Ice-I$_\mathrm{h}$ and found their origins due to a reordering of the hydrogen
693 % bond network in water near the interface\cite{Nada05}. While the work
694 % described so far has mainly focused on bulk water on ice, there is significant
695 % interest in thin films of water on ice surfaces as well.
696
697 % It is well known that the surface of ice exhibits a premelting layer at
698 % temperatures near the melting point, often called a quasi-liquid layer (QLL).
699 % Molecular dynamics simulations of the facets of ice-I$_\mathrm{h}$ exposed
700 % to vacuum performed by Conde, Vega and Patrykiejew have found QLL widths of
701 % approximately 10 \AA\ at 3 K below the melting point\cite{Conde08}.
702 % Similarly, Limmer and Chandler have used course grain simulations and
703 % statistical field theory to estimated QLL widths at the same temperature to
704 % be about 3 nm\cite{Limmer14}.
705 % Recently, Sazaki and Furukawa have developed an experimental technique with
706 % sufficient spatial and temporal resolution to visulaize and quantitatively
707 % analyze QLLs on ice crystals at temperatures near melting\cite{Sazaki10}. They
708 % have found the width of the QLLs perpindicular to the surface at -2.2$^{o}$C
709 % to be $\mathcal{O}$(\AA). They have also seen the formation of two immiscible
710 % QLLs, which displayed different stabilities and dynamics on the crystal
711 % surface\cite{Sazaki12}. Knowledge of the hydrophilicities of each
712 % of the crystal facets would help further our understanding of the properties
713 % and dynamics of the QLLs.
714
715 % Presented here is the follow up to our previous paper\cite{Louden13}, in which
716 % the basal and prismatic facets of an ice-I$_\mathrm{h}$/water interface were
717 % investigated where the ice was sheared relative to the liquid. By using a
718 % recently developed velocity shearing and scaling approach to reverse
719 % non-equilibrium molecular dynamics (VSS-RNEMD), simultaneous temperature and
720 % velocity gradients can be applied to the system, which allows for measurment
721 % of friction and thermal transport properties while maintaining a stable
722 % interfacial temperature\cite{Kuang12}. Structural analysis and dynamic
723 % correlation functions were used to probe the interfacial response to a shear,
724 % and the resulting solid/liquid kinetic friction coefficients were reported.
725 % In this paper we present the same analysis for the pyramidal and secondary
726 % prismatic facets, and show that the differential interfacial friction
727 % coefficients for the four facets of ice-I$_\mathrm{h}$ are determined by their
728 % relative hydrophilicity by means of dynamics water contact angle
729 % simulations.
730
731 % The local tetrahedral order parameter, $q(z)$, is given by
732 % \begin{equation}
733 % q(z) = \int_0^L \sum_{k=1}^{N} \Bigg(1 -\frac{3}{8}\sum_{i=1}^{3}
734 % \sum_{j=i+1}^{4} \bigg(\cos\psi_{ikj}+\frac{1}{3}\bigg)^2\Bigg)
735 % \delta(z_{k}-z)\mathrm{d}z \Bigg/ N_z
736 % \label{eq:qz}
737 % \end{equation}
738 % where $\psi_{ikj}$ is the angle formed between the oxygen sites of molecules
739 % $i$,$k$, and $j$, where the centeral oxygen is located within molecule $k$ and
740 % molecules $i$ and $j$ are two of the closest four water molecules
741 % around molecule $k$. All four closest neighbors of molecule $k$ are also
742 % required to reside within the first peak of the pair distribution function
743 % for molecule $k$ (typically $<$ 3.41 \AA\ for water).
744 % $N_z = \int\delta(z_k - z) \mathrm{d}z$ is a normalization factor to account
745 % for the varying population of molecules within each finite-width bin.
746
747
748 % The hydrophobicity or hydrophilicity of a surface can be described by the
749 % extent a droplet of water wets the surface. The contact angle formed between
750 % the solid and the liquid, $\theta$, which relates the free energies of the
751 % three interfaces involved, is given by Young's equation.
752 % \begin{equation}\label{young}
753 % \cos\theta = (\gamma_{sv} - \gamma_{sl})/\gamma_{lv}
754 % \end{equation}
755 % Here $\gamma_{sv}$, $\gamma_{sl}$, and $\gamma_{lv}$ are the free energies
756 % of the solid/vapor, solid/liquid, and liquid/vapor interfaces respectively.
757 % Large contact angles ($\theta$ $\gg$ 90\textsuperscript{o}) correspond to low
758 % wettability and hydrophobic surfaces, while small contact angles
759 % ($\theta$ $\ll$ 90\textsuperscript{o}) correspond to high wettability and
760 % hydrophilic surfaces. Experimentally, measurements of the contact angle
761 % of sessile drops has been used to quantify the extent of wetting on surfaces
762 % with thermally selective wetting charactaristics\cite{Tadanaga00,Liu04,Sun04},
763 % as well as nano-pillared surfaces with electrically tunable Cassie-Baxter and
764 % Wenzel states\cite{Herbertson06,Dhindsa06,Verplanck07,Ahuja08,Manukyan11}.
765 % Luzar and coworkers have done significant work modeling these transitions on
766 % nano-patterned surfaces\cite{Daub07,Daub10,Daub11,Ritchie12}, and have found
767 % the change in contact angle to be due to the external field perturbing the
768 % hydrogen bonding of the liquid/vapor interface\cite{Daub07}.
769
770
771
772 \end{article}
773
774 \begin{figure}
775 \includegraphics[width=\linewidth]{Droplet}
776 \caption{\label{fig:Droplet} Computational model of a droplet of
777 liquid water spreading over the prismatic $\{1~0~\bar{1}~0\}$ facet
778 of ice, before (left) and 2.6 ns after (right) being introduced to the
779 surface. The contact angle ($\theta$) shrinks as the simulation
780 proceeds, and the long-time behavior of this angle is used to
781 estimate the hydrophilicity of the facet.}
782 \end{figure}
783
784 \begin{figure}
785 \includegraphics[width=2in]{Shearing}
786 \caption{\label{fig:Shearing} Computational model of a slab of ice
787 being sheared through liquid water. In this figure, the ice is
788 presenting two copies of the prismatic $\{1~0~\bar{1}~0\}$ facet
789 towards the liquid phase. The RNEMD simulation exchanges both
790 linear momentum (indicated with arrows) and kinetic energy between
791 the central box and the box that spans the cell boundary. The
792 system responds with weak thermal gradient and a velocity profile
793 that shears the ice relative to the surrounding liquid.}
794 \end{figure}
795
796 \begin{figure}
797 \includegraphics[width=\linewidth]{ContactAngle}
798 \caption{\label{fig:ContactAngle} The dynamic contact angle of a
799 droplet after approaching each of the four ice facets. The decay to
800 an equilibrium contact angle displays similar dynamics. Although
801 all the surfaces are hydrophilic, the long-time behavior stabilizes
802 to significantly flatter droplets for the basal and pyramidal
803 facets. This suggests a difference in hydrophilicity for these
804 facets compared with the two prismatic facets.}
805 \end{figure}
806
807 \begin{figure}
808 \includegraphics[width=\linewidth]{Pyr_comic_strip}
809 \caption{\label{fig:pyrComic} Properties of the pyramidal interface
810 being sheared through water at 3.8 ms\textsuperscript{-1}. Lower
811 panel: the local tetrahedral order parameter, $q(z)$, (circles) and
812 the hyperbolic tangent fit (turquoise line). Middle panel: the
813 imposed thermal gradient required to maintain a fixed interfacial
814 temperature of 225 K. Upper panel: the transverse velocity gradient
815 that develops in response to an imposed momentum flux. The vertical
816 dotted lines indicate the locations of the midpoints of the two
817 interfaces.}
818 \end{figure}
819
820 % \begin{figure}
821 % \includegraphics[width=\linewidth]{SP_comic_strip}
822 % \caption{\label{fig:spComic} The secondary prismatic interface with a shear
823 % rate of 3.5 \
824 % ms\textsuperscript{-1}. Panel descriptions match those in figure \ref{fig:pyrComic}.}
825 % \end{figure}
826
827 \begin{figure}
828 \includegraphics[width=\linewidth]{Pyr-orient}
829 \caption{\label{fig:PyrOrient} The three decay constants of the
830 orientational time correlation function, $C_2(t)$, for water as a
831 function of distance from the center of the ice slab. The vertical
832 dashed line indicates the edge of the pyramidal ice slab determined
833 by the local order tetrahedral parameter. The control (circles) and
834 sheared (squares) simulations were fit using shifted-exponential
835 decay (see Eq. 9 in Ref. \citealp{Louden13}).}
836 \end{figure}
837
838 % \begin{figure}
839 % \includegraphics[width=\linewidth]{SP-orient-less}
840 % \caption{\label{fig:SPorient} Decay constants for $C_2(t)$ at the secondary
841 % prismatic face. Panel descriptions match those in \ref{fig:PyrOrient}.}
842 % \end{figure}
843
844
845 \begin{table}[h]
846 \centering
847 \caption{Sizes of the droplet and shearing simulations. Cell
848 dimensions are measured in \AA. \label{tab:method}}
849 \begin{tabular}{r|cccc|ccccc}
850 \toprule
851 \multirow{2}{*}{Interface} & \multicolumn{4}{c|}{Droplet} & \multicolumn{5}{c}{Shearing} \\
852 & $N_\mathrm{ice}$ & $N_\mathrm{droplet}$ & $L_x$ & $L_y$ & $N_\mathrm{ice}$ & $N_\mathrm{liquid}$ & $L_x$ & $L_y$ & $L_z$ \\
853 \midrule
854 Basal $\{0001\}$ & 12960 & 2048 & 134.70 & 140.04 & 900 & 1846 & 23.87 & 35.83 & 98.64 \\
855 Pyramidal $\{2~0~\bar{2}~1\}$ & 11136 & 2048 & 143.75 & 121.41 & 1216 & 2203 & 37.47 & 29.50 & 93.02 \\
856 Prismatic $\{1~0~\bar{1}~0\}$ & 9900 & 2048 & 110.04 & 115.00 & 3000 & 5464 & 35.95 & 35.65 & 205.77 \\
857 Secondary Prism $\{1~1~\bar{2}~0\}$ & 11520 & 2048 & 146.72 & 124.48 & 3840 & 8176 & 71.87 & 31.66 & 161.55 \\
858 \bottomrule
859 \end{tabular}
860 \end{table}
861
862
863 \begin{table}[h]
864 \centering
865 \caption{Structural and dynamic properties of the interfaces of Ice-I$_\mathrm{h}$
866 with water. Liquid-solid friction coefficients ($\kappa_x$ and $\kappa_y$) are expressed in 10\textsuperscript{-4} amu
867 \AA\textsuperscript{-2} fs\textsuperscript{-1}. \label{tab:kappa}}
868 \begin{tabular}{r|cc|cccc}
869 \toprule
870 \multirow{2}{*}{Interface} & \multicolumn{2}{c|}{Droplet} & \multicolumn{4}{c}{Shearing}\\
871 & $\theta_{\infty}$ ($^\circ$) & $k_\mathrm{spread}$ (ns\textsuperscript{-1}) &
872 $\kappa_{x}$ & $\kappa_{y}$ & $d_{q_{z}}$ (\AA) & $d_{\tau}$ (\AA) \\
873 \midrule
874 Basal $\{0001\}$ & $34.1 \pm 0.9$ &$0.60 \pm 0.07$
875 & $5.9 \pm 0.3$ & $6.5 \pm 0.8$ & $3.2 \pm 0.4$ & $2.6 \pm 0.8$ \\
876 Pyramidal $\{2~0~\bar{2}~1\}$ & $35 \pm 3$ & $0.7 \pm 0.1$ &
877 $5.8 \pm 0.4$ & $6.1 \pm 0.5$ & $3.2 \pm 0.2$ & $2.7 \pm 0.3$\\
878 Prismatic $\{1~0~\bar{1}~0\}$ & $45 \pm 3$ & $0.75 \pm 0.09$ &
879 $3.0 \pm 0.2$ & $3.0 \pm 0.1$ & $3.6 \pm 0.2$ & $4 \pm 2$ \\
880 Secondary Prism $\{1~1~\bar{2}~0\}$ & $43 \pm 2$ & $0.69 \pm 0.03$ &
881 $3.5 \pm 0.1$ & $3.3 \pm 0.2$ & $3.2 \pm 0.2$ & $3.4 \pm 0.5$ \\
882 \bottomrule
883 \end{tabular}
884 \end{table}
885
886 \end{document}