ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/iceWater2/iceWater2.tex
Revision: 4231
Committed: Mon Dec 8 15:43:23 2014 UTC (9 years, 9 months ago) by plouden
Content type: application/x-tex
File size: 37951 byte(s)
Log Message:
added images of systems and more paper revisions

File Contents

# Content
1 % ****** Start of file aipsamp.tex ******
2 %
3 % This file is part of the AIP files in the AIP distribution for REVTeX 4.
4 % Version 4.1 of REVTeX, October 2009
5 %
6 % Copyright (c) 2009 American Institute of Physics.
7 %
8 % See the AIP README file for restrictions and more information.
9 %
10 % TeX'ing this file requires that you have AMS-LaTeX 2.0 installed
11 % as well as the rest of the prerequisites for REVTeX 4.1
12 %
13 % It also requires running BibTeX. The commands are as follows:
14 %
15 % 1) latex aipsamp
16 % 2) bibtex aipsamp
17 % 3) latex aipsamp
18 % 4) latex aipsamp
19 %
20 % Use this file as a source of example code for your aip document.
21 % Use the file aiptemplate.tex as a template for your document.
22 \documentclass[%
23 aip,jcp,
24 amsmath,amssymb,
25 preprint,%
26 % reprint,%
27 %author-year,%
28 %author-numerical,%
29 jcp]{revtex4-1}
30
31 \usepackage{graphicx}% Include figure files
32 \usepackage{dcolumn}% Align table columns on decimal point
33 %\usepackage{bm}% bold math
34 \usepackage{times}
35 \usepackage[version=3]{mhchem} % this is a great package for formatting chemical reactions
36 \usepackage{url}
37
38 \begin{document}
39
40 \title{Friction at Water / Ice-I$_\mathrm{h}$ interfaces: Do the
41 Facets of Ice Have Different Hydrophilicity?}
42
43 \author{Patrick B. Louden}
44
45 \author{J. Daniel Gezelter}
46 \email{gezelter@nd.edu.}
47 \affiliation{Department of Chemistry and Biochemistry, University
48 of Notre Dame, Notre Dame, IN 46556}
49
50 \date{\today}
51
52 \begin{abstract}
53 In this follow up paper of the basal and prismatic facets of the
54 Ice-I$_\mathrm{h}$/water interface, we present the
55 pyramidal and secondary prismatic
56 interfaces for both the quiescent and sheared systems. The structural and
57 dynamic interfacial widths for all four crystal facets were found to be in good
58 agreement, and were found to be independent of the shear rate over the shear
59 rates investigated.
60 Decomposition of the molecular orientational time correlation function showed
61 different behavior for the short- and longer-time decay components approaching
62 normal to the interface. Lastly we show through calculation of the interfacial
63 friction coefficient that the basal and pyramidal facets are more
64 hydrophilic than the prismatic and secondary prismatic facets.
65
66 \end{abstract}
67
68 \maketitle
69
70 \section{Introduction}
71
72 % Why are ice and ice/water interfaces important?
73 There is significant interest in the properties of ice/ice and ice/water
74 interfaces in the geophysics community. Most commonly, the results of shearing
75 two ice blocks past one
76 another\cite{Casassa91, Sukhorukov13, Pritchard12, Lishman13} or the shearing
77 of ice through water\cite{Cuffey99, Bell08}. Using molecular dynamics
78 simulations, Samadashvili has recently shown that when two smooth ice slabs
79 slide past one another, a stable liquid-like layer develops between
80 them\cite{Samadashvili13}. To fundamentally understand these processes, a
81 molecular understanding of the ice/water interfaces is needed.
82
83 Investigation of the ice/water interface is also crucial in understanding
84 processes such as nucleation, crystal
85 growth,\cite{Han92, Granasy95, Vanfleet95} and crystal
86 melting\cite{Weber83, Han92, Sakai96, Sakai96B}. Insight gained to these
87 properties can also be applied to biological systems of interest, such as
88 the behavior of the antifreeze protein found in winter
89 flounder,\cite{Wierzbicki07,Chapsky97} and certain terrestial
90 arthropods.\cite{Duman:2001qy,Meister29012013} Elucidating the properties which
91 give rise to these processes through experimental techniques can be expensive,
92 complicated, and sometimes infeasible. However, through the use of molecular
93 dynamics simulations much of the problems of investigating these properties
94 are alleviated.
95
96 Understanding ice/water interfaces inherently begins with the isolated
97 systems. There has been extensive work parameterizing models for liquid water,
98 such as the SPC\cite{Berendsen81}, SPC/E\cite{Berendsen87},
99 TIP4P\cite{Jorgensen85}, TIP4P/2005\cite{Abascal05},
100 ($\dots$), and more recently, models for simulating
101 the solid phases of water, such as the TIP4P/Ice\cite{Abascal05b} model. The
102 melting point of various crystal structures of ice have been calculated for
103 many of these models
104 (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}),
105 and the partial or complete phase diagram for the model has been determined
106 (SPC/E\cite{Baez95,Bryk04b,Sanz04b}, TIP4P\cite{Sanz04,Sanz04b,Koyama04}, TIP5P\cite{Sanz04,Koyama04}).
107 Knowing the behavior and melting point for these models has enabled an initial
108 investigation of ice/water interfaces.
109
110 The Ice-I$_\mathrm{h}$/water quiescent interface has been extensively studied
111 over the past 30 years by theory and experiment. Haymet \emph{et al.} have
112 done significant work characterizing and quantifying the width of these
113 interfaces for the SPC,\cite{Karim90} SPC/E,\cite{Gay02,Bryk02},
114 CF1,\cite{Hayward01,Hayward02} and TIP4P\cite{Karim88} models for water. In
115 recent years, Haymet has focused on investigating the effects cations and
116 anions have on crystal nucleaion and
117 melting.\cite{Bryk04,Smith05,Wilson08,Wilson10} Nada and Furukawa have studied
118 the the basal- and prismatic-water interface width\cite{Nada95}, crystal
119 surface restructuring at temperatures approaching the melting
120 point\cite{Nada00}, and the mechanism of ice growth inhibition by antifreeze
121 proteins\cite{Nada08,Nada11,Nada12}. Nada has developed a six-site water model
122 for ice/water interfaces near the melting point\cite{Nada03}, and studied the
123 dependence of crystal growth shape on applied pressure\cite{Nada11b}. Using
124 this model, Nada and Furukawa have established differential
125 growth rates for the basal, prismatic, and secondary prismatic facets of
126 Ice-I$_\mathrm{h}$ and found their origins due to a reordering of the hydrogen
127 bond network in water near the interface\cite{Nada05}. While the work
128 described so far has mainly focused on bulk water on ice, there is significant
129 interest in thin films of water on ice surfaces as well.
130
131 It is well known that the surface of ice exhibits a premelting layer at
132 temperatures near the melting point, often called a quasi-liquid layer (QLL).
133 Molecular dynamics simulations of the facets of ice-I$_\mathrm{h}$ exposed
134 to vacuum performed by Conde, Vega and Patrykiejew have found QLL widths of
135 approximately 10 \AA\ at 3 K below the melting point\cite{Conde08}.
136 Similarly, Limmer and Chandler have used course grain simulations and
137 statistical field theory to estimated QLL widths at the same temperature to
138 be about 3 nm\cite{Limmer14}.
139 Recently, Sazaki and Furukawa have developed an experimental technique with
140 sufficient spatial and temporal resolution to visulaize and quantitatively
141 analyze QLLs on ice crystals at temperatures near melting\cite{Sazaki10}. They
142 have found the width of the QLLs perpindicular to the surface at -2.2$^{o}$C
143 to be $\mathcal{O}$(\AA). They have also seen the formation of two immiscible
144 QLLs, which displayed different stabilities and dynamics on the crystal
145 surface\cite{Sazaki12}. Knowledge of the hydrophilicities of each
146 of the crystal facets would help further our understanding of the properties
147 and dynamics of the QLLs.
148
149 Presented here is the follow up to our previous paper\cite{Louden13}, in which
150 the basal and prismatic facets of an ice-I$_\mathrm{h}$/water interface were
151 investigated where the ice was sheared relative to the liquid. By using a
152 recently developed velocity shearing and scaling approach to reverse
153 non-equilibrium molecular dynamics (VSS-RNEMD), simultaneous temperature and
154 velocity gradients can be applied to the system, which allows for measurment
155 of friction and thermal transport properties while maintaining a stable
156 interfacial temperature\cite{Kuang12}. Structural analysis and dynamic
157 correlation functions were used to probe the interfacial response to a shear,
158 and the resulting solid/liquid kinetic friction coefficients were reported.
159 In this paper we present the same analysis for the pyramidal and secondary
160 prismatic facets, and show that the differential interfacial friction
161 coefficients for the four facets of ice-I$_\mathrm{h}$ are determined by their
162 relative hydrophilicity by means of dynamics water contact angle simulations.
163
164 \section{Methodology}
165
166 \begin{figure}
167 \includegraphics[width=\linewidth]{Pyr_comic_strip}
168 \caption{\label{fig:pyrComic} The pyramidal interface with a shear
169 rate of 3.8 ms\textsuperscript{-1}. Lower panel: the local tetrahedral order
170 parameter, $q(z)$, (black circles) and the hyperbolic tangent fit (red line).
171 Middle panel: the imposed thermal gradient required to maintain a fixed
172 interfacial temperature. Upper panel: the transverse velocity gradient that
173 develops in response to an imposed momentum flux. The vertical dotted lines
174 indicate the locations of the midpoints of the two interfaces.}
175 \end{figure}
176
177 \begin{figure}
178 \includegraphics[width=\linewidth]{SP_comic_strip}
179 \caption{\label{fig:spComic} The secondary prismatic interface with a shear
180 rate of 3.5 \
181 ms\textsuperscript{-1}. Panel descriptions match those in figure \ref{fig:pyrComic}.}
182 \end{figure}
183
184 \subsection{Pyramidal and secondary prismatic system construction}
185
186 The construction of the pyramidal and secondary prismatic systems follows that
187 of
188 the basal and prismatic systems presented elsewhere\cite{Louden13}, however
189 the ice crystals and water boxes were equilibrated and combined at 50K
190 instead of 225K. The ice / water systems generated were then equilibrated
191 to 225K. The resulting pyramidal system was
192 $37.47 \times 29.50 \times 93.02$ \AA\ with 1216
193 SPC/E\cite{Berendsen87} molecules in the ice slab, and 2203 in the liquid
194 phase. The secondary
195 prismatic system generated was $71.87 \times 31.66 \times 161.55$ \AA\ with
196 3840
197 SPC/E molecules in the ice slab and 8176 molecules in the liquid phase.
198
199 \subsection{Computational details}
200 % Do we need to justify the sims at 225K?
201 % No crystal growth or shrinkage over 2 successive 1 ns NVT simulations for
202 % either the pyramidal or sec. prismatic ice/water systems.
203
204 The computational details performed here were equivalent to those reported
205 in our previous publication\cite{Louden13}. The only changes made to the
206 previously reported procedure were the following. VSS-RNEMD moves were
207 attempted every 2 fs instead of every 50 fs. This was done to minimize
208 the magnitude of each individual VSS-RNEMD perturbation to the system.
209
210 All pyramidal simulations were performed under the canonical (NVT) ensamble
211 except those
212 during which statistics were accumulated for the orientational correlation
213 function, which were performed under the microcanonical (NVE) ensamble. All
214 secondary prismatic
215 simulations were performed under the NVE ensamble.
216
217 \section{Results and discussion}
218 \subsection{Interfacial width}
219 In the literature there is good agreement that between the solid ice and
220 the bulk water, there exists a region of 'slush-like' water molecules.
221 In this region, the water molecules are structurely distinguishable and
222 behave differently than those of the solid ice or the bulk water.
223 The characteristics of this region have been defined by both structural
224 and dynamic properties; and its width has been measured by the change of these
225 properties from their bulk liquid values to those of the solid ice.
226 Examples of these properties include the density, the diffusion constant, and
227 the translational order profile. \cite{Bryk02,Karim90,Gay02,Hayward01,Hayward02,Karim88}
228
229 Since the VSS-RNEMD moves used to impose the thermal and velocity gradients
230 perturb the momenta of the water molecules in
231 the systems, parameters that depend on translational motion may give
232 faulty results. A stuructural parameter will be less effected by the
233 VSS-RNEMD perturbations to the system. Due to this, we have used the
234 local tetrahedral order parameter to quantify the width of the interface,
235 which was originally described by Kumar\cite{Kumar09} and
236 Errington\cite{Errington01}, and used by Bryk and Haymet in a previous study
237 of ice/water interfaces.\cite{Bryk04b}
238
239 The local tetrahedral order parameter, $q(z)$, is given by
240 \begin{equation}
241 q(z) = \int_0^L \sum_{k=1}^{N} \Bigg(1 -\frac{3}{8}\sum_{i=1}^{3}
242 \sum_{j=i+1}^{4} \bigg(\cos\psi_{ikj}+\frac{1}{3}\bigg)^2\Bigg)
243 \delta(z_{k}-z)\mathrm{d}z \Bigg/ N_z
244 \label{eq:qz}
245 \end{equation}
246 where $\psi_{ikj}$ is the angle formed between the oxygen sites of molecules
247 $i$,$k$, and $j$, where the centeral oxygen is located within molecule $k$ and
248 molecules $i$ and $j$ are two of the closest four water molecules
249 around molecule $k$. All four closest neighbors of molecule $k$ are also
250 required to reside within the first peak of the pair distribution function
251 for molecule $k$ (typically $<$ 3.41 \AA\ for water).
252 $N_z = \int\delta(z_k - z) \mathrm{d}z$ is a normalization factor to account
253 for the varying population of molecules within each finite-width bin.
254
255 To determine the width of the interfaces, each of the systems were
256 divided into 100 artificial bins along the
257 $z$-dimension, and the local tetrahedral order parameter, $q(z)$, was
258 time-averaged for each of the bins, resulting in a tetrahedrality profile of
259 the system. These profiles are shown across the $z$-dimension of the systems
260 in panel $a$ of Figures \ref{fig:pyrComic}
261 and \ref{fig:spComic} (black circles). The $q(z)$ function has a range of
262 (0,1), where a larger value indicates a more tetrahedral environment.
263 The $q(z)$ for the bulk liquid was found to be $\approx $ 0.77, while values of
264 $\approx $ 0.92 were more common for the ice. The tetrahedrality profiles were
265 fit using a hyperbolic tangent\cite{Louden13} designed to smoothly fit the
266 bulk to ice
267 transition, while accounting for the thermal influence on the profile by the
268 kinetic energy exchanges of the VSS-RNEMD moves. In panels $b$ and $c$, the
269 resulting thermal and velocity gradients from an imposed kinetic energy and
270 momentum fluxes can be seen. The verticle dotted
271 lines traversing all three panels indicate the midpoints of the interface
272 as determined by the hyperbolic tangent fit of the tetrahedrality profiles.
273
274 From fitting the tetrahedrality profiles for each of the 0.5 nanosecond
275 simulations (panel c of Figures \ref{fig:pyrComic} and \ref{fig:spComic})
276 by eq. 6\cite{Louden13},we find the interfacial width to be
277 3.2 $\pm$ 0.2 and 3.2 $\pm$ 0.2 \AA\ for the control system with no applied
278 momentum flux for both the pyramidal and secondary prismatic systems.
279 Over the range of shear rates investigated,
280 0.6 $\pm$ 0.2 $\mathrm{ms}^{-1} \rightarrow$ 5.6 $\pm$ 0.4 $\mathrm{ms}^{-1}$
281 for the pyramidal system and 0.9 $\pm$ 0.3 $\mathrm{ms}^{-1} \rightarrow$ 5.4
282 $\pm$ 0.1 $\mathrm{ms}^{-1}$ for the secondary prismatic, we found no
283 significant change in the interfacial width. This follows our previous
284 findings of the basal and
285 prismatic systems, in which the interfacial width was invarient of the
286 shear rate of the ice. The interfacial width of the quiescent basal and
287 prismatic systems was found to be 3.2 $\pm$ 0.4 \AA\ and 3.6 $\pm$ 0.2 \AA\
288 respectively, over the range of shear rates investigated, 0.6 $\pm$ 0.3
289 $\mathrm{ms}^{-1} \rightarrow$ 5.3 $\pm$ 0.5 $\mathrm{ms}^{-1}$ for the basal
290 system and 0.9 $\pm$ 0.2 $\mathrm{ms}^{-1} \rightarrow$ 4.5 $\pm$ 0.1
291 $\mathrm{ms}^{-1}$ for the prismatic.
292
293 These results indicate that the surface structure of the exposed ice crystal
294 has little to no effect on how far into the bulk the ice-like structural
295 ordering is. Also, it appears that the interface is not structurally effected
296 by shearing the ice through water.
297
298
299 \subsection{Orientational dynamics}
300 %Should we include the math here?
301 The orientational time correlation function,
302 \begin{equation}\label{C(t)1}
303 C_{2}(t)=\langle P_{2}(\mathbf{u}(0)\cdot \mathbf{u}(t))\rangle,
304 \end{equation}
305 helps indicate the local environment around the water molecules. The function
306 begins with an initial value of unity, and decays to zero as the water molecule
307 loses memory of its former orientation. Observing the rate at which this decay
308 occurs can provide insight to the mechanism and timescales for the relaxation.
309 In eq. \eqref{C(t)1}, $P_{2}$ is the second-order Legendre polynomial, and
310 $\mathbf{u}$ is the bisecting HOH vector. The angle brackets indicate
311 an ensemble average over all the water molecules in a given spatial region.
312
313 To investigate the dynamics of the water molecules across the interface, the
314 systems were divided in the $z$-dimension into bins, each $\approx$ 3 \AA\
315 wide, and eq. \eqref{C(t)1} was computed for each of the bins. A water
316 molecule was allocated to a particular bin if it was initially in the bin
317 at time zero. To compute eq. \eqref{C(t)1}, each 0.5 ns simulation was
318 followed by an additional 200 ps NVE simulation during which the
319 position and orientations of each molecule were recorded every 0.1 ps.
320
321 The data obtained for each bin was then fit to a triexponential decay given by
322 \begin{equation}\label{C(t)_fit}
323 C_{2}(t) \approx a e^{-t/\tau_\mathrm{short}} + b e^{-t/\tau_\mathrm{middle}} +\
324 c
325 e^{-t/\tau_\mathrm{long}} + (1-a-b-c)
326 \end{equation}
327 where $\tau_{short}$ corresponds to the librational motion of the water
328 molecules, $\tau_{middle}$ corresponds to jumps between the breaking and
329 making of hydrogen bonds, and $\tau_{long}$ corresponds to the translational
330 motion of the water molecules. The last term in \eqref{C(t)_fit} accounts
331 for the water molecules trapped in the ice which do not experience any
332 long-time orientational decay.
333
334 In Figures \ref{fig:PyrOrient} and \ref{fig:SPorient} we see the $z$-coordinate
335 profiles for the three decay constants, $\tau_{short}$ (panel a),
336 $\tau_{middle}$ (panel b),
337 and $\tau_{long}$ (panel c) for the pyramidal and secondary prismatic systems
338 respectively. The control experiments (no shear) are shown in black, and
339 an experiment with an imposed momentum flux is shown in red. The vertical
340 dotted line traversing all three panels denotes the midpoint of the
341 interface as determined by the local tetrahedral order parameter fitting.
342 In the liquid regions of both systems, we see that $\tau_{middle}$ and
343 $\tau_{long}$ have approximately consistent values of $3-6$ ps and $30-40$ ps,
344 resepctively, and increase in value as we approach the interface. Conversely,
345 in panel a, we see that $\tau_{short}$ decreases from the liquid value
346 of $72-76$ fs as we approach the interface. We believe this speed up is due to
347 the constrained motion of librations closer to the interface. Both the
348 approximate values for the decays and trends approaching the interface match
349 those reported previously for the basal and prismatic interfaces.
350
351 As done previously, we have attempted to quantify the distance, $d_{pyramidal}$
352 and $d_{secondary prismatic}$, from the
353 interface that the deviations from the bulk liquid values begin. This was done
354 by fitting the orientational decay constant $z$-profiles by
355 \begin{equation}\label{tauFit}
356 \tau(z)\approx\tau_{liquid}+(\tau_{wall}-\tau_{liquid})e^{-(z-z_{wall})/d}
357 \end{equation}
358 where $\tau_{liquid}$ and $\tau_{wall}$ are the liquid and projected wall
359 values of the decay constants, $z_{wall}$ is the location of the interface,
360 and $d$ is the displacement from the interface at which these deviations
361 occur. The values for $d_{pyramidal}$ and $d_{secondary prismatic}$ were
362 determined
363 for each of the decay constants, and then averaged for better statistics
364 ($\tau_{middle}$ was ommitted for secondary prismatic). For the pyramidal
365 system,
366 $d_{pyramidal}$ was found to be 2.7 \AA\ for both the control and the sheared
367 system. We found $d_{secondary prismatic}$ to be slightly larger than
368 $d_{pyramidal}$ for both the control and with an applied shear, with
369 displacements of $4$ \AA\ for the control system and $3$ \AA\ for the
370 experiment with the imposed momentum flux. These values are consistent with
371 those found for the basal ($d_{basal}\approx2.9$ \AA\ ) and prismatic
372 ($d_{prismatic}\approx3.5$ \AA\ ) systems.
373
374 \subsection{Coefficient of friction of the interfaces}
375 While investigating the kinetic coefficient of friction, there was found
376 to be a dependence for $\mu_k$
377 on the temperature of the liquid water in the system. We believe this
378 dependence
379 arrises from the sharp discontinuity of the viscosity for the SPC/E model
380 at temperatures approaching 200 K\cite{Kuang12}. Due to this, we propose
381 a weighting to the interfacial friction coefficient, $\kappa$ by the
382 shear viscosity of the fluid at 225 K. The interfacial friction coefficient
383 relates the shear stress with the relative velocity of the fluid normal to the
384 interface:
385 \begin{equation}\label{Shenyu-13}
386 j_{z}(p_{x}) = \kappa[v_{x}(fluid)-v_{x}(solid)]
387 \end{equation}
388 where $j_{z}(p_{x})$ is the applied momentum flux (shear stress) across $z$
389 in the
390 $x$-dimension, and $v_{x}$(fluid) and $v_{x}$(solid) are the velocities
391 directly adjacent to the interface. The shear viscosity, $\eta(T)$, of the
392 fluid can be determined under a linear response of the momentum
393 gradient to the applied shear stress by
394 \begin{equation}\label{Shenyu-11}
395 j_{z}(p_{x}) = \eta(T) \frac{\partial v_{x}}{\partial z}.
396 \end{equation}
397 Using eqs \eqref{Shenyu-13} and \eqref{Shenyu-11}, we can find the following
398 expression for $\kappa$,
399 \begin{equation}\label{kappa-1}
400 \kappa = \eta(T) \frac{\partial v_{x}}{\partial z}\frac{1}{[v_{x}(fluid)-v_{x}(solid)]}.
401 \end{equation}
402 Here is where we will introduce the weighting term of $\eta(225)/\eta(T)$
403 giving us
404 \begin{equation}\label{kappa-2}
405 \kappa = \frac{\eta(225)}{[v_{x}(fluid)-v_{x}(solid)]}\frac{\partial v_{x}}{\partial z}.
406 \end{equation}
407
408 To obtain the value of $\eta(225)$ for the SPC/E model, a $31.09 \times 29.38
409 \times 124.39$ \AA\ box with 3744 SPC/E liquid water molecules was
410 equilibrated to 225K,
411 and 5 unique shearing experiments were performed. Each experiment was
412 conducted in the NVE and were 5 ns in
413 length. The VSS were attempted every timestep, which was set to 2 fs.
414 For our SPC/E systems, we found $\eta(225)$ to be 0.0148 $\pm$ 0.0007 Pa s,
415 roughly ten times larger than the value found for 280 K SPC/E bulk water by
416 Kuang\cite{Kuang12}.
417
418 The interfacial friction coefficient, $\kappa$, can equivalently be expressed
419 as the ratio of the viscosity of the fluid to the slip length, $\delta$, which
420 is an indication of how 'slippery' the interface is.
421 \begin{equation}\label{kappa-3}
422 \kappa = \frac{\eta}{\delta}
423 \end{equation}
424 In each of the systems, the interfacial temperature was kept fixed to 225K,
425 which ensured the viscosity of the fluid at the
426 interace was approximately the same. Thus, any significant variation in
427 $\kappa$ between
428 the systems indicates differences in the 'slipperiness' of the interfaces.
429 As each of the ice systems are sheared relative to liquid water, the
430 'slipperiness' of the interface can be taken as an indication of how
431 hydrophobic or hydrophilic the interface is. The calculated $\kappa$ values
432 found for the four crystal facets of Ice-I$_\mathrm{h}$ investigated are shown
433 in Table \ref{tab:kappa}. The basal and pyramidal facets were found to have
434 similar values of $\kappa \approx$ 0.0006
435 (amu \AA\textsuperscript{-2} fs\textsuperscript{-1}), while values of
436 $\kappa \approx$ 0.0003 (amu \AA\textsuperscript{-2} fs\textsuperscript{-1})
437 were found for the prismatic and secondary prismatic systems.
438 These results indicate that the basal and pyramidal facets are
439 more hydrophilic than the prismatic and secondary prismatic facets.
440
441 \subsection{Dynamic water contact angle}
442 The hydrophobicity or hydrophilicity of a surface can be described by the
443 extent a droplet of water wets the surface. The contact angle formed between
444 the solid and the liquid, $\theta$, which relates the free energies of the
445 three interfaces involved, is given by Young's equation.
446 \begin{equation}\label{young}
447 \cos\theta = (\gamma_{sv} - \gamma_{sl})/\gamma_{lv}
448 \end{equation}
449 Here $\gamma_{sv}$, $\gamma_{sl}$, and $\gamma_{lv}$ are the free energies
450 of the solid/vapor, solid/liquid, and liquid/vapor interfaces respectively.
451 Large contact angles ($\theta$ $\gg$ 90\textsuperscript{o}) correspond to low
452 wettability and hydrophobic surfaces, while small contact angles
453 ($\theta$ $\ll$ 90\textsuperscript{o}) correspond to high wettability and
454 hydrophilic surfaces. Experimentally, measurements of the contact angle
455 of sessile drops has been used to quantify the extent of wetting on surfaces
456 with thermally selective wetting charactaristics\cite{Tadanaga00,Liu04,Sun04},
457 as well as nano-pillared surfaces with electrically tunable Cassie-Baxter and
458 Wenzel states\cite{Herbertson06,Dhindsa06,Verplanck07,Ahuja08,Manukyan11}.
459 Luzar and coworkers have done significant work modeling these transitions on
460 nano-patterned surfaces\cite{Daub07,Daub10,Daub11,Ritchie12}, and have found
461 the change in contact angle to be due to the external field perturbing the
462 hydrogen bonding of the liquid/vapor interface\cite{Daub07}.
463
464 Here, we will calculate the contact angle of a water droplet as it spreads
465 across each of the four ice I$_\mathrm{h}$ crystal facets in order to
466 determine the surface's relative hydrophilicites. The ice surfaces were
467 oriented so that the desired facet was exposed to the positive z dimension.
468 The sizes and number of molecules in each of the surfaces is given in Table
469 \ref{tab:ice_sheets}. Molecular restraints were applied to the center of mass
470 of the rigid bodies to prevent surface melting, however the molecules were
471 allowed to reorient themselves freely. The water doplet to be placed on the
472 surface contained 2048 SPC/E molecules, which has been found to produce
473 agreement for the Young contact angle extrapolated to an infinite drop
474 size\cite{Daub10}. The surfaces and droplet were equilibrated to 225 K, at
475 which time the droplet was placed 3-5 \AA\ above the surface at 5 unique
476 locations. Each simulation was 5 ns in length and conducted in the NVE
477 ensemble.
478
479
480
481
482 \begin{table}[h]
483 \centering
484 \caption{Phyiscal properties of the basal, prismatic, pyramidal, and secondary prismatic facets of Ice-I$_\mathrm{h}$}
485 \label{tab:kappa}
486 \begin{tabular}{|ccccc|} \hline
487 & \multicolumn{2}{c}{$\kappa_{Drag direction}$ (amu \AA\textsuperscript{-2} fs\textsuperscript{-1})} & & \\
488 Interface & $\kappa_{x}$ & $\kappa_{y}$ & $\theta_{\infty}$ & K_{spread} (ns^{-1}) \\ \hline
489 basal & $0.00059 \pm 0.00003$ & $0.00065 \pm 0.00008$ & $34.1 \pm 0.9$ & $0.60 \pm 0.07$ \\
490 pyramidal & $0.00058 \pm 0.00004$ & $0.00061 \pm 0.00005$ & $35 \pm 3$ & $0.7 \pm 0.1$ \\
491 prismatic & $0.00030 \pm 0.00002$ & $0.00030 \pm 0.00001$ & $46 \pm 4$ & $0.75 \pm 0.09$ \\
492 secondary prismatic & $0.00035 \pm 0.00001$ & $0.00033 \pm 0.00002$ & $42 \pm 2$ & $0.69 \pm 0.03$ \\ \hline
493 \end{tabular}
494 \end{table}
495
496
497
498 \begin{figure}
499 \includegraphics[width=\linewidth]{Pyr-orient}
500 \caption{\label{fig:PyrOrient} The three decay constants of the
501 orientational time correlation function, $C_2(t)$, for water as a function
502 of distance from the center of the ice slab. The vertical dashed line
503 indicates the edge of the pyramidal ice slab determined by the local order
504 tetrahedral parameter. The control (black circles) and sheared (red squares)
505 experiments were fit by a shifted exponential decay (Eq. 9\cite{Louden13})
506 shown by the black and red lines respectively. The upper two panels show that
507 translational and hydrogen bond making and breaking events slow down
508 through the interface while approaching the ice slab. The bottom most panel
509 shows the librational motion of the water molecules speeding up approaching
510 the ice block due to the confined region of space allowed for the molecules
511 to move in.}
512 \end{figure}
513
514 \begin{figure}
515 \includegraphics[width=\linewidth]{SP-orient-less}
516 \caption{\label{fig:SPorient} Decay constants for $C_2(t)$ at the secondary
517 prismatic face. Panel descriptions match those in \ref{fig:PyrOrient}.}
518 \end{figure}
519
520
521
522 \section{Conclusion}
523 We present the results of molecular dynamics simulations of the pyrmaidal
524 and secondary prismatic facets of an SPC/E model of the
525 Ice-I$_\mathrm{h}$/water interface. The ice was sheared through the liquid
526 water while being exposed to a thermal gradient to maintain a stable
527 interface by using the minimally perturbing VSS RNEMD method. In agreement
528 with our previous findings for the basal and prismatic facets, the interfacial
529 width was found to be independent of shear rate as measured by the local
530 tetrahedral order parameter. This width was found to be
531 3.2~$\pm$ 0.2~\AA\ for both the pyramidal and the secondary prismatic systems.
532 These values are in good agreement with our previously calculated interfacial
533 widths for the basal (3.2~$\pm$ 0.4~\AA\ ) and prismatic (3.6~$\pm$ 0.2~\AA\ )
534 systems.
535
536 Orientational dynamics of the Ice-I$_\mathrm{h}$/water interfaces were studied
537 by calculation of the orientational time correlation function at varying
538 displacements normal to the interface. The decays were fit
539 to a tri-exponential decay, where the three decay constants correspond to
540 the librational motion of the molecules driven by the restoring forces of
541 existing hydrogen bonds ($\tau_{short}$ $\mathcal{O}$(10 fs)), jumps between
542 two different hydrogen bonds ($\tau_{middle}$ $\mathcal{O}$(1 ps)), and
543 translational motion of the molecules ($\tau_{long}$ $\mathcal{O}$(100 ps)).
544 $\tau_{short}$ was found to decrease approaching the interface due to the
545 constrained motion of the molecules as the local environment becomes more
546 ice-like. Conversely, the two longer-time decay constants were found to
547 increase at small displacements from the interface. As seen in our previous
548 work on the basal and prismatic facets, there appears to be a dynamic
549 interface width at which deviations from the bulk liquid values occur.
550 We had previously found $d_{basal}$ and $d_{prismatic}$ to be approximately
551 2.8~\AA\ and 3.5~\AA. We found good agreement of these values for the
552 pyramidal and secondary prismatic systems with $d_{pyramidal}$ and
553 $d_{secondary prismatic}$ to be 2.7~\AA\ and 3~\AA. For all four of the
554 facets, no apparent dependence of the dynamic width on the shear rate was
555 found.
556
557 %Paragraph summarizing the Kappa values
558 The interfacial friction coefficient, $\kappa$, was determined for each facet
559 interface. We were able to reach an expression for $\kappa$ as a function of
560 the velocity profile of the system which is scaled by the viscosity of the liquid
561 at 225 K. In doing so, we have obtained an expression for $\kappa$ which is
562 independent of temperature differences of the liquid water at far displacements
563 from the interface. We found the basal and pyramidal facets to have
564 similar $\kappa$ values, of $\kappa \approx$ 0.0006
565 (amu \AA\textsuperscript{-2} fs\textsuperscript{-1}). However, the
566 prismatic and secondary prismatic facets were found to have $\kappa$ values of
567 $\kappa \approx$ 0.0003 (amu \AA\textsuperscript{-2} fs\textsuperscript{-1}).
568 As these ice facets are being sheared relative to liquid water, with the
569 structural and dynamic width of all four
570 interfaces being approximately the same, the difference in the coefficient of
571 friction indicates the hydrophilicity of the crystal facets are not
572 equivalent. Namely, that the basal and pyramidal facets of Ice-I$_\mathrm{h}$
573 are more hydrophilic than the prismatic and secondary prismatic facets.
574
575
576 \begin{acknowledgments}
577 Support for this project was provided by the National
578 Science Foundation under grant CHE-1362211. Computational time was
579 provided by the Center for Research Computing (CRC) at the
580 University of Notre Dame.
581 \end{acknowledgments}
582
583 \newpage
584
585 \bibliography{iceWater}
586
587 There is significant interest in the properties of ice/ice and ice/water
588 interfaces in the geophysics community. Most commonly, the results of shearing
589 two ice blocks past one
590 another\cite{Casassa91, Sukhorukov13, Pritchard12, Lishman13} or the shearing
591 of ice through water\cite{Cuffey99, Bell08}. Using molecular dynamics
592 simulations, Samadashvili has recently shown that when two smooth ice slabs
593 slide past one another, a stable liquid-like layer develops between
594 them\cite{Samadashvili13}. To fundamentally understand these processes, a
595 molecular understanding of the ice/water interfaces is needed.
596
597 Investigation of the ice/water interface is also crucial in understanding
598 processes such as nucleation, crystal
599 growth,\cite{Han92, Granasy95, Vanfleet95} and crystal
600 melting\cite{Weber83, Han92, Sakai96, Sakai96B}. Insight gained to these
601 properties can also be applied to biological systems of interest, such as
602 the behavior of the antifreeze protein found in winter
603 flounder,\cite{Wierzbicki07,Chapsky97} and certain terrestial
604 arthropods.\cite{Duman:2001qy,Meister29012013} Elucidating the properties which
605 give rise to these processes through experimental techniques can be expensive,
606 complicated, and sometimes infeasible. However, through the use of molecular
607 dynamics simulations much of the problems of investigating these properties
608 are alleviated.
609
610 Understanding ice/water interfaces inherently begins with the isolated
611 systems. There has been extensive work parameterizing models for liquid water,
612 such as the SPC\cite{Berendsen81}, SPC/E\cite{Berendsen87},
613 TIP4P\cite{Jorgensen85}, TIP4P/2005\cite{Abascal05},
614 ($\dots$), and more recently, models for simulating
615 the solid phases of water, such as the TIP4P/Ice\cite{Abascal05b} model. The
616 melting point of various crystal structures of ice have been calculated for
617 many of these models
618 (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}),
619 and the partial or complete phase diagram for the model has been determined
620 (SPC/E\cite{Baez95,Bryk04b,Sanz04b}, TIP4P\cite{Sanz04,Sanz04b,Koyama04}, TIP5P\cite{Sanz04,Koyama04}).
621 Knowing the behavior and melting point for these models has enabled an initial
622 investigation of ice/water interfaces.
623
624 The Ice-I$_\mathrm{h}$/water quiescent interface has been extensively studied
625 over the past 30 years by theory and experiment. Haymet \emph{et al.} have
626 done significant work characterizing and quantifying the width of these
627 interfaces for the SPC,\cite{Karim90} SPC/E,\cite{Gay02,Bryk02},
628 CF1,\cite{Hayward01,Hayward02} and TIP4P\cite{Karim88} models for water. In
629 recent years, Haymet has focused on investigating the effects cations and
630 anions have on crystal nucleaion and
631 melting.\cite{Bryk04,Smith05,Wilson08,Wilson10} Nada and Furukawa have studied
632 the the basal- and prismatic-water interface width\cite{Nada95}, crystal
633 surface restructuring at temperatures approaching the melting
634 point\cite{Nada00}, and the mechanism of ice growth inhibition by antifreeze
635 proteins\cite{Nada08,Nada11,Nada12}. Nada has developed a six-site water model
636 for ice/water interfaces near the melting point\cite{Nada03}, and studied the
637 dependence of crystal growth shape on applied pressure\cite{Nada11b}. Using
638 this model, Nada and Furukawa have established differential
639 growth rates for the basal, prismatic, and secondary prismatic facets of
640 Ice-I$_\mathrm{h}$ and found their origins due to a reordering of the hydrogen
641 bond network in water near the interface\cite{Nada05}. While the work
642 described so far has mainly focused on bulk water on ice, there is significant
643 interest in thin films of water on ice surfaces as well.
644
645 It is well known that the surface of ice exhibits a premelting layer at
646 temperatures near the melting point, often called a quasi-liquid layer (QLL).
647 Molecular dynamics simulations of the facets of ice-I$_\mathrm{h}$ exposed
648 to vacuum performed by Conde, Vega and Patrykiejew have found QLL widths of
649 approximately 10 \AA\ at 3 K below the melting point\cite{Conde08}.
650 Similarly, Limmer and Chandler have used course grain simulations and
651 statistical field theory to estimated QLL widths at the same temperature to
652 be about 3 nm\cite{Limmer14}.
653 Recently, Sazaki and Furukawa have developed an experimental technique with
654 sufficient spatial and temporal resolution to visulaize and quantitatively
655 analyze QLLs on ice crystals at temperatures near melting\cite{Sazaki10}. They
656 have found the width of the QLLs perpindicular to the surface at -2.2$^{o}$C
657 to be $\mathcal{O}$(\AA). They have also seen the formation of two immiscible
658 QLLs, which displayed different stabilities and dynamics on the crystal
659 surface\cite{Sazaki12}. Knowledge of the hydrophilicities of each
660 of the crystal facets would help further our understanding of the properties
661 and dynamics of the QLLs.
662
663 Presented here is the follow up to our previous paper\cite{Louden13}, in which
664 the basal and prismatic facets of an ice-I$_\mathrm{h}$/water interface were
665 investigated where the ice was sheared relative to the liquid. By using a
666 recently developed velocity shearing and scaling approach to reverse
667 non-equilibrium molecular dynamics (VSS-RNEMD), simultaneous temperature and
668 velocity gradients can be applied to the system, which allows for measurment
669 of friction and thermal transport properties while maintaining a stable
670 interfacial temperature\cite{Kuang12}. Structural analysis and dynamic
671 correlation functions were used to probe the interfacial response to a shear,
672 and the resulting solid/liquid kinetic friction coefficients were reported.
673 In this paper we present the same analysis for the pyramidal and secondary
674 prismatic facets, and show that the differential interfacial friction
675 coefficients for the four facets of ice-I$_\mathrm{h}$ are determined by their
676 relative hydrophilicity by means of dynamics water contact angle simulations.
677
678 \end{document}
679
680