ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/iceWater2/iceWater3.tex
Revision: 4236
Committed: Mon Dec 8 19:12:33 2014 UTC (9 years, 9 months ago) by plouden
Content type: application/x-tex
File size: 41631 byte(s)
Log Message:
Revised - Intro.

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
14
15
16 %% OPTIONAL MACRO DEFINITIONS
17 \def\s{\sigma}
18 %%%%%%%%%%%%
19 %% For PNAS Only:
20 \url{www.pnas.org/cgi/doi/10.1073/pnas.0709640104}
21 \copyrightyear{2008}
22 \issuedate{Issue Date}
23 \volume{Volume}
24 \issuenumber{Issue Number}
25 %\setcounter{page}{2687} %Set page number here if desired
26 %%%%%%%%%%%%
27
28 \begin{document}
29
30 \title{Friction at Water / Ice-I$_\mathrm{h}$ interfaces: Do the
31 Different Facets of Ice Have Different Hydrophilicity?}
32
33 \author{Patrick B. Louden
34 \and
35 J. Daniel Gezelter\affil{1}{Department of Chemistry and Biochemistry, University of Notre Dame, Notre Dame,
36 IN 46556}}
37
38 \contributor{Submitted to Proceedings of the National Academy of Sciences
39 of the United States of America}
40
41 %%%Newly updated.
42 %%% If significance statement need, then can use the below command otherwise just delete it.
43 \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.}
44
45 \maketitle
46
47 \begin{article}
48 \begin{abstract}
49 {In this follow up paper of the basal and prismatic facets of the
50 Ice-I$_\mathrm{h}$/water interface, we present the
51 pyramidal and secondary prismatic
52 interfaces for both the quiescent and sheared systems. The structural and
53 dynamic interfacial widths for all four crystal facets were found to be in good
54 agreement, and were found to be independent of the shear rate over the shear
55 rates investigated.
56 Decomposition of the molecular orientational time correlation function showed
57 different behavior for the short- and longer-time decay components approaching
58 normal to the interface. Lastly we show through calculation of the interfacial
59 friction coefficient and dynamic water contact angle measurement
60 that the basal and pyramidal facets are more
61 hydrophilic than the prismatic and secondary prismatic facets.}
62 \end{abstract}
63
64 \keywords{ice|water|interface|contact angle|molecular dynamics}
65
66 %\abbreviations{QLL, quasi liquid layer; MD, molecular dynamics}
67
68 \dropcap{T}he Ice-I$_\mathrm{h}$/water quiescent interface has been extensively studied
69 over the past 30 years by theory and experiment. Haymet \emph{et al.} have
70 done significant work characterizing and quantifying the width of these
71 interfaces for the SPC,\cite{Karim90} SPC/E,\cite{Gay02,Bryk02},
72 CF1,\cite{Hayward01,Hayward02} and TIP4P\cite{Karim88} models for water.\cite{Bryk04,Smith05,Wilson08,Wilson10} Nada and Furukawa have studied
73 the the basal- and prismatic-water interface width\cite{Nada95} and crystal
74 surface restructuring at temperatures approaching the melting
75 point\cite{Nada00}.
76
77 It is well known that the surface of ice exhibits a premelting layer at
78 temperatures near the melting point, often called a quasi-liquid layer (QLL).
79 Molecular dynamics simulations of the facets of ice-I$_\mathrm{h}$ exposed
80 to vacuum performed by Conde, Vega and Patrykiejew have found QLL widths of
81 approximately 10 \AA\ at 3 K below the melting point\cite{Conde08}.
82 Similarly, Limmer and Chandler have used course grain simulations and
83 statistical field theory to estimated QLL widths at the same temperature to
84 be about 3 nm\cite{Limmer14}.
85 Recently, Sazaki and Furukawa have developed an experimental technique with
86 sufficient spatial and temporal resolution to visulaize and quantitatively
87 analyze QLLs on ice crystals at temperatures near melting\cite{Sazaki10}. They
88 have found the width of the QLLs perpindicular to the surface at -2.2$^{o}$C
89 to be 3-4 \AA\ wide. They have also seen the formation of two immiscible
90 QLLs, which displayed different stabilities and dynamics on the crystal
91 surface\cite{Sazaki12}.
92
93 There is significant interest in the properties of ice/ice and ice/water
94 interfaces in the geophysics community. Most commonly, the results of shearing
95 two ice blocks past one
96 another\cite{Casassa91, Sukhorukov13, Pritchard12, Lishman13} or the shearing
97 of ice through water\cite{Cuffey99, Bell08}. Using molecular dynamics
98 simulations, Samadashvili has recently shown that when two smooth ice slabs
99 slide past one another, a stable liquid-like layer develops between
100 them\cite{Samadashvili13}. To fundamentally understand these processes, a
101 molecular understanding of the ice/water interfaces is needed.
102 Investigation of the ice/water interface is also crucial in understanding
103 processes such as nucleation, crystal
104 growth,\cite{Han92, Granasy95, Vanfleet95} and crystal
105 melting\cite{Weber83, Han92, Sakai96, Sakai96B}.
106
107 In a previous study\cite{Louden13}, we investigated
108 the basal and prismatic facets of an ice-I$_\mathrm{h}$/water
109 interface where the ice was sheared relative to the liquid. Using
110 velocity shearing and scaling approach to reverse
111 non-equilibrium molecular dynamics (VSS-RNEMD), simultaneous temperature and
112 velocity gradients were applied to the system, allowing for measurment
113 of friction and thermal transport properties while maintaining a stable
114 interfacial temperature\cite{Kuang12}. The resulting solid/liquid kinetic friction coefficients were
115 reported, and displayed a factor of two difference between the
116 basal and prismatic facets. We beleived this was due to an effective
117 difference in the contact of the water with the different facets.
118
119 Surfaces exhibit varying interactions with water, and are
120 charactarized as either being hydrophobic or hydrophilic based on the
121 extent of these interactions. Hydrophobic surfaces have unfavorable
122 solid-liquid interactions, and result in water maintaining a spherical
123 or droplet shape on the surface. This occurs due to the liquid-liquid
124 intermolecular forces being stronger than the solid-liquid
125 interaction. Conversely, hydrophilic surfaces display large water
126 spreading over the surface. Here solid-liquid interactions are
127 stronger than the liquid-liquid interactions, which results in the
128 water spreading out over the surface.
129
130 The hydrophobicity or hydrophilicity of a surface can be described by the
131 extent a droplet of water wets the surface. The contact angle formed between
132 the solid and the liquid, $\theta$, which relates the free energies of the
133 three interfaces involved, is given by Young's equation.
134 \begin{equation}\label{young}
135 \cos\theta = (\gamma_{sv} - \gamma_{sl})/\gamma_{lv}
136 \end{equation}
137 Here $\gamma_{sv}$, $\gamma_{sl}$, and $\gamma_{lv}$ are the free energies
138 of the solid/vapor, solid/liquid, and liquid/vapor interfaces respectively.
139 Large contact angles ($\theta$ $\gg$ 90\textsuperscript{o}) correspond to low
140 wettability and hydrophobic surfaces, while small contact angles
141 ($\theta$ $\ll$ 90\textsuperscript{o}) correspond to high wettability and
142 hydrophilic surfaces. Experimentally, measurements of the contact angle
143 of sessile drops has been used to quantify the extent of wetting on surfaces
144 with thermally selective wetting charactaristics\cite{Tadanaga00,Liu04,Sun04},
145 as well as nano-pillared surfaces with electrically tunable Cassie-Baxter and
146 Wenzel states\cite{Herbertson06,Dhindsa06,Verplanck07,Ahuja08,Manukyan11}.
147 Luzar and coworkers have done significant work modeling these transitions on
148 nano-patterned surfaces\cite{Daub07,Daub10,Daub11,Ritchie12}, and have found
149 the change in contact angle to be due to the external field perturbing the
150 hydrogen bonding of the liquid/vapor interface\cite{Daub07}.
151
152 In this paper we present the same analysis for the pyramidal and secondary
153 prismatic facets, and show that the differential interfacial friction
154 coefficients for the four facets of ice-I$_\mathrm{h}$ are determined by their
155 relative hydrophilicity by means of dynamics water contact angle simulations.
156
157 \section{Methodology}
158
159 \subsection{Water Model}
160 Understanding ice/water interfaces inherently begins with the isolated
161 systems. There has been extensive work parameterizing models for liquid water,
162 such as the SPC\cite{Berendsen81}, SPC/E\cite{Berendsen87},
163 TIP4P\cite{Jorgensen85}, TIP4P/2005\cite{Abascal05},
164 ($\dots$), and more recently, models for simulating
165 the solid phases of water, such as the TIP4P/Ice\cite{Abascal05b} model. The
166 melting point of various crystal structures of ice have been calculated for
167 many of these models
168 (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}),
169 and the partial or complete phase diagram for the model has been determined
170 (SPC/E\cite{Baez95,Bryk04b,Sanz04b}, TIP4P\cite{Sanz04,Sanz04b,Koyama04}, TIP5P\cite{Sanz04,Koyama04}).
171
172 Haymet et al. have studied the quiescent Ice-I$_\mathrm{h}$/water interface
173 using the rigid SPC, SPC/E, TIP4P, and the flexible CF1 water models, and has seen good
174 agreement for structural and dynamic measurements of the interfacial
175 width. Given the expansive size of our systems of interest, and the
176 apparent independence of water model on interfacial width, we have chosen to use the rigid SPC/E
177 water model in this study.
178
179 \subsection{Pyramidal and secondary prismatic ice/water interface construction}
180 To construct the pyramidal and secondary prismatic ice/water systems,
181 first a proton-ordered zero dipole crystal of ice-I$_\mathrm{h}$ with exposed strips
182 of H-atoms and lone pairs was constructed from Structure 6 of Hirsch
183 and Ojam\"{a}e's recent paper\cite{Hirsch04}. The crystal was then cut
184 along the plane of the desired facet, and reoriented so that the
185 $z$-axis was perpdicular to the exposed face. Two orthoganol cuts were
186 then made to the crystal such that perfect periodic replication could
187 be perfromed in the $x$ and $y$ dimensions. The slab was then
188 replicated along the $x$ and $y$ axes until the desired crystal size
189 was obtained. Liquid water boxes were created having identical
190 dimensions (in $x$ and$y$) as the ice blocks, and a $z$ dimension of
191 three times that of the ice block. Each of the ice slabs and water
192 boxes were independently equilibrated to 50K, and the resulting
193 systems were merged by carving out any liquid water molecules within 3
194 \AA\ of any atoms in the ice slabs. Each of the combined ice/water
195 systems were then equilibrated to 225K, which was found to be a stable
196 temperature for each of the interfaces over a 5 ns simulation.
197 For a more detailed explanation of
198 the ice/water systems construction, please refer to a previous
199 paper\cite{Louden13}. The resulting dimensions, number of ice, and liquid water molecules
200 contained in each of these systems can be seen in Table \ref{tab:method}.
201 \subsection{Shearing simulations}
202 % Do we need to justify the sims at 225K?
203 % No crystal growth or shrinkage over 2 successive 1 ns NVT simulations for
204 % either the pyramidal or sec. prismatic ice/water systems.
205 To perform the shearing simulations, the velocity shearing and scaling
206 varient of reverse nonequilibrium molecular dynamics (VSS-RNEMD) was
207 conducted. This method performs a series of simultaneous velocity
208 exchanges between two regions of the simulation cell, to
209 simultaneously create a velocity and temperature gradient. The thermal
210 gradient is necessary when performing shearing simulations as to
211 prevent frictional heating from the shear from melting the
212 interface. For more details on the VSS-RNEMD method please refer to a
213 pervious paper\cite{Louden13}.
214
215 The computational details performed here were equivalent to those reported
216 in a previous publication\cite{Louden13}, with the following changes.
217 VSS-RNEMD moves were attempted every 2 fs instead of every 50 fs. This was done to minimize
218 the magnitude of each individual VSS-RNEMD perturbation to the system.
219 All pyramidal simulations were performed under the canonical (NVT) ensamble
220 except those during which configurations were accumulated for the orientational correlation
221 function, which were performed under the microcanonical (NVE) ensamble. All
222 secondary prismatic simulations were performed under the NVE ensamble.
223
224 \subsection{Droplet simulations}
225 To construct ice surfaces to perform water contact angle calculations
226 on, ice crystals were created as described earlier (see Pyramidal and
227 secondary prismatic ice/water interface construction). The crystals
228 were then cut from the negative $z$ dimension, ensuring the remaining
229 ice crystal was thicker in $z$ than the potential cutoff. The crystals
230 were then replicated in $x$ and $y$ until a sufficiently large surface
231 had been created. The sizes and number of molecules in each of the surfaces is given in Table
232 \ref{tab:ice_sheets}. Molecular restraints were applied to the center of mass
233 of the rigid bodies to prevent surface melting, however the molecules were
234 allowed to reorient themselves freely. The water doplet contained 2048
235 SPC/E molecules, which has been found to produce
236 agreement for the Young contact angle extrapolated to an infinite drop
237 size\cite{Daub10}. The surfaces and droplet were independently
238 equilibrated to 225 K, at which time the droplet was placed 3-5 \AA\
239 above the positive $z$ dimension of the surface at 5 unique
240 locations. Each simulation was 5 ns in length and conducted in the NVE ensemble.
241
242
243 \section{Results and discussion}
244 \subsection{Interfacial width}
245 In the literature there is good agreement that between the solid ice and
246 the bulk water, there exists a region of 'slush-like' water molecules.
247 In this region, the water molecules are structurely distinguishable and
248 behave differently than those of the solid ice or the bulk water.
249 The characteristics of this region have been defined by both structural
250 and dynamic properties; and its width has been measured by the change of these
251 properties from their bulk liquid values to those of the solid ice.
252 Examples of these properties include the density, the diffusion constant, and
253 the translational order profile. \cite{Bryk02,Karim90,Gay02,Hayward01,Hayward02,Karim88}
254
255 Since the VSS-RNEMD moves used to impose the thermal and velocity gradients
256 perturb the momenta of the water molecules in
257 the systems, parameters that depend on translational motion may give
258 faulty results. A stuructural parameter will be less effected by the
259 VSS-RNEMD perturbations to the system. Due to this, we have used the
260 local tetrahedral order parameter (Eq 5\cite{Louden13} to quantify the width of the interface,
261 which was originally described by Kumar\cite{Kumar09} and
262 Errington\cite{Errington01}, and used by Bryk and Haymet in a previous study
263 of ice/water interfaces.\cite{Bryk04b}
264
265 To determine the width of the interfaces, each of the systems were
266 divided into 100 artificial bins along the
267 $z$-dimension, and the local tetrahedral order parameter, $q(z)$, was
268 time-averaged for each of the bins, resulting in a tetrahedrality profile of
269 the system. These profiles are shown across the $z$-dimension of the systems
270 in panel $a$ of Figures \ref{fig:pyrComic}
271 and \ref{fig:spComic} (black circles). The $q(z)$ function has a range of
272 (0,1), where a larger value indicates a more tetrahedral environment.
273 The $q(z)$ for the bulk liquid was found to be $\approx $ 0.77, while values of
274 $\approx $ 0.92 were more common for the ice. The tetrahedrality profiles were
275 fit using a hyperbolic tangent\cite{Louden13} designed to smoothly fit the
276 bulk to ice
277 transition, while accounting for the thermal influence on the profile by the
278 kinetic energy exchanges of the VSS-RNEMD moves. In panels $b$ and $c$, the
279 resulting thermal and velocity gradients from an imposed kinetic energy and
280 momentum fluxes can be seen. The verticle dotted
281 lines traversing all three panels indicate the midpoints of the interface
282 as determined by the hyperbolic tangent fit of the tetrahedrality profiles.
283
284 From fitting the tetrahedrality profiles for each of the 0.5 nanosecond
285 simulations (panel c of Figures \ref{fig:pyrComic} and \ref{fig:spComic})
286 by eq. 6\cite{Louden13},we find the interfacial width to be
287 3.2 $\pm$ 0.2 and 3.2 $\pm$ 0.2 \AA\ for the control system with no applied
288 momentum flux for both the pyramidal and secondary prismatic systems.
289 Over the range of shear rates investigated,
290 0.6 $\pm$ 0.2 $\mathrm{ms}^{-1} \rightarrow$ 5.6 $\pm$ 0.4 $\mathrm{ms}^{-1}$
291 for the pyramidal system and 0.9 $\pm$ 0.3 $\mathrm{ms}^{-1} \rightarrow$ 5.4
292 $\pm$ 0.1 $\mathrm{ms}^{-1}$ for the secondary prismatic, we found no
293 significant change in the interfacial width. This follows our previous
294 findings of the basal and
295 prismatic systems, in which the interfacial width was invarient of the
296 shear rate of the ice. The interfacial width of the quiescent basal and
297 prismatic systems was found to be 3.2 $\pm$ 0.4 \AA\ and 3.6 $\pm$ 0.2 \AA\
298 respectively, over the range of shear rates investigated, 0.6 $\pm$ 0.3
299 $\mathrm{ms}^{-1} \rightarrow$ 5.3 $\pm$ 0.5 $\mathrm{ms}^{-1}$ for the basal
300 system and 0.9 $\pm$ 0.2 $\mathrm{ms}^{-1} \rightarrow$ 4.5 $\pm$ 0.1
301 $\mathrm{ms}^{-1}$ for the prismatic.
302
303 These results indicate that the surface structure of the exposed ice crystal
304 has little to no effect on how far into the bulk the ice-like structural
305 ordering is. Also, it appears that the interface is not structurally effected
306 by the movement of water over the ice.
307
308
309 \subsection{Orientational dynamics}
310 %Should we include the math here?
311 The orientational time correlation function,
312 \begin{equation}\label{C(t)1}
313 C_{2}(t)=\langle P_{2}(\mathbf{u}(0)\cdot \mathbf{u}(t))\rangle,
314 \end{equation}
315 helps indicate the local environment around the water molecules. The function
316 begins with an initial value of unity, and decays to zero as the water molecule
317 loses memory of its former orientation. Observing the rate at which this decay
318 occurs can provide insight to the mechanism and timescales for the relaxation.
319 In eq. \eqref{C(t)1}, $P_{2}$ is the second-order Legendre polynomial, and
320 $\mathbf{u}$ is the bisecting HOH vector. The angle brackets indicate
321 an ensemble average over all the water molecules in a given spatial region.
322
323 To investigate the dynamics of the water molecules across the interface, the
324 systems were divided in the $z$-dimension into bins, each $\approx$ 3 \AA\
325 wide, and eq. \eqref{C(t)1} was computed for each of the bins. A water
326 molecule was allocated to a particular bin if it was initially in the bin
327 at time zero. To compute eq. \eqref{C(t)1}, each 0.5 ns simulation was
328 followed by an additional 200 ps NVE simulation during which the
329 position and orientations of each molecule were recorded every 0.1 ps.
330
331 The data obtained for each bin was then fit to a triexponential decay
332 with the three decay constants
333 $\tau_{short}$ corresponding to the librational motion of the water
334 molecules, $\tau_{middle}$ corresponding to jumps between the breaking and
335 making of hydrogen bonds, and $\tau_{long}$ corresponding to the translational
336 motion of the water molecules. An additive constant in the fit accounts
337 for the water molecules trapped in the ice which do not experience any
338 long-time orientational decay.
339
340 In Figures \ref{fig:PyrOrient} and \ref{fig:SPorient} we see the $z$-coordinate
341 profiles for the three decay constants, $\tau_{short}$ (panel a),
342 $\tau_{middle}$ (panel b),
343 and $\tau_{long}$ (panel c) for the pyramidal and secondary prismatic systems
344 respectively. The control experiments (no shear) are shown in black, and
345 an experiment with an imposed momentum flux is shown in red. The vertical
346 dotted line traversing all three panels denotes the midpoint of the
347 interface as determined by the local tetrahedral order parameter fitting.
348 In the liquid regions of both systems, we see that $\tau_{middle}$ and
349 $\tau_{long}$ have approximately consistent values of $3-6$ ps and $30-40$ ps,
350 resepctively, and increase in value as we approach the interface. Conversely,
351 in panel a, we see that $\tau_{short}$ decreases from the liquid value
352 of $72-76$ fs as we approach the interface. We believe this speed up is due to
353 the constrained motion of librations closer to the interface. Both the
354 approximate values for the decays and trends approaching the interface match
355 those reported previously for the basal and prismatic interfaces.
356
357 As done previously, we have attempted to quantify the distance, $d_{pyramidal}$
358 and $d_{secondary prismatic}$, from the
359 interface that the deviations from the bulk liquid values begin. This was done
360 by fitting the orientational decay constant $z$-profiles by
361 \begin{equation}\label{tauFit}
362 \tau(z)\approx\tau_{liquid}+(\tau_{wall}-\tau_{liquid})e^{-(z-z_{wall})/d}
363 \end{equation}
364 where $\tau_{liquid}$ and $\tau_{wall}$ are the liquid and projected wall
365 values of the decay constants, $z_{wall}$ is the location of the interface,
366 and $d$ is the displacement from the interface at which these deviations
367 occur. The values for $d_{pyramidal}$ and $d_{secondary prismatic}$ were
368 determined
369 for each of the decay constants, and then averaged for better statistics
370 ($\tau_{middle}$ was ommitted for secondary prismatic). For the pyramidal
371 system,
372 $d_{pyramidal}$ was found to be 2.7 \AA\ for both the control and the sheared
373 system. We found $d_{secondary prismatic}$ to be slightly larger than
374 $d_{pyramidal}$ for both the control and with an applied shear, with
375 displacements of $4$ \AA\ for the control system and $3$ \AA\ for the
376 experiment with the imposed momentum flux. These values are consistent with
377 those found for the basal ($d_{basal}\approx2.9$ \AA\ ) and prismatic
378 ($d_{prismatic}\approx3.5$ \AA\ ) systems.
379
380 \subsection{Coefficient of friction of the interfaces}
381 While investigating the kinetic coefficient of friction, there was found
382 to be a dependence for $\mu_k$
383 on the temperature of the liquid water in the system. We believe this
384 dependence
385 arrises from the sharp discontinuity of the viscosity for the SPC/E model
386 at temperatures approaching 200 K\cite{Kuang12}. Due to this, we propose
387 a weighting to the interfacial friction coefficient, $\kappa$ by the
388 shear viscosity of the fluid at 225 K. The interfacial friction coefficient
389 relates the shear stress with the relative velocity of the fluid normal to the
390 interface:
391 \begin{equation}\label{Shenyu-13}
392 j_{z}(p_{x}) = \kappa[v_{x}(fluid)-v_{x}(solid)]
393 \end{equation}
394 where $j_{z}(p_{x})$ is the applied momentum flux (shear stress) across $z$
395 in the
396 $x$-dimension, and $v_{x}$(fluid) and $v_{x}$(solid) are the velocities
397 directly adjacent to the interface. The shear viscosity, $\eta(T)$, of the
398 fluid can be determined under a linear response of the momentum
399 gradient to the applied shear stress by
400 \begin{equation}\label{Shenyu-11}
401 j_{z}(p_{x}) = \eta(T) \frac{\partial v_{x}}{\partial z}.
402 \end{equation}
403 Using eqs \eqref{Shenyu-13} and \eqref{Shenyu-11}, we can find the following
404 expression for $\kappa$,
405 \begin{equation}\label{kappa-1}
406 \kappa = \eta(T) \frac{\partial v_{x}}{\partial z}\frac{1}{[v_{x}(fluid)-v_{x}(solid)]}.
407 \end{equation}
408 Here is where we will introduce the weighting term of $\eta(225)/\eta(T)$
409 giving us
410 \begin{equation}\label{kappa-2}
411 \kappa = \frac{\eta(225)}{[v_{x}(fluid)-v_{x}(solid)]}\frac{\partial v_{x}}{\partial z}.
412 \end{equation}
413
414 To obtain the value of $\eta(225)$ for the SPC/E model, a $31.09 \times 29.38
415 \times 124.39$ \AA\ box with 3744 SPC/E liquid water molecules was
416 equilibrated to 225K,
417 and 5 unique shearing experiments were performed. Each experiment was
418 conducted in the NVE and were 5 ns in
419 length. The VSS were attempted every timestep, which was set to 2 fs.
420 For our SPC/E systems, we found $\eta(225)$ to be 0.0148 $\pm$ 0.0007 Pa s,
421 roughly ten times larger than the value found for 280 K SPC/E bulk water by
422 Kuang\cite{Kuang12}.
423
424 The interfacial friction coefficient, $\kappa$, can equivalently be expressed
425 as the ratio of the viscosity of the fluid to the slip length, $\delta$, which
426 is an indication of how 'slippery' the interface is.
427 \begin{equation}\label{kappa-3}
428 \kappa = \frac{\eta}{\delta}
429 \end{equation}
430 In each of the systems, the interfacial temperature was kept fixed to 225K,
431 which ensured the viscosity of the fluid at the
432 interace was approximately the same. Thus, any significant variation in
433 $\kappa$ between
434 the systems indicates differences in the 'slipperiness' of the interfaces.
435 As each of the ice systems are sheared relative to liquid water, the
436 'slipperiness' of the interface can be taken as an indication of how
437 hydrophobic or hydrophilic the interface is. The calculated $\kappa$ values
438 found for the four crystal facets of Ice-I$_\mathrm{h}$ investigated are shown
439 in Table \ref{tab:kappa}. The basal and pyramidal facets were found to have
440 similar values of $\kappa \approx$ 0.0006
441 (amu \AA\textsuperscript{-2} fs\textsuperscript{-1}), while values of
442 $\kappa \approx$ 0.0003 (amu \AA\textsuperscript{-2} fs\textsuperscript{-1})
443 were found for the prismatic and secondary prismatic systems.
444 These results indicate that the basal and pyramidal facets are
445 more hydrophilic than the prismatic and secondary prismatic facets.
446
447 \subsection{Dynamic water contact angle}
448
449
450
451
452 \section{Conclusion}
453 We present the results of molecular dynamics simulations of the pyrmaidal
454 and secondary prismatic facets of an SPC/E model of the
455 Ice-I$_\mathrm{h}$/water interface. The ice was sheared through the liquid
456 water while being exposed to a thermal gradient to maintain a stable
457 interface by using the minimally perturbing VSS RNEMD method. In agreement
458 with our previous findings for the basal and prismatic facets, the interfacial
459 width was found to be independent of shear rate as measured by the local
460 tetrahedral order parameter. This width was found to be
461 3.2~$\pm$ 0.2~\AA\ for both the pyramidal and the secondary prismatic systems.
462 These values are in good agreement with our previously calculated interfacial
463 widths for the basal (3.2~$\pm$ 0.4~\AA\ ) and prismatic (3.6~$\pm$ 0.2~\AA\ )
464 systems.
465
466 Orientational dynamics of the Ice-I$_\mathrm{h}$/water interfaces were studied
467 by calculation of the orientational time correlation function at varying
468 displacements normal to the interface. The decays were fit
469 to a tri-exponential decay, where the three decay constants correspond to
470 the librational motion of the molecules driven by the restoring forces of
471 existing hydrogen bonds ($\tau_{short}$ $\mathcal{O}$(10 fs)), jumps between
472 two different hydrogen bonds ($\tau_{middle}$ $\mathcal{O}$(1 ps)), and
473 translational motion of the molecules ($\tau_{long}$ $\mathcal{O}$(100 ps)).
474 $\tau_{short}$ was found to decrease approaching the interface due to the
475 constrained motion of the molecules as the local environment becomes more
476 ice-like. Conversely, the two longer-time decay constants were found to
477 increase at small displacements from the interface. As seen in our previous
478 work on the basal and prismatic facets, there appears to be a dynamic
479 interface width at which deviations from the bulk liquid values occur.
480 We had previously found $d_{basal}$ and $d_{prismatic}$ to be approximately
481 2.8~\AA\ and 3.5~\AA. We found good agreement of these values for the
482 pyramidal and secondary prismatic systems with $d_{pyramidal}$ and
483 $d_{secondary prismatic}$ to be 2.7~\AA\ and 3~\AA. For all four of the
484 facets, no apparent dependence of the dynamic width on the shear rate was
485 found.
486
487 %Paragraph summarizing the Kappa values
488 The interfacial friction coefficient, $\kappa$, was determined for each facet
489 interface. We were able to reach an expression for $\kappa$ as a function of
490 the velocity profile of the system which is scaled by the viscosity of the liquid
491 at 225 K. In doing so, we have obtained an expression for $\kappa$ which is
492 independent of temperature differences of the liquid water at far displacements
493 from the interface. We found the basal and pyramidal facets to have
494 similar $\kappa$ values, of $\kappa \approx$ 0.0006
495 (amu \AA\textsuperscript{-2} fs\textsuperscript{-1}). However, the
496 prismatic and secondary prismatic facets were found to have $\kappa$ values of
497 $\kappa \approx$ 0.0003 (amu \AA\textsuperscript{-2} fs\textsuperscript{-1}).
498 As these ice facets are being sheared relative to liquid water, with the
499 structural and dynamic width of all four
500 interfaces being approximately the same, the difference in the coefficient of
501 friction indicates the hydrophilicity of the crystal facets are not
502 equivalent. Namely, that the basal and pyramidal facets of Ice-I$_\mathrm{h}$
503 are more hydrophilic than the prismatic and secondary prismatic facets.
504
505
506 \begin{acknowledgments}
507 Support for this project was provided by the National
508 Science Foundation under grant CHE-1362211. Computational time was
509 provided by the Center for Research Computing (CRC) at the
510 University of Notre Dame.
511 \end{acknowledgments}
512
513 \newpage
514
515 \bibliography{iceWater.bib}
516
517 There is significant interest in the properties of ice/ice and ice/water
518 interfaces in the geophysics community. Most commonly, the results of shearing
519 two ice blocks past one
520 another\cite{Casassa91, Sukhorukov13, Pritchard12, Lishman13} or the shearing
521 of ice through water\cite{Cuffey99, Bell08}. Using molecular dynamics
522 simulations, Samadashvili has recently shown that when two smooth ice slabs
523 slide past one another, a stable liquid-like layer develops between
524 them\cite{Samadashvili13}. To fundamentally understand these processes, a
525 molecular understanding of the ice/water interfaces is needed.
526
527 Investigation of the ice/water interface is also crucial in understanding
528 processes such as nucleation, crystal
529 growth,\cite{Han92, Granasy95, Vanfleet95} and crystal
530 melting\cite{Weber83, Han92, Sakai96, Sakai96B}. Insight gained to these
531 properties can also be applied to biological systems of interest, such as
532 the behavior of the antifreeze protein found in winter
533 flounder,\cite{Wierzbicki07,Chapsky97} and certain terrestial
534 arthropods.\cite{Duman:2001qy,Meister29012013} Elucidating the properties which
535 give rise to these processes through experimental techniques can be expensive,
536 complicated, and sometimes infeasible. However, through the use of molecular
537 dynamics simulations much of the problems of investigating these properties
538 are alleviated.
539
540 Understanding ice/water interfaces inherently begins with the isolated
541 systems. There has been extensive work parameterizing models for liquid water,
542 such as the SPC\cite{Berendsen81}, SPC/E\cite{Berendsen87},
543 TIP4P\cite{Jorgensen85}, TIP4P/2005\cite{Abascal05},
544 ($\dots$), and more recently, models for simulating
545 the solid phases of water, such as the TIP4P/Ice\cite{Abascal05b} model. The
546 melting point of various crystal structures of ice have been calculated for
547 many of these models
548 (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}),
549 and the partial or complete phase diagram for the model has been determined
550 (SPC/E\cite{Baez95,Bryk04b,Sanz04b}, TIP4P\cite{Sanz04,Sanz04b,Koyama04}, TIP5P\cite{Sanz04,Koyama04}).
551 Knowing the behavior and melting point for these models has enabled an initial
552 investigation of ice/water interfaces.
553
554 The Ice-I$_\mathrm{h}$/water quiescent interface has been extensively studied
555 over the past 30 years by theory and experiment. Haymet \emph{et al.} have
556 done significant work characterizing and quantifying the width of these
557 interfaces for the SPC,\cite{Karim90} SPC/E,\cite{Gay02,Bryk02},
558 CF1,\cite{Hayward01,Hayward02} and TIP4P\cite{Karim88} models for water. In
559 recent years, Haymet has focused on investigating the effects cations and
560 anions have on crystal nucleaion and
561 melting.\cite{Bryk04,Smith05,Wilson08,Wilson10} Nada and Furukawa have studied
562 the the basal- and prismatic-water interface width\cite{Nada95}, crystal
563 surface restructuring at temperatures approaching the melting
564 point\cite{Nada00}, and the mechanism of ice growth inhibition by antifreeze
565 proteins\cite{Nada08,Nada11,Nada12}. Nada has developed a six-site water model
566 for ice/water interfaces near the melting point\cite{Nada03}, and studied the
567 dependence of crystal growth shape on applied pressure\cite{Nada11b}. Using
568 this model, Nada and Furukawa have established differential
569 growth rates for the basal, prismatic, and secondary prismatic facets of
570 Ice-I$_\mathrm{h}$ and found their origins due to a reordering of the hydrogen
571 bond network in water near the interface\cite{Nada05}. While the work
572 described so far has mainly focused on bulk water on ice, there is significant
573 interest in thin films of water on ice surfaces as well.
574
575 It is well known that the surface of ice exhibits a premelting layer at
576 temperatures near the melting point, often called a quasi-liquid layer (QLL).
577 Molecular dynamics simulations of the facets of ice-I$_\mathrm{h}$ exposed
578 to vacuum performed by Conde, Vega and Patrykiejew have found QLL widths of
579 approximately 10 \AA\ at 3 K below the melting point\cite{Conde08}.
580 Similarly, Limmer and Chandler have used course grain simulations and
581 statistical field theory to estimated QLL widths at the same temperature to
582 be about 3 nm\cite{Limmer14}.
583 Recently, Sazaki and Furukawa have developed an experimental technique with
584 sufficient spatial and temporal resolution to visulaize and quantitatively
585 analyze QLLs on ice crystals at temperatures near melting\cite{Sazaki10}. They
586 have found the width of the QLLs perpindicular to the surface at -2.2$^{o}$C
587 to be $\mathcal{O}$(\AA). They have also seen the formation of two immiscible
588 QLLs, which displayed different stabilities and dynamics on the crystal
589 surface\cite{Sazaki12}. Knowledge of the hydrophilicities of each
590 of the crystal facets would help further our understanding of the properties
591 and dynamics of the QLLs.
592
593 Presented here is the follow up to our previous paper\cite{Louden13}, in which
594 the basal and prismatic facets of an ice-I$_\mathrm{h}$/water interface were
595 investigated where the ice was sheared relative to the liquid. By using a
596 recently developed velocity shearing and scaling approach to reverse
597 non-equilibrium molecular dynamics (VSS-RNEMD), simultaneous temperature and
598 velocity gradients can be applied to the system, which allows for measurment
599 of friction and thermal transport properties while maintaining a stable
600 interfacial temperature\cite{Kuang12}. Structural analysis and dynamic
601 correlation functions were used to probe the interfacial response to a shear,
602 and the resulting solid/liquid kinetic friction coefficients were reported.
603 In this paper we present the same analysis for the pyramidal and secondary
604 prismatic facets, and show that the differential interfacial friction
605 coefficients for the four facets of ice-I$_\mathrm{h}$ are determined by their
606 relative hydrophilicity by means of dynamics water contact angle
607 simulations.
608
609 The local tetrahedral order parameter, $q(z)$, is given by
610 \begin{equation}
611 q(z) = \int_0^L \sum_{k=1}^{N} \Bigg(1 -\frac{3}{8}\sum_{i=1}^{3}
612 \sum_{j=i+1}^{4} \bigg(\cos\psi_{ikj}+\frac{1}{3}\bigg)^2\Bigg)
613 \delta(z_{k}-z)\mathrm{d}z \Bigg/ N_z
614 \label{eq:qz}
615 \end{equation}
616 where $\psi_{ikj}$ is the angle formed between the oxygen sites of molecules
617 $i$,$k$, and $j$, where the centeral oxygen is located within molecule $k$ and
618 molecules $i$ and $j$ are two of the closest four water molecules
619 around molecule $k$. All four closest neighbors of molecule $k$ are also
620 required to reside within the first peak of the pair distribution function
621 for molecule $k$ (typically $<$ 3.41 \AA\ for water).
622 $N_z = \int\delta(z_k - z) \mathrm{d}z$ is a normalization factor to account
623 for the varying population of molecules within each finite-width bin.
624
625
626 The hydrophobicity or hydrophilicity of a surface can be described by the
627 extent a droplet of water wets the surface. The contact angle formed between
628 the solid and the liquid, $\theta$, which relates the free energies of the
629 three interfaces involved, is given by Young's equation.
630 \begin{equation}\label{young}
631 \cos\theta = (\gamma_{sv} - \gamma_{sl})/\gamma_{lv}
632 \end{equation}
633 Here $\gamma_{sv}$, $\gamma_{sl}$, and $\gamma_{lv}$ are the free energies
634 of the solid/vapor, solid/liquid, and liquid/vapor interfaces respectively.
635 Large contact angles ($\theta$ $\gg$ 90\textsuperscript{o}) correspond to low
636 wettability and hydrophobic surfaces, while small contact angles
637 ($\theta$ $\ll$ 90\textsuperscript{o}) correspond to high wettability and
638 hydrophilic surfaces. Experimentally, measurements of the contact angle
639 of sessile drops has been used to quantify the extent of wetting on surfaces
640 with thermally selective wetting charactaristics\cite{Tadanaga00,Liu04,Sun04},
641 as well as nano-pillared surfaces with electrically tunable Cassie-Baxter and
642 Wenzel states\cite{Herbertson06,Dhindsa06,Verplanck07,Ahuja08,Manukyan11}.
643 Luzar and coworkers have done significant work modeling these transitions on
644 nano-patterned surfaces\cite{Daub07,Daub10,Daub11,Ritchie12}, and have found
645 the change in contact angle to be due to the external field perturbing the
646 hydrogen bonding of the liquid/vapor interface\cite{Daub07}.
647
648
649
650 \end{article}
651
652 \begin{figure}
653 \includegraphics[width=\linewidth]{Pyr_comic_strip}
654 \caption{\label{fig:pyrComic} The pyramidal interface with a shear
655 rate of 3.8 ms\textsuperscript{-1}. Lower panel: the local tetrahedral order
656 parameter, $q(z)$, (black circles) and the hyperbolic tangent fit (red line).
657 Middle panel: the imposed thermal gradient required to maintain a fixed
658 interfacial temperature. Upper panel: the transverse velocity gradient that
659 develops in response to an imposed momentum flux. The vertical dotted lines
660 indicate the locations of the midpoints of the two interfaces.}
661 \end{figure}
662
663 \begin{figure}
664 \includegraphics[width=\linewidth]{SP_comic_strip}
665 \caption{\label{fig:spComic} The secondary prismatic interface with a shear
666 rate of 3.5 \
667 ms\textsuperscript{-1}. Panel descriptions match those in figure \ref{fig:pyrComic}.}
668 \end{figure}
669
670 \begin{figure}
671 \includegraphics[width=\linewidth]{Pyr-orient}
672 \caption{\label{fig:PyrOrient} The three decay constants of the
673 orientational time correlation function, $C_2(t)$, for water as a function
674 of distance from the center of the ice slab. The vertical dashed line
675 indicates the edge of the pyramidal ice slab determined by the local order
676 tetrahedral parameter. The control (black circles) and sheared (red squares)
677 experiments were fit by a shifted exponential decay (Eq. 9\cite{Louden13})
678 shown by the black and red lines respectively. The upper two panels show that
679 translational and hydrogen bond making and breaking events slow down
680 through the interface while approaching the ice slab. The bottom most panel
681 shows the librational motion of the water molecules speeding up approaching
682 the ice block due to the confined region of space allowed for the molecules
683 to move in.}
684 \end{figure}
685
686 \begin{figure}
687 \includegraphics[width=\linewidth]{SP-orient-less}
688 \caption{\label{fig:SPorient} Decay constants for $C_2(t)$ at the secondary
689 prismatic face. Panel descriptions match those in \ref{fig:PyrOrient}.}
690 \end{figure}
691
692
693 \begin{table}[h]
694 \centering
695 \caption{Phyiscal properties of the basal, prismatic, pyramidal, and secondary prismatic facets of Ice-I$_\mathrm{h}$}
696 \label{tab:kappa}
697 \begin{tabular}{|ccccc|} \hline
698 & \multicolumn{2}{c}{$\kappa_{Drag direction}$
699 (x10\textsuperscript{-4} amu \AA\textsuperscript{-2} fs\textsuperscript{-1})} & & \\
700 Interface & $\kappa_{x}$ & $\kappa_{y}$ & $\theta_{\infty}$ & $K_{spread} (ns^{-1})$ \\ \hline
701 basal & $5.9 \pm 0.3$ & $6.5 \pm 0.8$ & $34.1 \pm 0.9$ & $0.60 \pm 0.07$ \\
702 pyramidal & $5.8 \pm 0.4$ & $6.1 \pm 0.5$ & $35 \pm 3$ & $0.7 \pm 0.1$ \\
703 prismatic & $3.0 \pm 0.2$ & $3.0 \pm 0.1$ & $45 \pm 3$ & $0.75 \pm 0.09$ \\
704 secondary prismatic & $3.5 \pm 0.1$ & $3.3 \pm 0.2$ & $42 \pm 2$ & $0.69 \pm 0.03$ \\ \hline
705 \end{tabular}
706 \end{table}
707
708
709 \begin{table}[h]
710 \centering
711 \caption{Shearing and Droplet simulation parameters}
712 \label{tab:method}
713 \begin{tabular}{|cccc|ccc|} \hline
714 & \multicolumn{3}{c}{Shearing} & \multicolumn{3}{c}{Droplet}\\
715 Interface & $N_{ice}$ & $N_{liquid}$ & Lx, Ly, Lz (\AA) &
716 $N_{ice}$ & $N_{droplet}$ & Lx, Ly (\AA) \\ \hline
717 Basal & 900 & 1846 & (23.87, 35.83, 98.64) & 12960 & 2048 & (134.70, 140.04)\\
718 Prismatic & 3000 & 5464 & (35.95, 35.65, 205.77) & 9900 & 2048 &
719 (110.04, 115.00)\\
720 Pyramidal & 1216 & 2203& (37.47, 29.50, 93.02) & 11136 & 2048 &
721 (143.75, 121.41)\\
722 Secondary Prismatic & 3840 & 8176 & (71.87, 31.66, 161.55) & 11520 &
723 2048 & (146.72, 124.48)\\
724 \hline
725 \end{tabular}
726 \end{table}
727
728 \end{document}