ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/iceWater2/iceWater3.tex
Revision: 4235
Committed: Mon Dec 8 18:31:51 2014 UTC (9 years, 9 months ago) by plouden
Content type: application/x-tex
File size: 41524 byte(s)
Log Message:
Revised Methodology - Droplet Simulations.

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{I}n this article we study the evolution of ``almost-sharp'' fronts
69 %for the surface quasi-geostrophic equation. This 2-D active scalar
70 %equation reads for the surface quasi-geostrophic equation.
71 %\begin{equation}
72 %\mfrac{D \theta}{Dt}=\mfrac{\pr \theta}{\pr t} + u\cdot \nabla
73 %\theta=0 \label{qg1}
74 %\end{equation}
75
76 The Ice-I$_\mathrm{h}$/water quiescent interface has been extensively studied
77 over the past 30 years by theory and experiment. Haymet \emph{et al.} have
78 done significant work characterizing and quantifying the width of these
79 interfaces for the SPC,\cite{Karim90} SPC/E,\cite{Gay02,Bryk02},
80 CF1,\cite{Hayward01,Hayward02} and TIP4P\cite{Karim88} models for water.\cite{Bryk04,Smith05,Wilson08,Wilson10} Nada and Furukawa have studied
81 the the basal- and prismatic-water interface width\cite{Nada95} and crystal
82 surface restructuring at temperatures approaching the melting
83 point\cite{Nada00}.
84
85 It is well known that the surface of ice exhibits a premelting layer at
86 temperatures near the melting point, often called a quasi-liquid layer (QLL).
87 Molecular dynamics simulations of the facets of ice-I$_\mathrm{h}$ exposed
88 to vacuum performed by Conde, Vega and Patrykiejew have found QLL widths of
89 approximately 10 \AA\ at 3 K below the melting point\cite{Conde08}.
90 Similarly, Limmer and Chandler have used course grain simulations and
91 statistical field theory to estimated QLL widths at the same temperature to
92 be about 3 nm\cite{Limmer14}.
93 Recently, Sazaki and Furukawa have developed an experimental technique with
94 sufficient spatial and temporal resolution to visulaize and quantitatively
95 analyze QLLs on ice crystals at temperatures near melting\cite{Sazaki10}. They
96 have found the width of the QLLs perpindicular to the surface at -2.2$^{o}$C
97 to be 3-4 \AA\ wide. They have also seen the formation of two immiscible
98 QLLs, which displayed different stabilities and dynamics on the crystal
99 surface\cite{Sazaki12}.
100
101 There is significant interest in the properties of ice/ice and ice/water
102 interfaces in the geophysics community. Most commonly, the results of shearing
103 two ice blocks past one
104 another\cite{Casassa91, Sukhorukov13, Pritchard12, Lishman13} or the shearing
105 of ice through water\cite{Cuffey99, Bell08}. Using molecular dynamics
106 simulations, Samadashvili has recently shown that when two smooth ice slabs
107 slide past one another, a stable liquid-like layer develops between
108 them\cite{Samadashvili13}. To fundamentally understand these processes, a
109 molecular understanding of the ice/water interfaces is needed.
110 Investigation of the ice/water interface is also crucial in understanding
111 processes such as nucleation, crystal
112 growth,\cite{Han92, Granasy95, Vanfleet95} and crystal
113 melting\cite{Weber83, Han92, Sakai96, Sakai96B}.
114
115 In a previous study\cite{Louden13}, we investigated
116 the basal and prismatic facets of an ice-I$_\mathrm{h}$/water
117 interface where the ice was sheared relative to the liquid. Using
118 velocity shearing and scaling approach to reverse
119 non-equilibrium molecular dynamics (VSS-RNEMD), simultaneous temperature and
120 velocity gradients were applied to the system, allowing for measurment
121 of friction and thermal transport properties while maintaining a stable
122 interfacial temperature\cite{Kuang12}.
123
124 Paragraph here about hydrophobicity and hydrophilicity, maybe move up
125 more in the paper as well. Talk about physically what it means for a
126 surface to by hydrophobic or hydrophilic, and then we move into
127 how do we define it (mathematically) and then measure the degree
128 of wetting experimentally and theoretically.
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 The resulting solid/liquid kinetic friction coefficients were
153 reported, and displayed a factor of two difference between the
154 basal and prismatic facets.
155 In this paper we present the same analysis for the pyramidal and secondary
156 prismatic facets, and show that the differential interfacial friction
157 coefficients for the four facets of ice-I$_\mathrm{h}$ are determined by their
158 relative hydrophilicity by means of dynamics water contact angle simulations.
159
160 \section{Methodology}
161
162 \subsection{Water Model}
163 Understanding ice/water interfaces inherently begins with the isolated
164 systems. There has been extensive work parameterizing models for liquid water,
165 such as the SPC\cite{Berendsen81}, SPC/E\cite{Berendsen87},
166 TIP4P\cite{Jorgensen85}, TIP4P/2005\cite{Abascal05},
167 ($\dots$), and more recently, models for simulating
168 the solid phases of water, such as the TIP4P/Ice\cite{Abascal05b} model. The
169 melting point of various crystal structures of ice have been calculated for
170 many of these models
171 (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}),
172 and the partial or complete phase diagram for the model has been determined
173 (SPC/E\cite{Baez95,Bryk04b,Sanz04b}, TIP4P\cite{Sanz04,Sanz04b,Koyama04}, TIP5P\cite{Sanz04,Koyama04}).
174
175 Haymet et al. have studied the quiescent Ice-I$_\mathrm{h}$/water interface
176 using the rigid SPC, SPC/E, TIP4P, and the flexible CF1 water models, and has seen good
177 agreement for structural and dynamic measurements of the interfacial
178 width. Given the expansive size of our systems of interest, and the
179 apparent independence of water model on interfacial width, we have chosen to use the rigid SPC/E
180 water model in this study.
181
182 \subsection{Pyramidal and secondary prismatic ice/water interface construction}
183 To construct the pyramidal and secondary prismatic ice/water systems,
184 first a proton-ordered zero dipole crystal of ice-I$_\mathrm{h}$ with exposed strips
185 of H-atoms and lone pairs was constructed from Structure 6 of Hirsch
186 and Ojam\"{a}e's recent paper\cite{Hirsch04}. The crystal was then cut
187 along the plane of the desired facet, and reoriented so that the
188 $z$-axis was perpdicular to the exposed face. Two orthoganol cuts were
189 then made to the crystal such that perfect periodic replication could
190 be perfromed in the $x$ and $y$ dimensions. The slab was then
191 replicated along the $x$ and $y$ axes until the desired crystal size
192 was obtained. Liquid water boxes were created having identical
193 dimensions (in $x$ and$y$) as the ice blocks, and a $z$ dimension of
194 three times that of the ice block. Each of the ice slabs and water
195 boxes were independently equilibrated to 50K, and the resulting
196 systems were merged by carving out any liquid water molecules within 3
197 \AA\ of any atoms in the ice slabs. Each of the combined ice/water
198 systems were then equilibrated to 225K, which was found to be a stable
199 temperature for each of the interfaces over a 5 ns simulation.
200 For a more detailed explanation of
201 the ice/water systems construction, please refer to a previous
202 paper\cite{Louden13}. The resulting dimensions, number of ice, and liquid water molecules
203 contained in each of these systems can be seen in Table \ref{tab:method}.
204 \subsection{Shearing simulations}
205 % Do we need to justify the sims at 225K?
206 % No crystal growth or shrinkage over 2 successive 1 ns NVT simulations for
207 % either the pyramidal or sec. prismatic ice/water systems.
208 To perform the shearing simulations, the velocity shearing and scaling
209 varient of reverse nonequilibrium molecular dynamics (VSS-RNEMD) was
210 conducted. This method performs a series of simultaneous velocity
211 exchanges between two regions of the simulation cell, to
212 simultaneously create a velocity and temperature gradient. The thermal
213 gradient is necessary when performing shearing simulations as to
214 prevent frictional heating from the shear from melting the
215 interface. For more details on the VSS-RNEMD method please refer to a
216 pervious paper\cite{Louden13}.
217
218 The computational details performed here were equivalent to those reported
219 in a previous publication\cite{Louden13}, with the following changes.
220 VSS-RNEMD moves were attempted every 2 fs instead of every 50 fs. This was done to minimize
221 the magnitude of each individual VSS-RNEMD perturbation to the system.
222 All pyramidal simulations were performed under the canonical (NVT) ensamble
223 except those during which configurations were accumulated for the orientational correlation
224 function, which were performed under the microcanonical (NVE) ensamble. All
225 secondary prismatic simulations were performed under the NVE ensamble.
226
227 \subsection{Droplet simulations}
228 To construct ice surfaces to perform water contact angle calculations
229 on, ice crystals were created as described earlier (see Pyramidal and
230 secondary prismatic ice/water interface construction). The crystals
231 were then cut from the negative $z$ dimension, ensuring the remaining
232 ice crystal was thicker in $z$ than the potential cutoff. The crystals
233 were then replicated in $x$ and $y$ until a sufficiently large surface
234 had been created. The sizes and number of molecules in each of the surfaces is given in Table
235 \ref{tab:ice_sheets}. Molecular restraints were applied to the center of mass
236 of the rigid bodies to prevent surface melting, however the molecules were
237 allowed to reorient themselves freely. The water doplet contained 2048
238 SPC/E molecules, which has been found to produce
239 agreement for the Young contact angle extrapolated to an infinite drop
240 size\cite{Daub10}. The surfaces and droplet were independently
241 equilibrated to 225 K, at which time the droplet was placed 3-5 \AA\
242 above the positive $z$ dimension of the surface at 5 unique
243 locations. Each simulation was 5 ns in length and conducted in the NVE ensemble.
244
245
246 \section{Results and discussion}
247 \subsection{Interfacial width}
248 In the literature there is good agreement that between the solid ice and
249 the bulk water, there exists a region of 'slush-like' water molecules.
250 In this region, the water molecules are structurely distinguishable and
251 behave differently than those of the solid ice or the bulk water.
252 The characteristics of this region have been defined by both structural
253 and dynamic properties; and its width has been measured by the change of these
254 properties from their bulk liquid values to those of the solid ice.
255 Examples of these properties include the density, the diffusion constant, and
256 the translational order profile. \cite{Bryk02,Karim90,Gay02,Hayward01,Hayward02,Karim88}
257
258 Since the VSS-RNEMD moves used to impose the thermal and velocity gradients
259 perturb the momenta of the water molecules in
260 the systems, parameters that depend on translational motion may give
261 faulty results. A stuructural parameter will be less effected by the
262 VSS-RNEMD perturbations to the system. Due to this, we have used the
263 local tetrahedral order parameter (Eq 5\cite{Louden13} to quantify the width of the interface,
264 which was originally described by Kumar\cite{Kumar09} and
265 Errington\cite{Errington01}, and used by Bryk and Haymet in a previous study
266 of ice/water interfaces.\cite{Bryk04b}
267
268 To determine the width of the interfaces, each of the systems were
269 divided into 100 artificial bins along the
270 $z$-dimension, and the local tetrahedral order parameter, $q(z)$, was
271 time-averaged for each of the bins, resulting in a tetrahedrality profile of
272 the system. These profiles are shown across the $z$-dimension of the systems
273 in panel $a$ of Figures \ref{fig:pyrComic}
274 and \ref{fig:spComic} (black circles). The $q(z)$ function has a range of
275 (0,1), where a larger value indicates a more tetrahedral environment.
276 The $q(z)$ for the bulk liquid was found to be $\approx $ 0.77, while values of
277 $\approx $ 0.92 were more common for the ice. The tetrahedrality profiles were
278 fit using a hyperbolic tangent\cite{Louden13} designed to smoothly fit the
279 bulk to ice
280 transition, while accounting for the thermal influence on the profile by the
281 kinetic energy exchanges of the VSS-RNEMD moves. In panels $b$ and $c$, the
282 resulting thermal and velocity gradients from an imposed kinetic energy and
283 momentum fluxes can be seen. The verticle dotted
284 lines traversing all three panels indicate the midpoints of the interface
285 as determined by the hyperbolic tangent fit of the tetrahedrality profiles.
286
287 From fitting the tetrahedrality profiles for each of the 0.5 nanosecond
288 simulations (panel c of Figures \ref{fig:pyrComic} and \ref{fig:spComic})
289 by eq. 6\cite{Louden13},we find the interfacial width to be
290 3.2 $\pm$ 0.2 and 3.2 $\pm$ 0.2 \AA\ for the control system with no applied
291 momentum flux for both the pyramidal and secondary prismatic systems.
292 Over the range of shear rates investigated,
293 0.6 $\pm$ 0.2 $\mathrm{ms}^{-1} \rightarrow$ 5.6 $\pm$ 0.4 $\mathrm{ms}^{-1}$
294 for the pyramidal system and 0.9 $\pm$ 0.3 $\mathrm{ms}^{-1} \rightarrow$ 5.4
295 $\pm$ 0.1 $\mathrm{ms}^{-1}$ for the secondary prismatic, we found no
296 significant change in the interfacial width. This follows our previous
297 findings of the basal and
298 prismatic systems, in which the interfacial width was invarient of the
299 shear rate of the ice. The interfacial width of the quiescent basal and
300 prismatic systems was found to be 3.2 $\pm$ 0.4 \AA\ and 3.6 $\pm$ 0.2 \AA\
301 respectively, over the range of shear rates investigated, 0.6 $\pm$ 0.3
302 $\mathrm{ms}^{-1} \rightarrow$ 5.3 $\pm$ 0.5 $\mathrm{ms}^{-1}$ for the basal
303 system and 0.9 $\pm$ 0.2 $\mathrm{ms}^{-1} \rightarrow$ 4.5 $\pm$ 0.1
304 $\mathrm{ms}^{-1}$ for the prismatic.
305
306 These results indicate that the surface structure of the exposed ice crystal
307 has little to no effect on how far into the bulk the ice-like structural
308 ordering is. Also, it appears that the interface is not structurally effected
309 by shearing the ice through water.
310
311
312 \subsection{Orientational dynamics}
313 %Should we include the math here?
314 The orientational time correlation function,
315 \begin{equation}\label{C(t)1}
316 C_{2}(t)=\langle P_{2}(\mathbf{u}(0)\cdot \mathbf{u}(t))\rangle,
317 \end{equation}
318 helps indicate the local environment around the water molecules. The function
319 begins with an initial value of unity, and decays to zero as the water molecule
320 loses memory of its former orientation. Observing the rate at which this decay
321 occurs can provide insight to the mechanism and timescales for the relaxation.
322 In eq. \eqref{C(t)1}, $P_{2}$ is the second-order Legendre polynomial, and
323 $\mathbf{u}$ is the bisecting HOH vector. The angle brackets indicate
324 an ensemble average over all the water molecules in a given spatial region.
325
326 To investigate the dynamics of the water molecules across the interface, the
327 systems were divided in the $z$-dimension into bins, each $\approx$ 3 \AA\
328 wide, and eq. \eqref{C(t)1} was computed for each of the bins. A water
329 molecule was allocated to a particular bin if it was initially in the bin
330 at time zero. To compute eq. \eqref{C(t)1}, each 0.5 ns simulation was
331 followed by an additional 200 ps NVE simulation during which the
332 position and orientations of each molecule were recorded every 0.1 ps.
333
334 The data obtained for each bin was then fit to a triexponential decay
335 with the three decay constants
336 $\tau_{short}$ corresponding to the librational motion of the water
337 molecules, $\tau_{middle}$ corresponding to jumps between the breaking and
338 making of hydrogen bonds, and $\tau_{long}$ corresponding to the translational
339 motion of the water molecules. An additive constant in the fit accounts
340 for the water molecules trapped in the ice which do not experience any
341 long-time orientational decay.
342
343 In Figures \ref{fig:PyrOrient} and \ref{fig:SPorient} we see the $z$-coordinate
344 profiles for the three decay constants, $\tau_{short}$ (panel a),
345 $\tau_{middle}$ (panel b),
346 and $\tau_{long}$ (panel c) for the pyramidal and secondary prismatic systems
347 respectively. The control experiments (no shear) are shown in black, and
348 an experiment with an imposed momentum flux is shown in red. The vertical
349 dotted line traversing all three panels denotes the midpoint of the
350 interface as determined by the local tetrahedral order parameter fitting.
351 In the liquid regions of both systems, we see that $\tau_{middle}$ and
352 $\tau_{long}$ have approximately consistent values of $3-6$ ps and $30-40$ ps,
353 resepctively, and increase in value as we approach the interface. Conversely,
354 in panel a, we see that $\tau_{short}$ decreases from the liquid value
355 of $72-76$ fs as we approach the interface. We believe this speed up is due to
356 the constrained motion of librations closer to the interface. Both the
357 approximate values for the decays and trends approaching the interface match
358 those reported previously for the basal and prismatic interfaces.
359
360 As done previously, we have attempted to quantify the distance, $d_{pyramidal}$
361 and $d_{secondary prismatic}$, from the
362 interface that the deviations from the bulk liquid values begin. This was done
363 by fitting the orientational decay constant $z$-profiles by
364 \begin{equation}\label{tauFit}
365 \tau(z)\approx\tau_{liquid}+(\tau_{wall}-\tau_{liquid})e^{-(z-z_{wall})/d}
366 \end{equation}
367 where $\tau_{liquid}$ and $\tau_{wall}$ are the liquid and projected wall
368 values of the decay constants, $z_{wall}$ is the location of the interface,
369 and $d$ is the displacement from the interface at which these deviations
370 occur. The values for $d_{pyramidal}$ and $d_{secondary prismatic}$ were
371 determined
372 for each of the decay constants, and then averaged for better statistics
373 ($\tau_{middle}$ was ommitted for secondary prismatic). For the pyramidal
374 system,
375 $d_{pyramidal}$ was found to be 2.7 \AA\ for both the control and the sheared
376 system. We found $d_{secondary prismatic}$ to be slightly larger than
377 $d_{pyramidal}$ for both the control and with an applied shear, with
378 displacements of $4$ \AA\ for the control system and $3$ \AA\ for the
379 experiment with the imposed momentum flux. These values are consistent with
380 those found for the basal ($d_{basal}\approx2.9$ \AA\ ) and prismatic
381 ($d_{prismatic}\approx3.5$ \AA\ ) systems.
382
383 \subsection{Coefficient of friction of the interfaces}
384 While investigating the kinetic coefficient of friction, there was found
385 to be a dependence for $\mu_k$
386 on the temperature of the liquid water in the system. We believe this
387 dependence
388 arrises from the sharp discontinuity of the viscosity for the SPC/E model
389 at temperatures approaching 200 K\cite{Kuang12}. Due to this, we propose
390 a weighting to the interfacial friction coefficient, $\kappa$ by the
391 shear viscosity of the fluid at 225 K. The interfacial friction coefficient
392 relates the shear stress with the relative velocity of the fluid normal to the
393 interface:
394 \begin{equation}\label{Shenyu-13}
395 j_{z}(p_{x}) = \kappa[v_{x}(fluid)-v_{x}(solid)]
396 \end{equation}
397 where $j_{z}(p_{x})$ is the applied momentum flux (shear stress) across $z$
398 in the
399 $x$-dimension, and $v_{x}$(fluid) and $v_{x}$(solid) are the velocities
400 directly adjacent to the interface. The shear viscosity, $\eta(T)$, of the
401 fluid can be determined under a linear response of the momentum
402 gradient to the applied shear stress by
403 \begin{equation}\label{Shenyu-11}
404 j_{z}(p_{x}) = \eta(T) \frac{\partial v_{x}}{\partial z}.
405 \end{equation}
406 Using eqs \eqref{Shenyu-13} and \eqref{Shenyu-11}, we can find the following
407 expression for $\kappa$,
408 \begin{equation}\label{kappa-1}
409 \kappa = \eta(T) \frac{\partial v_{x}}{\partial z}\frac{1}{[v_{x}(fluid)-v_{x}(solid)]}.
410 \end{equation}
411 Here is where we will introduce the weighting term of $\eta(225)/\eta(T)$
412 giving us
413 \begin{equation}\label{kappa-2}
414 \kappa = \frac{\eta(225)}{[v_{x}(fluid)-v_{x}(solid)]}\frac{\partial v_{x}}{\partial z}.
415 \end{equation}
416
417 To obtain the value of $\eta(225)$ for the SPC/E model, a $31.09 \times 29.38
418 \times 124.39$ \AA\ box with 3744 SPC/E liquid water molecules was
419 equilibrated to 225K,
420 and 5 unique shearing experiments were performed. Each experiment was
421 conducted in the NVE and were 5 ns in
422 length. The VSS were attempted every timestep, which was set to 2 fs.
423 For our SPC/E systems, we found $\eta(225)$ to be 0.0148 $\pm$ 0.0007 Pa s,
424 roughly ten times larger than the value found for 280 K SPC/E bulk water by
425 Kuang\cite{Kuang12}.
426
427 The interfacial friction coefficient, $\kappa$, can equivalently be expressed
428 as the ratio of the viscosity of the fluid to the slip length, $\delta$, which
429 is an indication of how 'slippery' the interface is.
430 \begin{equation}\label{kappa-3}
431 \kappa = \frac{\eta}{\delta}
432 \end{equation}
433 In each of the systems, the interfacial temperature was kept fixed to 225K,
434 which ensured the viscosity of the fluid at the
435 interace was approximately the same. Thus, any significant variation in
436 $\kappa$ between
437 the systems indicates differences in the 'slipperiness' of the interfaces.
438 As each of the ice systems are sheared relative to liquid water, the
439 'slipperiness' of the interface can be taken as an indication of how
440 hydrophobic or hydrophilic the interface is. The calculated $\kappa$ values
441 found for the four crystal facets of Ice-I$_\mathrm{h}$ investigated are shown
442 in Table \ref{tab:kappa}. The basal and pyramidal facets were found to have
443 similar values of $\kappa \approx$ 0.0006
444 (amu \AA\textsuperscript{-2} fs\textsuperscript{-1}), while values of
445 $\kappa \approx$ 0.0003 (amu \AA\textsuperscript{-2} fs\textsuperscript{-1})
446 were found for the prismatic and secondary prismatic systems.
447 These results indicate that the basal and pyramidal facets are
448 more hydrophilic than the prismatic and secondary prismatic facets.
449
450 \subsection{Dynamic water contact angle}
451
452
453
454
455 \section{Conclusion}
456 We present the results of molecular dynamics simulations of the pyrmaidal
457 and secondary prismatic facets of an SPC/E model of the
458 Ice-I$_\mathrm{h}$/water interface. The ice was sheared through the liquid
459 water while being exposed to a thermal gradient to maintain a stable
460 interface by using the minimally perturbing VSS RNEMD method. In agreement
461 with our previous findings for the basal and prismatic facets, the interfacial
462 width was found to be independent of shear rate as measured by the local
463 tetrahedral order parameter. This width was found to be
464 3.2~$\pm$ 0.2~\AA\ for both the pyramidal and the secondary prismatic systems.
465 These values are in good agreement with our previously calculated interfacial
466 widths for the basal (3.2~$\pm$ 0.4~\AA\ ) and prismatic (3.6~$\pm$ 0.2~\AA\ )
467 systems.
468
469 Orientational dynamics of the Ice-I$_\mathrm{h}$/water interfaces were studied
470 by calculation of the orientational time correlation function at varying
471 displacements normal to the interface. The decays were fit
472 to a tri-exponential decay, where the three decay constants correspond to
473 the librational motion of the molecules driven by the restoring forces of
474 existing hydrogen bonds ($\tau_{short}$ $\mathcal{O}$(10 fs)), jumps between
475 two different hydrogen bonds ($\tau_{middle}$ $\mathcal{O}$(1 ps)), and
476 translational motion of the molecules ($\tau_{long}$ $\mathcal{O}$(100 ps)).
477 $\tau_{short}$ was found to decrease approaching the interface due to the
478 constrained motion of the molecules as the local environment becomes more
479 ice-like. Conversely, the two longer-time decay constants were found to
480 increase at small displacements from the interface. As seen in our previous
481 work on the basal and prismatic facets, there appears to be a dynamic
482 interface width at which deviations from the bulk liquid values occur.
483 We had previously found $d_{basal}$ and $d_{prismatic}$ to be approximately
484 2.8~\AA\ and 3.5~\AA. We found good agreement of these values for the
485 pyramidal and secondary prismatic systems with $d_{pyramidal}$ and
486 $d_{secondary prismatic}$ to be 2.7~\AA\ and 3~\AA. For all four of the
487 facets, no apparent dependence of the dynamic width on the shear rate was
488 found.
489
490 %Paragraph summarizing the Kappa values
491 The interfacial friction coefficient, $\kappa$, was determined for each facet
492 interface. We were able to reach an expression for $\kappa$ as a function of
493 the velocity profile of the system which is scaled by the viscosity of the liquid
494 at 225 K. In doing so, we have obtained an expression for $\kappa$ which is
495 independent of temperature differences of the liquid water at far displacements
496 from the interface. We found the basal and pyramidal facets to have
497 similar $\kappa$ values, of $\kappa \approx$ 0.0006
498 (amu \AA\textsuperscript{-2} fs\textsuperscript{-1}). However, the
499 prismatic and secondary prismatic facets were found to have $\kappa$ values of
500 $\kappa \approx$ 0.0003 (amu \AA\textsuperscript{-2} fs\textsuperscript{-1}).
501 As these ice facets are being sheared relative to liquid water, with the
502 structural and dynamic width of all four
503 interfaces being approximately the same, the difference in the coefficient of
504 friction indicates the hydrophilicity of the crystal facets are not
505 equivalent. Namely, that the basal and pyramidal facets of Ice-I$_\mathrm{h}$
506 are more hydrophilic than the prismatic and secondary prismatic facets.
507
508
509 \begin{acknowledgments}
510 Support for this project was provided by the National
511 Science Foundation under grant CHE-1362211. Computational time was
512 provided by the Center for Research Computing (CRC) at the
513 University of Notre Dame.
514 \end{acknowledgments}
515
516 \newpage
517
518 \bibliography{iceWater.bib}
519
520 There is significant interest in the properties of ice/ice and ice/water
521 interfaces in the geophysics community. Most commonly, the results of shearing
522 two ice blocks past one
523 another\cite{Casassa91, Sukhorukov13, Pritchard12, Lishman13} or the shearing
524 of ice through water\cite{Cuffey99, Bell08}. Using molecular dynamics
525 simulations, Samadashvili has recently shown that when two smooth ice slabs
526 slide past one another, a stable liquid-like layer develops between
527 them\cite{Samadashvili13}. To fundamentally understand these processes, a
528 molecular understanding of the ice/water interfaces is needed.
529
530 Investigation of the ice/water interface is also crucial in understanding
531 processes such as nucleation, crystal
532 growth,\cite{Han92, Granasy95, Vanfleet95} and crystal
533 melting\cite{Weber83, Han92, Sakai96, Sakai96B}. Insight gained to these
534 properties can also be applied to biological systems of interest, such as
535 the behavior of the antifreeze protein found in winter
536 flounder,\cite{Wierzbicki07,Chapsky97} and certain terrestial
537 arthropods.\cite{Duman:2001qy,Meister29012013} Elucidating the properties which
538 give rise to these processes through experimental techniques can be expensive,
539 complicated, and sometimes infeasible. However, through the use of molecular
540 dynamics simulations much of the problems of investigating these properties
541 are alleviated.
542
543 Understanding ice/water interfaces inherently begins with the isolated
544 systems. There has been extensive work parameterizing models for liquid water,
545 such as the SPC\cite{Berendsen81}, SPC/E\cite{Berendsen87},
546 TIP4P\cite{Jorgensen85}, TIP4P/2005\cite{Abascal05},
547 ($\dots$), and more recently, models for simulating
548 the solid phases of water, such as the TIP4P/Ice\cite{Abascal05b} model. The
549 melting point of various crystal structures of ice have been calculated for
550 many of these models
551 (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}),
552 and the partial or complete phase diagram for the model has been determined
553 (SPC/E\cite{Baez95,Bryk04b,Sanz04b}, TIP4P\cite{Sanz04,Sanz04b,Koyama04}, TIP5P\cite{Sanz04,Koyama04}).
554 Knowing the behavior and melting point for these models has enabled an initial
555 investigation of ice/water interfaces.
556
557 The Ice-I$_\mathrm{h}$/water quiescent interface has been extensively studied
558 over the past 30 years by theory and experiment. Haymet \emph{et al.} have
559 done significant work characterizing and quantifying the width of these
560 interfaces for the SPC,\cite{Karim90} SPC/E,\cite{Gay02,Bryk02},
561 CF1,\cite{Hayward01,Hayward02} and TIP4P\cite{Karim88} models for water. In
562 recent years, Haymet has focused on investigating the effects cations and
563 anions have on crystal nucleaion and
564 melting.\cite{Bryk04,Smith05,Wilson08,Wilson10} Nada and Furukawa have studied
565 the the basal- and prismatic-water interface width\cite{Nada95}, crystal
566 surface restructuring at temperatures approaching the melting
567 point\cite{Nada00}, and the mechanism of ice growth inhibition by antifreeze
568 proteins\cite{Nada08,Nada11,Nada12}. Nada has developed a six-site water model
569 for ice/water interfaces near the melting point\cite{Nada03}, and studied the
570 dependence of crystal growth shape on applied pressure\cite{Nada11b}. Using
571 this model, Nada and Furukawa have established differential
572 growth rates for the basal, prismatic, and secondary prismatic facets of
573 Ice-I$_\mathrm{h}$ and found their origins due to a reordering of the hydrogen
574 bond network in water near the interface\cite{Nada05}. While the work
575 described so far has mainly focused on bulk water on ice, there is significant
576 interest in thin films of water on ice surfaces as well.
577
578 It is well known that the surface of ice exhibits a premelting layer at
579 temperatures near the melting point, often called a quasi-liquid layer (QLL).
580 Molecular dynamics simulations of the facets of ice-I$_\mathrm{h}$ exposed
581 to vacuum performed by Conde, Vega and Patrykiejew have found QLL widths of
582 approximately 10 \AA\ at 3 K below the melting point\cite{Conde08}.
583 Similarly, Limmer and Chandler have used course grain simulations and
584 statistical field theory to estimated QLL widths at the same temperature to
585 be about 3 nm\cite{Limmer14}.
586 Recently, Sazaki and Furukawa have developed an experimental technique with
587 sufficient spatial and temporal resolution to visulaize and quantitatively
588 analyze QLLs on ice crystals at temperatures near melting\cite{Sazaki10}. They
589 have found the width of the QLLs perpindicular to the surface at -2.2$^{o}$C
590 to be $\mathcal{O}$(\AA). They have also seen the formation of two immiscible
591 QLLs, which displayed different stabilities and dynamics on the crystal
592 surface\cite{Sazaki12}. Knowledge of the hydrophilicities of each
593 of the crystal facets would help further our understanding of the properties
594 and dynamics of the QLLs.
595
596 Presented here is the follow up to our previous paper\cite{Louden13}, in which
597 the basal and prismatic facets of an ice-I$_\mathrm{h}$/water interface were
598 investigated where the ice was sheared relative to the liquid. By using a
599 recently developed velocity shearing and scaling approach to reverse
600 non-equilibrium molecular dynamics (VSS-RNEMD), simultaneous temperature and
601 velocity gradients can be applied to the system, which allows for measurment
602 of friction and thermal transport properties while maintaining a stable
603 interfacial temperature\cite{Kuang12}. Structural analysis and dynamic
604 correlation functions were used to probe the interfacial response to a shear,
605 and the resulting solid/liquid kinetic friction coefficients were reported.
606 In this paper we present the same analysis for the pyramidal and secondary
607 prismatic facets, and show that the differential interfacial friction
608 coefficients for the four facets of ice-I$_\mathrm{h}$ are determined by their
609 relative hydrophilicity by means of dynamics water contact angle
610 simulations.
611
612 The local tetrahedral order parameter, $q(z)$, is given by
613 \begin{equation}
614 q(z) = \int_0^L \sum_{k=1}^{N} \Bigg(1 -\frac{3}{8}\sum_{i=1}^{3}
615 \sum_{j=i+1}^{4} \bigg(\cos\psi_{ikj}+\frac{1}{3}\bigg)^2\Bigg)
616 \delta(z_{k}-z)\mathrm{d}z \Bigg/ N_z
617 \label{eq:qz}
618 \end{equation}
619 where $\psi_{ikj}$ is the angle formed between the oxygen sites of molecules
620 $i$,$k$, and $j$, where the centeral oxygen is located within molecule $k$ and
621 molecules $i$ and $j$ are two of the closest four water molecules
622 around molecule $k$. All four closest neighbors of molecule $k$ are also
623 required to reside within the first peak of the pair distribution function
624 for molecule $k$ (typically $<$ 3.41 \AA\ for water).
625 $N_z = \int\delta(z_k - z) \mathrm{d}z$ is a normalization factor to account
626 for the varying population of molecules within each finite-width bin.
627
628
629 The hydrophobicity or hydrophilicity of a surface can be described by the
630 extent a droplet of water wets the surface. The contact angle formed between
631 the solid and the liquid, $\theta$, which relates the free energies of the
632 three interfaces involved, is given by Young's equation.
633 \begin{equation}\label{young}
634 \cos\theta = (\gamma_{sv} - \gamma_{sl})/\gamma_{lv}
635 \end{equation}
636 Here $\gamma_{sv}$, $\gamma_{sl}$, and $\gamma_{lv}$ are the free energies
637 of the solid/vapor, solid/liquid, and liquid/vapor interfaces respectively.
638 Large contact angles ($\theta$ $\gg$ 90\textsuperscript{o}) correspond to low
639 wettability and hydrophobic surfaces, while small contact angles
640 ($\theta$ $\ll$ 90\textsuperscript{o}) correspond to high wettability and
641 hydrophilic surfaces. Experimentally, measurements of the contact angle
642 of sessile drops has been used to quantify the extent of wetting on surfaces
643 with thermally selective wetting charactaristics\cite{Tadanaga00,Liu04,Sun04},
644 as well as nano-pillared surfaces with electrically tunable Cassie-Baxter and
645 Wenzel states\cite{Herbertson06,Dhindsa06,Verplanck07,Ahuja08,Manukyan11}.
646 Luzar and coworkers have done significant work modeling these transitions on
647 nano-patterned surfaces\cite{Daub07,Daub10,Daub11,Ritchie12}, and have found
648 the change in contact angle to be due to the external field perturbing the
649 hydrogen bonding of the liquid/vapor interface\cite{Daub07}.
650
651
652
653 \end{article}
654
655 \begin{figure}
656 \includegraphics[width=\linewidth]{Pyr_comic_strip}
657 \caption{\label{fig:pyrComic} The pyramidal interface with a shear
658 rate of 3.8 ms\textsuperscript{-1}. Lower panel: the local tetrahedral order
659 parameter, $q(z)$, (black circles) and the hyperbolic tangent fit (red line).
660 Middle panel: the imposed thermal gradient required to maintain a fixed
661 interfacial temperature. Upper panel: the transverse velocity gradient that
662 develops in response to an imposed momentum flux. The vertical dotted lines
663 indicate the locations of the midpoints of the two interfaces.}
664 \end{figure}
665
666 \begin{figure}
667 \includegraphics[width=\linewidth]{SP_comic_strip}
668 \caption{\label{fig:spComic} The secondary prismatic interface with a shear
669 rate of 3.5 \
670 ms\textsuperscript{-1}. Panel descriptions match those in figure \ref{fig:pyrComic}.}
671 \end{figure}
672
673 \begin{figure}
674 \includegraphics[width=\linewidth]{Pyr-orient}
675 \caption{\label{fig:PyrOrient} The three decay constants of the
676 orientational time correlation function, $C_2(t)$, for water as a function
677 of distance from the center of the ice slab. The vertical dashed line
678 indicates the edge of the pyramidal ice slab determined by the local order
679 tetrahedral parameter. The control (black circles) and sheared (red squares)
680 experiments were fit by a shifted exponential decay (Eq. 9\cite{Louden13})
681 shown by the black and red lines respectively. The upper two panels show that
682 translational and hydrogen bond making and breaking events slow down
683 through the interface while approaching the ice slab. The bottom most panel
684 shows the librational motion of the water molecules speeding up approaching
685 the ice block due to the confined region of space allowed for the molecules
686 to move in.}
687 \end{figure}
688
689 \begin{figure}
690 \includegraphics[width=\linewidth]{SP-orient-less}
691 \caption{\label{fig:SPorient} Decay constants for $C_2(t)$ at the secondary
692 prismatic face. Panel descriptions match those in \ref{fig:PyrOrient}.}
693 \end{figure}
694
695
696 \begin{table}[h]
697 \centering
698 \caption{Phyiscal properties of the basal, prismatic, pyramidal, and secondary prismatic facets of Ice-I$_\mathrm{h}$}
699 \label{tab:kappa}
700 \begin{tabular}{|ccccc|} \hline
701 & \multicolumn{2}{c}{$\kappa_{Drag direction}$
702 (x10\textsuperscript{-4} amu \AA\textsuperscript{-2} fs\textsuperscript{-1})} & & \\
703 Interface & $\kappa_{x}$ & $\kappa_{y}$ & $\theta_{\infty}$ & $K_{spread} (ns^{-1})$ \\ \hline
704 basal & $5.9 \pm 0.3$ & $6.5 \pm 0.8$ & $34.1 \pm 0.9$ & $0.60 \pm 0.07$ \\
705 pyramidal & $5.8 \pm 0.4$ & $6.1 \pm 0.5$ & $35 \pm 3$ & $0.7 \pm 0.1$ \\
706 prismatic & $3.0 \pm 0.2$ & $3.0 \pm 0.1$ & $45 \pm 3$ & $0.75 \pm 0.09$ \\
707 secondary prismatic & $3.5 \pm 0.1$ & $3.3 \pm 0.2$ & $42 \pm 2$ & $0.69 \pm 0.03$ \\ \hline
708 \end{tabular}
709 \end{table}
710
711
712 \begin{table}[h]
713 \centering
714 \caption{Shearing and Droplet simulation parameters}
715 \label{tab:method}
716 \begin{tabular}{|cccc|ccc|} \hline
717 & \multicolumn{3}{c}{Shearing} & \multicolumn{3}{c}{Droplet}\\
718 Interface & $N_{ice}$ & $N_{liquid}$ & Lx, Ly, Lz (\AA) &
719 $N_{ice}$ & $N_{droplet}$ & Lx, Ly (\AA) \\ \hline
720 Basal & 900 & 1846 & (23.87, 35.83, 98.64) & 12960 & 2048 & (134.70, 140.04)\\
721 Prismatic & 3000 & 5464 & (35.95, 35.65, 205.77) & 9900 & 2048 &
722 (110.04, 115.00)\\
723 Pyramidal & 1216 & 2203& (37.47, 29.50, 93.02) & 11136 & 2048 &
724 (143.75, 121.41)\\
725 Secondary Prismatic & 3840 & 8176 & (71.87, 31.66, 161.55) & 11520 &
726 2048 & (146.72, 124.48)\\
727 \hline
728 \end{tabular}
729 \end{table}
730
731 \end{document}