ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/iceWater2/iceWater3.tex
Revision: 4245
Committed: Wed Dec 10 20:49:40 2014 UTC (9 years, 9 months ago) by gezelter
Content type: application/x-tex
File size: 45429 byte(s)
Log Message:
New stuff

File Contents

# Content
1 %% PNAStwoS.tex
2 %% Sample file to use for PNAS articles prepared in LaTeX
3 %% For two column PNAS articles
4 %% Version1: Apr 15, 2008
5 %% Version2: Oct 04, 2013
6
7 %% BASIC CLASS FILE
8 \documentclass{pnastwo}
9
10 %% ADDITIONAL OPTIONAL STYLE FILES Font specification
11
12 %\usepackage{PNASTWOF}
13 \usepackage[version=3]{mhchem}
14 \usepackage[round,numbers,sort&compress]{natbib}
15 \usepackage{fixltx2e}
16 \bibpunct{(}{)}{,}{n}{,}{,}
17 \bibliographystyle{pnas2011}
18
19 %% OPTIONAL MACRO DEFINITIONS
20 \def\s{\sigma}
21 %%%%%%%%%%%%
22 %% For PNAS Only:
23 %\url{www.pnas.org/cgi/doi/10.1073/pnas.0709640104}
24 \copyrightyear{2014}
25 \issuedate{Issue Date}
26 \volume{Volume}
27 \issuenumber{Issue Number}
28 %\setcounter{page}{2687} %Set page number here if desired
29 %%%%%%%%%%%%
30
31 \begin{document}
32
33 \title{Friction at water / ice-I\textsubscript{h} interfaces: Do the
34 different facets of ice have different hydrophilicities?}
35
36 \author{Patrick B. Louden\affil{1}{Department of Chemistry and Biochemistry, University of Notre Dame, Notre Dame,
37 IN 46556}
38 \and
39 J. Daniel Gezelter\affil{1}{}}
40
41 \contributor{Submitted to Proceedings of the National Academy of Sciences
42 of the United States of America}
43
44 %%%Newly updated.
45 %%% If significance statement need, then can use the below command otherwise just delete it.
46 %\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.}
47
48 \maketitle
49
50 \begin{article}
51 \begin{abstract}
52 In this paper we present evidence that some of the crystal facets
53 of ice-I$_\mathrm{h}$ posess structural features that can halve
54 the effective hydrophilicity of the ice/water interface. The
55 spreading dynamics of liquid water droplets on ice facets exhibits
56 long-time behavior that differs substantially for the prismatic
57 $\{1~0~\bar{1}~0\}$ and secondary prism $\{1~1~\bar{2}~0\}$ facets
58 when compared with the basal $\{0001\}$ and pyramidal
59 $\{2~0~\bar{2}~1\}$ facets. We also present the results of
60 simulations of solid-liquid friction of the same four crystal
61 facets being drawn through liquid water. Both simulation
62 techniques provide evidence that the two prismatic faces have an
63 effective surface area in contact with the liquid water of
64 approximately half of the total surface area of the crystal. The
65 ice / water interfacial widths for all four crystal facets are
66 similar (using both structural and dynamic measures), and were
67 found to be independent of the shear rate. Additionally,
68 decomposition of orientational time correlation functions show
69 position-dependence for the short- and longer-time decay
70 components close to the interface.
71 \end{abstract}
72
73 \keywords{ice | water | interfaces | hydrophobicity}
74 \abbreviations{QLL, quasi liquid layer; MD, molecular dynamics; RNEMD,
75 reverse non-equilibrium molecular dynamics}
76
77 \dropcap{S}urfaces can be characterized as hydrophobic or hydrophilic
78 based on the strength of the interactions with water. Hydrophobic
79 surfaces do not have strong enough interactions with water to overcome
80 the internal attraction between molecules in the liquid phase, and the
81 degree of hydrophilicity of a surface can be described by the extent a
82 droplet can spread out over the surface. The contact angle formed
83 between the solid and the liquid depends on the free energies of the
84 three interfaces involved, and is given by Young's
85 equation.\cite{Young}
86 \begin{equation}\label{young}
87 \cos\theta = (\gamma_{sv} - \gamma_{sl})/\gamma_{lv} .
88 \end{equation}
89 Here $\gamma_{sv}$, $\gamma_{sl}$, and $\gamma_{lv}$ are the free
90 energies of the solid/vapor, solid/liquid, and liquid/vapor interfaces
91 respectively. Large contact angles, $\theta > 90^{\circ}$, correspond
92 to hydrophobic surfaces with low wettability, while small contact
93 angles, $\theta < 90^{\circ}$, correspond to hydrophilic surfaces.
94 Experimentally, measurements of the contact angle of sessile drops is
95 often used to quantify the extent of wetting on surfaces with
96 thermally selective wetting
97 characteristics.\cite{Tadanaga00,Liu04,Sun04}
98
99 Nanometer-scale structural features of a solid surface can influence
100 the hydrophilicity to a surprising degree. Small changes in the
101 heights and widths of nano-pillars can change a surface from
102 superhydrophobic, $\theta \ge 150^{\circ}$, to hydrophilic, $\theta
103 \sim 0^{\circ}$.\cite{CBW} This is often referred to as the
104 Cassie-Baxter to Wenzel transition. Nano-pillared surfaces with
105 electrically tunable Cassie-Baxter and Wenzel states have also been
106 observed.\cite{Herbertson06,Dhindsa06,Verplanck07,Ahuja08,Manukyan11}
107 Luzar and coworkers have modeled these transitions on nano-patterned
108 surfaces\cite{Daub07,Daub10,Daub11,Ritchie12}, and have found the
109 change in contact angle is due to the field-induced perturbation of
110 hydrogen bonding at the liquid/vapor interface.\cite{Daub07}
111
112 One would expect the interfaces of ice to be highly hydrophilic (and
113 possibly the most hydrophilic of all solid surfaces). In this paper we
114 present evidence that some of the crystal facets of ice-I$_\mathrm{h}$
115 have structural features that can halve the effective hydrophilicity.
116 Our evidence for this comes from molecular dynamics (MD) simulations
117 of the spreading dynamics of liquid droplets on these facets, as well
118 as reverse non-equilibrium molecular dynamics (RNEMD) simulations of
119 solid-liquid friction.
120
121 Quiescent ice-I$_\mathrm{h}$/water interfaces have been studied
122 extensively using computer simulations. Haymet \textit{et al.}
123 characterized and measured the width of these interfaces for the
124 SPC~\cite{Karim90}, SPC/E~\cite{Gay02,Bryk02},
125 CF1~\cite{Hayward01,Hayward02} and TIP4P~\cite{Karim88} models, in
126 both neat water and with solvated
127 ions~\cite{Bryk04,Smith05,Wilson08,Wilson10}. Nada and Furukawa have
128 studied the width of basal/water and prismatic/water
129 interfaces~\cite{Nada95} as well as crystal restructuring at
130 temperatures approaching the melting point~\cite{Nada00}.
131
132 The surface of ice exhibits a premelting layer, often called a
133 quasi-liquid layer (QLL), at temperatures near the melting point. MD
134 simulations of the facets of ice-I$_\mathrm{h}$ exposed to vacuum have
135 found QLL widths of approximately 10 \AA\ at 3 K below the melting
136 point.\cite{Conde08} Similarly, Limmer and Chandler have used the mW
137 water model~\cite{Molinero09} and statistical field theory to estimate
138 QLL widths at similar temperatures to be about 3 nm.\cite{Limmer14}
139
140 Recently, Sazaki and Furukawa have developed a technique using laser
141 confocal microscopy combined with differential interference contrast
142 microscopy that has sufficient spatial and temporal resolution to
143 visulaize and quantitatively analyze QLLs on ice crystals at
144 temperatures near melting.\cite{Sazaki10} They have found the width of
145 the QLLs perpindicular to the surface at -2.2$^{o}$C to be 3-4 \AA\
146 wide. They have also seen the formation of two immiscible QLLs, which
147 displayed different dynamics on the crystal surface.\cite{Sazaki12}
148
149 There is now significant interest in the \textit{tribological}
150 properties of ice/ice and ice/water interfaces in the geophysics
151 community. Understanding the dynamics of solid-solid shearing that is
152 mediated by a liquid layer\cite{Cuffey99, Bell08} will aid in
153 understanding the macroscopic motion of large ice
154 masses.\cite{Casassa91, Sukhorukov13, Pritchard12, Lishman13}
155
156 Using molecular dynamics simulations, Samadashvili has recently shown
157 that when two smooth ice slabs slide past one another, a stable
158 liquid-like layer develops between them.\cite{Samadashvili13} In a
159 previous study, our RNEMD simulations of ice-I$_\mathrm{h}$ shearing
160 through liquid water have provided quantitative estimates of the
161 solid-liquid kinetic friction coefficients.\cite{Louden13} These
162 displayed a factor of two difference between the basal and prismatic
163 facets. The friction was found to be independent of shear direction
164 relative to the surface orientation. We attributed facet-based
165 difference in liquid-solid friction to the 6.5 \AA\ corrugation of the
166 prismatic face which reduces the effective surface area of the ice
167 that is in direct contact with liquid water.
168
169 In the sections that follow, we outline the methodology used to
170 simulate droplet-spreading dynamics using standard MD and tribological
171 properties using RNEMD simulations. These simulation methods give
172 complementary results that point to the prismatic and secondary prism
173 facets having roughly half of their surface area in direct contact
174 with the liquid.
175
176 \section{Methodology}
177 \subsection{Construction of the Ice / Water Interfaces}
178 To construct the four interfacial ice/water systems, a proton-ordered,
179 zero-dipole crystal of ice-I$_\mathrm{h}$ with exposed strips of
180 H-atoms and lone pairs was constructed using Structure 6 of Hirsch and
181 Ojam\"{a}e's set of orthorhombic representations for
182 ice-I$_{h}$.\cite{Hirsch04} This crystal structure was cleaved along
183 the four different facets being studied. The exposed face was
184 reoriented normal to the $z$-axis of the simulation cell, and the
185 structures were and extended to form large exposed facets in
186 rectangular box geometries. Liquid water boxes were created with
187 identical dimensions (in $x$ and $y$) as the ice, and a $z$ dimension
188 of three times that of the ice block, and a density corresponding to
189 $\sim$ 1 g / cm$^3$. Each of the ice slabs and water boxes were
190 independently equilibrated, and the resulting systems were merged by
191 carving out any liquid water molecules within 3 \AA\ of any atoms in
192 the ice slabs. Each of the combined ice/water systems were then
193 equilibrated at 225K, which is the liquid-ice coexistence temperature
194 for SPC/E water.\cite{} Ref. \citealp{Louden13} contains a more
195 detailed explanation of the construction of ice/water interfaces. The
196 resulting dimensions, number of ice, and liquid water molecules
197 contained in each of these systems can be seen in Table
198 \ref{tab:method}.
199
200 We used SPC/E Why? Extensively characterized over a wide range of
201 liquid conditions. Well-studied phase diagram. Reasonably accurate
202 crystalline free energies. Mostly avoids spurious crystalline
203 morphologies like ice-i and ice-B. Most importantly, the use of SPC/E
204 has been well characterized in previous ice/water interfacial studies.
205
206
207
208 There has been extensive work parameterizing good models for liquid
209 water over a wide range of conditions. The melting points of various
210 crystal structures of ice have been calculated for many of these
211 models (SPC\cite{Karim90,Abascal07},
212 SPC/E\cite{Baez95,Arbuckle02,Gay02,Bryk02,Bryk04b,Sanz04b,Fernandez06,Abascal07,Vrbka07},
213 TIP4P\cite{Karim88,Gao00,Sanz04,Sanz04b,Koyama04,Wang05,Fernandez06,Abascal07},
214 TIP5P\cite{Sanz04,Koyama04,Wang05,Fernandez06,Abascal07}), and the
215 partial or complete phase diagram for the model has been determined
216 (SPC/E\cite{Baez95,Bryk04b,Sanz04b},
217 TIP4P\cite{Sanz04,Sanz04b,Koyama04}, TIP5P\cite{Sanz04,Koyama04}).
218
219
220 such as the SPC\cite{Berendsen81}, SPC/E\cite{Berendsen87},
221 TIP4P\cite{Jorgensen85}, TIP4P/2005\cite{Abascal05}, ($\dots$), and
222 more recently, models for simulating the solid phases of water, such
223 as the TIP4P/Ice\cite{Abascal05b} model.
224
225 Haymet et al. have studied the quiescent Ice-I$_\mathrm{h}$/water
226 interface using the rigid SPC, SPC/E, TIP4P, and the flexible CF1
227 water models, and has seen good agreement for structural and dynamic
228 measurements of the interfacial width. Given the expansive size of our
229 systems of interest, and the apparent independence of water model on
230 interfacial width, we have chosen to use the rigid SPC/E water model
231 in this study.
232
233 \subsection{Shearing simulations (interfaces in bulk water)}
234 % Should we mention number of runs, sim times, etc. ?
235 To perform the shearing simulations, the velocity shearing and scaling
236 varient of reverse nonequilibrium molecular dynamics (VSS-RNEMD) was
237 conducted. This method performs a series of simultaneous velocity
238 exchanges between two regions of the simulation cell, to
239 simultaneously create a velocity and temperature gradient. The thermal
240 gradient is necessary when performing shearing simulations as to
241 prevent frictional heating from the shear from melting the
242 interface. For more details on the VSS-RNEMD method please refer to a
243 pervious paper\cite{Louden13}.
244
245 The computational details performed here were equivalent to those reported
246 in a previous publication\cite{Louden13}, with the following changes.
247 VSS-RNEMD moves were attempted every 2 fs instead of every 50 fs. This was done to minimize
248 the magnitude of each individual VSS-RNEMD perturbation to the system.
249 All pyramidal simulations were performed under the canonical (NVT) ensamble
250 except those during which configurations were accumulated for the orientational correlation
251 function, which were performed under the microcanonical (NVE) ensamble. All
252 secondary prismatic simulations were performed under the NVE ensamble.
253
254 \subsection{Droplet simulations}
255 Ice interfaces with a thickness of $\sim 30 \AA$ were created as
256 described above, but were not solvated in a liquid box. The crystals
257 were then replicated along the $x$ and $y$ axes (parallel to the
258 surface) until a large surface had been created. The sizes and
259 numbers of molecules in each of the surfaces is given in Table
260 \ref{tab:ice_sheets}. Weak translational restraining potentials with
261 spring constants of XXXX were applied to the center of mass of each
262 molecule in order to prevent surface melting, although the molecules
263 were allowed to reorient freely. A water doplet containing 2048 SPC/E
264 molecules was created separately. Droplets of this size can produce
265 agreement with the Young contact angle extrapolated to an infinite
266 drop size\cite{Daub10}. The surfaces and droplet were independently
267 equilibrated to 225 K, at which time the droplet was placed 3-5 \AA\
268 above the surface. Five statistically independent simulations were
269 carried out for each facet, and the droplet was placed at unique $x$
270 and $y$ locations for each of these simulations. Each simulation was
271 5 ns in length and conducted in the microcanonical (NVE) ensemble.
272
273 \section{Results and discussion}
274 \subsection{Interfacial width}
275 In the literature there is good agreement that between the solid ice and
276 the bulk water, there exists a region of 'slush-like' water molecules.
277 In this region, the water molecules are structurely distinguishable and
278 behave differently than those of the solid ice or the bulk water.
279 The characteristics of this region have been defined by both structural
280 and dynamic properties; and its width has been measured by the change of these
281 properties from their bulk liquid values to those of the solid ice.
282 Examples of these properties include the density, the diffusion constant, and
283 the translational order profile. \cite{Bryk02,Karim90,Gay02,Hayward01,Hayward02,Karim88}
284
285 Since the VSS-RNEMD moves used to impose the thermal and velocity gradients
286 perturb the momenta of the water molecules in
287 the systems, parameters that depend on translational motion may give
288 faulty results. A stuructural parameter will be less effected by the
289 VSS-RNEMD perturbations to the system. Due to this, we have used the
290 local tetrahedral order parameter (Eq 5\cite{Louden13} to quantify the width of the interface,
291 which was originally described by Kumar\cite{Kumar09} and
292 Errington\cite{Errington01}, and used by Bryk and Haymet in a previous study
293 of ice/water interfaces.\cite{Bryk04b}
294
295 To determine the width of the interfaces, each of the systems were
296 divided into 100 artificial bins along the
297 $z$-dimension, and the local tetrahedral order parameter, $q(z)$, was
298 time-averaged for each of the bins, resulting in a tetrahedrality profile of
299 the system. These profiles are shown across the $z$-dimension of the systems
300 in panel $a$ of Figures \ref{fig:pyrComic}
301 and \ref{fig:spComic} (black circles). The $q(z)$ function has a range of
302 (0,1), where a larger value indicates a more tetrahedral environment.
303 The $q(z)$ for the bulk liquid was found to be $\approx $ 0.77, while values of
304 $\approx $ 0.92 were more common for the ice. The tetrahedrality profiles were
305 fit using a hyperbolic tangent\cite{Louden13} designed to smoothly fit the
306 bulk to ice
307 transition, while accounting for the thermal influence on the profile by the
308 kinetic energy exchanges of the VSS-RNEMD moves. In panels $b$ and $c$, the
309 resulting thermal and velocity gradients from an imposed kinetic energy and
310 momentum fluxes can be seen. The verticle dotted
311 lines traversing all three panels indicate the midpoints of the interface
312 as determined by the hyperbolic tangent fit of the tetrahedrality profiles.
313
314 From fitting the tetrahedrality profiles for each of the 0.5 nanosecond
315 simulations (panel c of Figures \ref{fig:pyrComic} and \ref{fig:spComic})
316 by eq. 6\cite{Louden13},we find the interfacial width to be
317 3.2 $\pm$ 0.2 and 3.2 $\pm$ 0.2 \AA\ for the control system with no applied
318 momentum flux for both the pyramidal and secondary prismatic systems.
319 Over the range of shear rates investigated,
320 0.6 $\pm$ 0.2 $\mathrm{ms}^{-1} \rightarrow$ 5.6 $\pm$ 0.4 $\mathrm{ms}^{-1}$
321 for the pyramidal system and 0.9 $\pm$ 0.3 $\mathrm{ms}^{-1} \rightarrow$ 5.4
322 $\pm$ 0.1 $\mathrm{ms}^{-1}$ for the secondary prismatic, we found no
323 significant change in the interfacial width. This follows our previous
324 findings of the basal and
325 prismatic systems, in which the interfacial width was invarient of the
326 shear rate of the ice. The interfacial width of the quiescent basal and
327 prismatic systems was found to be 3.2 $\pm$ 0.4 \AA\ and 3.6 $\pm$ 0.2 \AA\
328 respectively, over the range of shear rates investigated, 0.6 $\pm$ 0.3
329 $\mathrm{ms}^{-1} \rightarrow$ 5.3 $\pm$ 0.5 $\mathrm{ms}^{-1}$ for the basal
330 system and 0.9 $\pm$ 0.2 $\mathrm{ms}^{-1} \rightarrow$ 4.5 $\pm$ 0.1
331 $\mathrm{ms}^{-1}$ for the prismatic.
332
333 These results indicate that the surface structure of the exposed ice crystal
334 has little to no effect on how far into the bulk the ice-like structural
335 ordering is. Also, it appears that the interface is not structurally effected
336 by the movement of water over the ice.
337
338
339 \subsection{Orientational dynamics}
340 %Should we include the math here?
341 The orientational time correlation function,
342 \begin{equation}\label{C(t)1}
343 C_{2}(t)=\langle P_{2}(\mathbf{u}(0)\cdot \mathbf{u}(t))\rangle,
344 \end{equation}
345 helps indicate the local environment around the water molecules. The function
346 begins with an initial value of unity, and decays to zero as the water molecule
347 loses memory of its former orientation. Observing the rate at which this decay
348 occurs can provide insight to the mechanism and timescales for the relaxation.
349 In eq. \eqref{C(t)1}, $P_{2}$ is the second-order Legendre polynomial, and
350 $\mathbf{u}$ is the bisecting HOH vector. The angle brackets indicate
351 an ensemble average over all the water molecules in a given spatial region.
352
353 To investigate the dynamics of the water molecules across the interface, the
354 systems were divided in the $z$-dimension into bins, each $\approx$ 3 \AA\
355 wide, and eq. \eqref{C(t)1} was computed for each of the bins. A water
356 molecule was allocated to a particular bin if it was initially in the bin
357 at time zero. To compute eq. \eqref{C(t)1}, each 0.5 ns simulation was
358 followed by an additional 200 ps NVE simulation during which the
359 position and orientations of each molecule were recorded every 0.1 ps.
360
361 The data obtained for each bin was then fit to a triexponential decay
362 with the three decay constants
363 $\tau_{short}$ corresponding to the librational motion of the water
364 molecules, $\tau_{middle}$ corresponding to jumps between the breaking and
365 making of hydrogen bonds, and $\tau_{long}$ corresponding to the translational
366 motion of the water molecules. An additive constant in the fit accounts
367 for the water molecules trapped in the ice which do not experience any
368 long-time orientational decay.
369
370 In Figures \ref{fig:PyrOrient} and \ref{fig:SPorient} we see the $z$-coordinate
371 profiles for the three decay constants, $\tau_{short}$ (panel a),
372 $\tau_{middle}$ (panel b),
373 and $\tau_{long}$ (panel c) for the pyramidal and secondary prismatic systems
374 respectively. The control experiments (no shear) are shown in black, and
375 an experiment with an imposed momentum flux is shown in red. The vertical
376 dotted line traversing all three panels denotes the midpoint of the
377 interface as determined by the local tetrahedral order parameter fitting.
378 In the liquid regions of both systems, we see that $\tau_{middle}$ and
379 $\tau_{long}$ have approximately consistent values of $3-6$ ps and $30-40$ ps,
380 resepctively, and increase in value as we approach the interface. Conversely,
381 in panel a, we see that $\tau_{short}$ decreases from the liquid value
382 of $72-76$ fs as we approach the interface. We believe this speed up is due to
383 the constrained motion of librations closer to the interface. Both the
384 approximate values for the decays and trends approaching the interface match
385 those reported previously for the basal and prismatic interfaces.
386
387 As done previously, we have attempted to quantify the distance, $d_{pyramidal}$
388 and $d_{secondary prismatic}$, from the
389 interface that the deviations from the bulk liquid values begin. This was done
390 by fitting the orientational decay constant $z$-profiles by
391 \begin{equation}\label{tauFit}
392 \tau(z)\approx\tau_{liquid}+(\tau_{wall}-\tau_{liquid})e^{-(z-z_{wall})/d}
393 \end{equation}
394 where $\tau_{liquid}$ and $\tau_{wall}$ are the liquid and projected wall
395 values of the decay constants, $z_{wall}$ is the location of the interface,
396 and $d$ is the displacement from the interface at which these deviations
397 occur. The values for $d_{pyramidal}$ and $d_{secondary prismatic}$ were
398 determined
399 for each of the decay constants, and then averaged for better statistics
400 ($\tau_{middle}$ was ommitted for secondary prismatic). For the pyramidal
401 system,
402 $d_{pyramidal}$ was found to be 2.7 \AA\ for both the control and the sheared
403 system. We found $d_{secondary prismatic}$ to be slightly larger than
404 $d_{pyramidal}$ for both the control and with an applied shear, with
405 displacements of $4$ \AA\ for the control system and $3$ \AA\ for the
406 experiment with the imposed momentum flux. These values are consistent with
407 those found for the basal ($d_{basal}\approx2.9$ \AA\ ) and prismatic
408 ($d_{prismatic}\approx3.5$ \AA\ ) systems.
409
410 \subsection{Coefficient of friction of the interfaces}
411 While investigating the kinetic coefficient of friction, there was found
412 to be a dependence for $\mu_k$
413 on the temperature of the liquid water in the system. We believe this
414 dependence
415 arrises from the sharp discontinuity of the viscosity for the SPC/E model
416 at temperatures approaching 200 K\cite{Kuang12}. Due to this, we propose
417 a weighting to the interfacial friction coefficient, $\kappa$ by the
418 shear viscosity of the fluid at 225 K. The interfacial friction coefficient
419 relates the shear stress with the relative velocity of the fluid normal to the
420 interface:
421 \begin{equation}\label{Shenyu-13}
422 j_{z}(p_{x}) = \kappa[v_{x}(fluid)-v_{x}(solid)]
423 \end{equation}
424 where $j_{z}(p_{x})$ is the applied momentum flux (shear stress) across $z$
425 in the
426 $x$-dimension, and $v_{x}$(fluid) and $v_{x}$(solid) are the velocities
427 directly adjacent to the interface. The shear viscosity, $\eta(T)$, of the
428 fluid can be determined under a linear response of the momentum
429 gradient to the applied shear stress by
430 \begin{equation}\label{Shenyu-11}
431 j_{z}(p_{x}) = \eta(T) \frac{\partial v_{x}}{\partial z}.
432 \end{equation}
433 Using eqs \eqref{Shenyu-13} and \eqref{Shenyu-11}, we can find the following
434 expression for $\kappa$,
435 \begin{equation}\label{kappa-1}
436 \kappa = \eta(T) \frac{\partial v_{x}}{\partial z}\frac{1}{[v_{x}(fluid)-v_{x}(solid)]}.
437 \end{equation}
438 Here is where we will introduce the weighting term of $\eta(225)/\eta(T)$
439 giving us
440 \begin{equation}\label{kappa-2}
441 \kappa = \frac{\eta(225)}{[v_{x}(fluid)-v_{x}(solid)]}\frac{\partial v_{x}}{\partial z}.
442 \end{equation}
443
444 To obtain the value of $\eta(225)$ for the SPC/E model, a $31.09 \times 29.38
445 \times 124.39$ \AA\ box with 3744 SPC/E liquid water molecules was
446 equilibrated to 225K,
447 and 5 unique shearing experiments were performed. Each experiment was
448 conducted in the NVE and were 5 ns in
449 length. The VSS were attempted every timestep, which was set to 2 fs.
450 For our SPC/E systems, we found $\eta(225)$ to be 0.0148 $\pm$ 0.0007 Pa s,
451 roughly ten times larger than the value found for 280 K SPC/E bulk water by
452 Kuang\cite{Kuang12}.
453
454 The interfacial friction coefficient, $\kappa$, can equivalently be expressed
455 as the ratio of the viscosity of the fluid to the slip length, $\delta$, which
456 is an indication of how 'slippery' the interface is.
457 \begin{equation}\label{kappa-3}
458 \kappa = \frac{\eta}{\delta}
459 \end{equation}
460 In each of the systems, the interfacial temperature was kept fixed to 225K,
461 which ensured the viscosity of the fluid at the
462 interace was approximately the same. Thus, any significant variation in
463 $\kappa$ between
464 the systems indicates differences in the 'slipperiness' of the interfaces.
465 As each of the ice systems are sheared relative to liquid water, the
466 'slipperiness' of the interface can be taken as an indication of how
467 hydrophobic or hydrophilic the interface is. The calculated $\kappa$ values
468 found for the four crystal facets of Ice-I$_\mathrm{h}$ investigated are shown
469 in Table \ref{tab:kappa}. The basal and pyramidal facets were found to have
470 similar values of $\kappa \approx$ 0.0006
471 (amu \AA\textsuperscript{-2} fs\textsuperscript{-1}), while values of
472 $\kappa \approx$ 0.0003 (amu \AA\textsuperscript{-2} fs\textsuperscript{-1})
473 were found for the prismatic and secondary prismatic systems.
474 These results indicate that the basal and pyramidal facets are
475 more hydrophilic than the prismatic and secondary prismatic facets.
476
477 \subsection{Dynamic water contact angle}
478
479 To determine the extent of wetting for each of the four crystal
480 facets, water contact angle simuations were performed. Contact angles
481 were obtained from these simulations by two methods. In the first
482 method, the contact angle was obtained from the $z$-center of mass
483 ($z_{c.m.}$) of the water droplet as described by Hautman and Klein\cite{Hautman91}
484 and utilized by Hirvi and Pakkanen in their investigation of water
485 droplets on polyethylene and poly(vinyl chloride)
486 surface\cite{Hirvi06}. At each snapshot of the simulation, the contact
487 angle, $\theta$, was found by
488 \begin{equation}\label{contact_1}
489 \langle z_{c.m.}\rangle = 2^{-4/3}R_{0}\bigg(\frac{1-cos\theta}{2+cos\theta}\bigg)^{1/3}\frac{3+cos\theta}{2+cos\theta} ,
490 \end{equation}
491 where $R_{0}$ is the radius of the free water droplet. In the second
492 method, the contact angle was obtained from fitting the droplet's
493 $z$-profile after radial averaging to a
494 circle as described by Ruijter, Blake, and
495 Coninck\cite{Ruijter99}. At each snapshot of the simulation the water droplet was
496 broken into bins, and the location of bin containing half-bulk density was
497 stored. Due to fluctuations close to the ice, all bins located within
498 2.0 \AA\ of the ice were discarded. The remaining stored bins were
499 then fit by a circle, whose tangential intersection with the ice plane could
500 be used to calculate the water
501 contact angle. These results proved noisey and unreliable when
502 compared with the first method, for these purposes we omit the data
503 from the second method.
504
505 The resulting water contact angle profiles generated by the first method
506 had an initial value of 180$^{o}$, and decayed over time. Each of
507 these profiles were fit to a biexponential decay, with a short time
508 piece to account for the water droplet initially adhering to the
509 surface, a long time piece describing the spreading of the droplet
510 over the surface, and an additive constant to capture the infinite
511 decay of the contact angle. We have found that the rate of the water
512 droplet spreading across all four crystal facets to be $\approx$ 0.7
513 ns$^{-1}$. However, the basal and pyramidal facets
514 had an infinite decay value for $\theta$ of $\approx$ 35$^{o}$, while
515 prismatic and secondary prismatic had values for $\theta$ near
516 43$^{o}$ as seen in Table \ref{tab:kappa}. This indicates that the
517 basal and pyramidal facets are more hydrophilic than the prismatic and
518 secondary prismatic. This is in good agreement
519 with our calculations of friction coefficients, in which the basal
520 and pyramidal had a higher coefficient of kinetic friction than the
521 prismatic and secondary prismatic. Due to this, we beleive that the
522 differences in friction coefficients can be attributed to the varying
523 hydrophilicities of the facets.
524
525 \section{Conclusion}
526 We present the results of molecular dynamics simulations of the basal,
527 prismatic, pyrmaidal
528 and secondary prismatic facets of an SPC/E model of the
529 Ice-I$_\mathrm{h}$/water interface, and show that the differential
530 coefficients of friction among the four facets are due to their
531 relative hydrophilicities by means
532 of water contact angle calculations. To obtain the coefficients of
533 friction, the ice was sheared through the liquid
534 water while being exposed to a thermal gradient to maintain a stable
535 interface by using the minimally perturbing VSS RNEMD method. Water
536 contact angles are obtained by fitting the spreading of a liquid water
537 droplet over the crystal facets.
538
539 In agreement with our previous findings for the basal and prismatic facets, the interfacial
540 width of the prismatic and secondary prismatic crystal faces were
541 found to be independent of shear rate as measured by the local
542 tetrahedral order parameter. This width was found to be
543 3.2~$\pm$ 0.2~\AA\ for both the pyramidal and the secondary prismatic systems.
544 These values are in good agreement with our previously calculated interfacial
545 widths for the basal (3.2~$\pm$ 0.4~\AA\ ) and prismatic (3.6~$\pm$ 0.2~\AA\ )
546 systems.
547
548 Orientational dynamics of the Ice-I$_\mathrm{h}$/water interfaces were studied
549 by calculation of the orientational time correlation function at varying
550 displacements normal to the interface. The decays were fit
551 to a tri-exponential decay, where the three decay constants correspond to
552 the librational motion of the molecules driven by the restoring forces of
553 existing hydrogen bonds ($\tau_{short}$ $\mathcal{O}$(10 fs)), jumps between
554 two different hydrogen bonds ($\tau_{middle}$ $\mathcal{O}$(1 ps)), and
555 translational motion of the molecules ($\tau_{long}$ $\mathcal{O}$(100 ps)).
556 $\tau_{short}$ was found to decrease approaching the interface due to the
557 constrained motion of the molecules as the local environment becomes more
558 ice-like. Conversely, the two longer-time decay constants were found to
559 increase at small displacements from the interface. As seen in our previous
560 work on the basal and prismatic facets, there appears to be a dynamic
561 interface width at which deviations from the bulk liquid values occur.
562 We had previously found $d_{basal}$ and $d_{prismatic}$ to be approximately
563 2.8~\AA\ and 3.5~\AA. We found good agreement of these values for the
564 pyramidal and secondary prismatic systems with $d_{pyramidal}$ and
565 $d_{secondary prismatic}$ to be 2.7~\AA\ and 3~\AA. For all four of the
566 facets, no apparent dependence of the dynamic width on the shear rate was
567 found.
568
569 The interfacial friction coefficient, $\kappa$, was determined for each facet
570 interface. We were able to reach an expression for $\kappa$ as a function of
571 the velocity profile of the system which is scaled by the viscosity of the liquid
572 at 225 K. In doing so, we have obtained an expression for $\kappa$ which is
573 independent of temperature differences of the liquid water at far displacements
574 from the interface. We found the basal and pyramidal facets to have
575 similar $\kappa$ values, of $\kappa \approx$ 6
576 (x$10^{-4}$amu \AA\textsuperscript{-2} fs\textsuperscript{-1}). However, the
577 prismatic and secondary prismatic facets were found to have $\kappa$ values of
578 $\kappa \approx$ 3 (x$10^{-4}$amu \AA\textsuperscript{-2} fs\textsuperscript{-1}).
579 Believing this difference was due to the relative hydrophilicities of
580 the crystal faces, we have calculated the infinite decay of the water
581 contact angle, $\theta_{\infty}$, by watching the spreading of a water
582 droplet over the surface of the crystal facets. We have found
583 $\theta_{\infty}$ for the basal and pyramidal faces to be $\approx$ 34
584 degrees, while obtaining $\theta_{\infty}$ of $\approx$ 43 degrees for
585 the prismatic and secondary prismatic faces. This indicates that the
586 basal and pyramidal faces of ice-I$_\mathrm{h}$ are more hydrophilic
587 than the prismatic and secondary prismatic. These results also seem to
588 explain the differential friction coefficients obtained through the
589 shearing simulations, namely, that the coefficients of friction of the
590 ice-I$_\mathrm{h}$ crystal facets are governed by their inherent
591 hydrophilicities.
592
593
594 \begin{acknowledgments}
595 Support for this project was provided by the National
596 Science Foundation under grant CHE-1362211. Computational time was
597 provided by the Center for Research Computing (CRC) at the
598 University of Notre Dame.
599 \end{acknowledgments}
600
601 \bibliography{iceWater}
602 % *****************************************
603 % There is significant interest in the properties of ice/ice and ice/water
604 % interfaces in the geophysics community. Most commonly, the results of shearing
605 % two ice blocks past one
606 % another\cite{Casassa91, Sukhorukov13, Pritchard12, Lishman13} or the shearing
607 % of ice through water\cite{Cuffey99, Bell08}. Using molecular dynamics
608 % simulations, Samadashvili has recently shown that when two smooth ice slabs
609 % slide past one another, a stable liquid-like layer develops between
610 % them\cite{Samadashvili13}. To fundamentally understand these processes, a
611 % molecular understanding of the ice/water interfaces is needed.
612
613 % Investigation of the ice/water interface is also crucial in understanding
614 % processes such as nucleation, crystal
615 % growth,\cite{Han92, Granasy95, Vanfleet95} and crystal
616 % melting\cite{Weber83, Han92, Sakai96, Sakai96B}. Insight gained to these
617 % properties can also be applied to biological systems of interest, such as
618 % the behavior of the antifreeze protein found in winter
619 % flounder,\cite{Wierzbicki07,Chapsky97} and certain terrestial
620 % arthropods.\cite{Duman:2001qy,Meister29012013} Elucidating the properties which
621 % give rise to these processes through experimental techniques can be expensive,
622 % complicated, and sometimes infeasible. However, through the use of molecular
623 % dynamics simulations much of the problems of investigating these properties
624 % are alleviated.
625
626 % Understanding ice/water interfaces inherently begins with the isolated
627 % systems. There has been extensive work parameterizing models for liquid water,
628 % such as the SPC\cite{Berendsen81}, SPC/E\cite{Berendsen87},
629 % TIP4P\cite{Jorgensen85}, TIP4P/2005\cite{Abascal05},
630 % ($\dots$), and more recently, models for simulating
631 % the solid phases of water, such as the TIP4P/Ice\cite{Abascal05b} model. The
632 % melting point of various crystal structures of ice have been calculated for
633 % many of these models
634 % (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}),
635 % and the partial or complete phase diagram for the model has been determined
636 % (SPC/E\cite{Baez95,Bryk04b,Sanz04b}, TIP4P\cite{Sanz04,Sanz04b,Koyama04}, TIP5P\cite{Sanz04,Koyama04}).
637 % Knowing the behavior and melting point for these models has enabled an initial
638 % investigation of ice/water interfaces.
639
640 % The Ice-I$_\mathrm{h}$/water quiescent interface has been extensively studied
641 % over the past 30 years by theory and experiment. Haymet \emph{et al.} have
642 % done significant work characterizing and quantifying the width of these
643 % interfaces for the SPC,\cite{Karim90} SPC/E,\cite{Gay02,Bryk02},
644 % CF1,\cite{Hayward01,Hayward02} and TIP4P\cite{Karim88} models for water. In
645 % recent years, Haymet has focused on investigating the effects cations and
646 % anions have on crystal nucleaion and
647 % melting.\cite{Bryk04,Smith05,Wilson08,Wilson10} Nada and Furukawa have studied
648 % the the basal- and prismatic-water interface width\cite{Nada95}, crystal
649 % surface restructuring at temperatures approaching the melting
650 % point\cite{Nada00}, and the mechanism of ice growth inhibition by antifreeze
651 % proteins\cite{Nada08,Nada11,Nada12}. Nada has developed a six-site water model
652 % for ice/water interfaces near the melting point\cite{Nada03}, and studied the
653 % dependence of crystal growth shape on applied pressure\cite{Nada11b}. Using
654 % this model, Nada and Furukawa have established differential
655 % growth rates for the basal, prismatic, and secondary prismatic facets of
656 % Ice-I$_\mathrm{h}$ and found their origins due to a reordering of the hydrogen
657 % bond network in water near the interface\cite{Nada05}. While the work
658 % described so far has mainly focused on bulk water on ice, there is significant
659 % interest in thin films of water on ice surfaces as well.
660
661 % It is well known that the surface of ice exhibits a premelting layer at
662 % temperatures near the melting point, often called a quasi-liquid layer (QLL).
663 % Molecular dynamics simulations of the facets of ice-I$_\mathrm{h}$ exposed
664 % to vacuum performed by Conde, Vega and Patrykiejew have found QLL widths of
665 % approximately 10 \AA\ at 3 K below the melting point\cite{Conde08}.
666 % Similarly, Limmer and Chandler have used course grain simulations and
667 % statistical field theory to estimated QLL widths at the same temperature to
668 % be about 3 nm\cite{Limmer14}.
669 % Recently, Sazaki and Furukawa have developed an experimental technique with
670 % sufficient spatial and temporal resolution to visulaize and quantitatively
671 % analyze QLLs on ice crystals at temperatures near melting\cite{Sazaki10}. They
672 % have found the width of the QLLs perpindicular to the surface at -2.2$^{o}$C
673 % to be $\mathcal{O}$(\AA). They have also seen the formation of two immiscible
674 % QLLs, which displayed different stabilities and dynamics on the crystal
675 % surface\cite{Sazaki12}. Knowledge of the hydrophilicities of each
676 % of the crystal facets would help further our understanding of the properties
677 % and dynamics of the QLLs.
678
679 % Presented here is the follow up to our previous paper\cite{Louden13}, in which
680 % the basal and prismatic facets of an ice-I$_\mathrm{h}$/water interface were
681 % investigated where the ice was sheared relative to the liquid. By using a
682 % recently developed velocity shearing and scaling approach to reverse
683 % non-equilibrium molecular dynamics (VSS-RNEMD), simultaneous temperature and
684 % velocity gradients can be applied to the system, which allows for measurment
685 % of friction and thermal transport properties while maintaining a stable
686 % interfacial temperature\cite{Kuang12}. Structural analysis and dynamic
687 % correlation functions were used to probe the interfacial response to a shear,
688 % and the resulting solid/liquid kinetic friction coefficients were reported.
689 % In this paper we present the same analysis for the pyramidal and secondary
690 % prismatic facets, and show that the differential interfacial friction
691 % coefficients for the four facets of ice-I$_\mathrm{h}$ are determined by their
692 % relative hydrophilicity by means of dynamics water contact angle
693 % simulations.
694
695 % The local tetrahedral order parameter, $q(z)$, is given by
696 % \begin{equation}
697 % q(z) = \int_0^L \sum_{k=1}^{N} \Bigg(1 -\frac{3}{8}\sum_{i=1}^{3}
698 % \sum_{j=i+1}^{4} \bigg(\cos\psi_{ikj}+\frac{1}{3}\bigg)^2\Bigg)
699 % \delta(z_{k}-z)\mathrm{d}z \Bigg/ N_z
700 % \label{eq:qz}
701 % \end{equation}
702 % where $\psi_{ikj}$ is the angle formed between the oxygen sites of molecules
703 % $i$,$k$, and $j$, where the centeral oxygen is located within molecule $k$ and
704 % molecules $i$ and $j$ are two of the closest four water molecules
705 % around molecule $k$. All four closest neighbors of molecule $k$ are also
706 % required to reside within the first peak of the pair distribution function
707 % for molecule $k$ (typically $<$ 3.41 \AA\ for water).
708 % $N_z = \int\delta(z_k - z) \mathrm{d}z$ is a normalization factor to account
709 % for the varying population of molecules within each finite-width bin.
710
711
712 % The hydrophobicity or hydrophilicity of a surface can be described by the
713 % extent a droplet of water wets the surface. The contact angle formed between
714 % the solid and the liquid, $\theta$, which relates the free energies of the
715 % three interfaces involved, is given by Young's equation.
716 % \begin{equation}\label{young}
717 % \cos\theta = (\gamma_{sv} - \gamma_{sl})/\gamma_{lv}
718 % \end{equation}
719 % Here $\gamma_{sv}$, $\gamma_{sl}$, and $\gamma_{lv}$ are the free energies
720 % of the solid/vapor, solid/liquid, and liquid/vapor interfaces respectively.
721 % Large contact angles ($\theta$ $\gg$ 90\textsuperscript{o}) correspond to low
722 % wettability and hydrophobic surfaces, while small contact angles
723 % ($\theta$ $\ll$ 90\textsuperscript{o}) correspond to high wettability and
724 % hydrophilic surfaces. Experimentally, measurements of the contact angle
725 % of sessile drops has been used to quantify the extent of wetting on surfaces
726 % with thermally selective wetting charactaristics\cite{Tadanaga00,Liu04,Sun04},
727 % as well as nano-pillared surfaces with electrically tunable Cassie-Baxter and
728 % Wenzel states\cite{Herbertson06,Dhindsa06,Verplanck07,Ahuja08,Manukyan11}.
729 % Luzar and coworkers have done significant work modeling these transitions on
730 % nano-patterned surfaces\cite{Daub07,Daub10,Daub11,Ritchie12}, and have found
731 % the change in contact angle to be due to the external field perturbing the
732 % hydrogen bonding of the liquid/vapor interface\cite{Daub07}.
733
734
735
736 \end{article}
737
738 \begin{figure}
739 \includegraphics[width=\linewidth]{Pyr_comic_strip}
740 \caption{\label{fig:pyrComic} The pyramidal interface with a shear
741 rate of 3.8 ms\textsuperscript{-1}. Lower panel: the local tetrahedral order
742 parameter, $q(z)$, (black circles) and the hyperbolic tangent fit (red line).
743 Middle panel: the imposed thermal gradient required to maintain a fixed
744 interfacial temperature. Upper panel: the transverse velocity gradient that
745 develops in response to an imposed momentum flux. The vertical dotted lines
746 indicate the locations of the midpoints of the two interfaces.}
747 \end{figure}
748
749 \begin{figure}
750 \includegraphics[width=\linewidth]{SP_comic_strip}
751 \caption{\label{fig:spComic} The secondary prismatic interface with a shear
752 rate of 3.5 \
753 ms\textsuperscript{-1}. Panel descriptions match those in figure \ref{fig:pyrComic}.}
754 \end{figure}
755
756 \begin{figure}
757 \includegraphics[width=\linewidth]{Pyr-orient}
758 \caption{\label{fig:PyrOrient} The three decay constants of the
759 orientational time correlation function, $C_2(t)$, for water as a function
760 of distance from the center of the ice slab. The vertical dashed line
761 indicates the edge of the pyramidal ice slab determined by the local order
762 tetrahedral parameter. The control (black circles) and sheared (red squares)
763 experiments were fit by a shifted exponential decay (Eq. 9\cite{Louden13})
764 shown by the black and red lines respectively. The upper two panels show that
765 translational and hydrogen bond making and breaking events slow down
766 through the interface while approaching the ice slab. The bottom most panel
767 shows the librational motion of the water molecules speeding up approaching
768 the ice block due to the confined region of space allowed for the molecules
769 to move in.}
770 \end{figure}
771
772 \begin{figure}
773 \includegraphics[width=\linewidth]{SP-orient-less}
774 \caption{\label{fig:SPorient} Decay constants for $C_2(t)$ at the secondary
775 prismatic face. Panel descriptions match those in \ref{fig:PyrOrient}.}
776 \end{figure}
777
778
779 \begin{table}[h]
780 \centering
781 \caption{Phyiscal properties of the basal, prismatic, pyramidal, and secondary prismatic facets of Ice-I$_\mathrm{h}$}
782 \label{tab:kappa}
783 \begin{tabular}{|ccccc|} \hline
784 & \multicolumn{2}{c}{$\kappa_{Drag direction}$
785 (x10\textsuperscript{-4} amu \AA\textsuperscript{-2} fs\textsuperscript{-1})} & & \\
786 Interface & $\kappa_{x}$ & $\kappa_{y}$ & $\theta_{\infty}$ & $K_{spread} (ns^{-1})$ \\ \hline
787 basal & $5.9 \pm 0.3$ & $6.5 \pm 0.8$ & $34.1 \pm 0.9$ & $0.60 \pm 0.07$ \\
788 pyramidal & $5.8 \pm 0.4$ & $6.1 \pm 0.5$ & $35 \pm 3$ & $0.7 \pm 0.1$ \\
789 prismatic & $3.0 \pm 0.2$ & $3.0 \pm 0.1$ & $45 \pm 3$ & $0.75 \pm 0.09$ \\
790 secondary prismatic & $3.5 \pm 0.1$ & $3.3 \pm 0.2$ & $42 \pm 2$ & $0.69 \pm 0.03$ \\ \hline
791 \end{tabular}
792 \end{table}
793
794
795 \begin{table}[h]
796 \centering
797 \caption{Shearing and Droplet simulation parameters}
798 \label{tab:method}
799 \begin{tabular}{|cccc|ccc|} \hline
800 & \multicolumn{3}{c}{Shearing} & \multicolumn{3}{c}{Droplet}\\
801 Interface & $N_{ice}$ & $N_{liquid}$ & Lx, Ly, Lz (\AA) &
802 $N_{ice}$ & $N_{droplet}$ & Lx, Ly (\AA) \\ \hline
803 Basal & 900 & 1846 & (23.87, 35.83, 98.64) & 12960 & 2048 & (134.70, 140.04)\\
804 Prismatic & 3000 & 5464 & (35.95, 35.65, 205.77) & 9900 & 2048 &
805 (110.04, 115.00)\\
806 Pyramidal & 1216 & 2203& (37.47, 29.50, 93.02) & 11136 & 2048 &
807 (143.75, 121.41)\\
808 Secondary Prismatic & 3840 & 8176 & (71.87, 31.66, 161.55) & 11520 &
809 2048 & (146.72, 124.48)\\
810 \hline
811 \end{tabular}
812 \end{table}
813
814 \end{document}