ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/iceWater2/iceWater3.tex
Revision: 4234
Committed: Mon Dec 8 17:43:24 2014 UTC (9 years, 9 months ago) by plouden
Content type: application/x-tex
File size: 40624 byte(s)
Log Message:
Revised Methodology - System Construction

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 that the basal and pyramidal facets are more
60 hydrophilic than the prismatic and secondary prismatic facets.}
61 \end{abstract}
62
63 \keywords{ice|water|interface|contact angle|molecular dynamics}
64
65 \abbreviations{QLL, quasi liquid layer; MD, molecular dynamics}
66
67 %\dropcap{I}n this article we study the evolution of ``almost-sharp'' fronts
68 %for the surface quasi-geostrophic equation. This 2-D active scalar
69 %equation reads for the surface quasi-geostrophic equation.
70 %\begin{equation}
71 %\mfrac{D \theta}{Dt}=\mfrac{\pr \theta}{\pr t} + u\cdot \nabla
72 %\theta=0 \label{qg1}
73 %\end{equation}
74
75 The Ice-I$_\mathrm{h}$/water quiescent interface has been extensively studied
76 over the past 30 years by theory and experiment. Haymet \emph{et al.} have
77 done significant work characterizing and quantifying the width of these
78 interfaces for the SPC,\cite{Karim90} SPC/E,\cite{Gay02,Bryk02},
79 CF1,\cite{Hayward01,Hayward02} and TIP4P\cite{Karim88} models for water.\cite{Bryk04,Smith05,Wilson08,Wilson10} Nada and Furukawa have studied
80 the the basal- and prismatic-water interface width\cite{Nada95} and crystal
81 surface restructuring at temperatures approaching the melting
82 point\cite{Nada00}.
83
84 It is well known that the surface of ice exhibits a premelting layer at
85 temperatures near the melting point, often called a quasi-liquid layer (QLL).
86 Molecular dynamics simulations of the facets of ice-I$_\mathrm{h}$ exposed
87 to vacuum performed by Conde, Vega and Patrykiejew have found QLL widths of
88 approximately 10 \AA\ at 3 K below the melting point\cite{Conde08}.
89 Similarly, Limmer and Chandler have used course grain simulations and
90 statistical field theory to estimated QLL widths at the same temperature to
91 be about 3 nm\cite{Limmer14}.
92 Recently, Sazaki and Furukawa have developed an experimental technique with
93 sufficient spatial and temporal resolution to visulaize and quantitatively
94 analyze QLLs on ice crystals at temperatures near melting\cite{Sazaki10}. They
95 have found the width of the QLLs perpindicular to the surface at -2.2$^{o}$C
96 to be 3-4 \AA\ wide. They have also seen the formation of two immiscible
97 QLLs, which displayed different stabilities and dynamics on the crystal
98 surface\cite{Sazaki12}.
99
100 There is significant interest in the properties of ice/ice and ice/water
101 interfaces in the geophysics community. Most commonly, the results of shearing
102 two ice blocks past one
103 another\cite{Casassa91, Sukhorukov13, Pritchard12, Lishman13} or the shearing
104 of ice through water\cite{Cuffey99, Bell08}. Using molecular dynamics
105 simulations, Samadashvili has recently shown that when two smooth ice slabs
106 slide past one another, a stable liquid-like layer develops between
107 them\cite{Samadashvili13}. To fundamentally understand these processes, a
108 molecular understanding of the ice/water interfaces is needed.
109 Investigation of the ice/water interface is also crucial in understanding
110 processes such as nucleation, crystal
111 growth,\cite{Han92, Granasy95, Vanfleet95} and crystal
112 melting\cite{Weber83, Han92, Sakai96, Sakai96B}.
113
114 In a previous study\cite{Louden13}, we investigated
115 the basal and prismatic facets of an ice-I$_\mathrm{h}$/water
116 interface where the ice was sheared relative to the liquid. Using
117 velocity shearing and scaling approach to reverse
118 non-equilibrium molecular dynamics (VSS-RNEMD), simultaneous temperature and
119 velocity gradients were applied to the system, allowing for measurment
120 of friction and thermal transport properties while maintaining a stable
121 interfacial temperature\cite{Kuang12}.
122
123 Paragraph here about hydrophobicity and hydrophilicity, maybe move up
124 more in the paper as well. Talk about physically what it means for a
125 surface to by hydrophobic or hydrophilic, and then we move into
126 how do we define it (mathematically) and then measure the degree
127 of wetting experimentally and theoretically.
128
129 The hydrophobicity or hydrophilicity of a surface can be described by the
130 extent a droplet of water wets the surface. The contact angle formed between
131 the solid and the liquid, $\theta$, which relates the free energies of the
132 three interfaces involved, is given by Young's equation.
133 \begin{equation}\label{young}
134 \cos\theta = (\gamma_{sv} - \gamma_{sl})/\gamma_{lv}
135 \end{equation}
136 Here $\gamma_{sv}$, $\gamma_{sl}$, and $\gamma_{lv}$ are the free energies
137 of the solid/vapor, solid/liquid, and liquid/vapor interfaces respectively.
138 Large contact angles ($\theta$ $\gg$ 90\textsuperscript{o}) correspond to low
139 wettability and hydrophobic surfaces, while small contact angles
140 ($\theta$ $\ll$ 90\textsuperscript{o}) correspond to high wettability and
141 hydrophilic surfaces. Experimentally, measurements of the contact angle
142 of sessile drops has been used to quantify the extent of wetting on surfaces
143 with thermally selective wetting charactaristics\cite{Tadanaga00,Liu04,Sun04},
144 as well as nano-pillared surfaces with electrically tunable Cassie-Baxter and
145 Wenzel states\cite{Herbertson06,Dhindsa06,Verplanck07,Ahuja08,Manukyan11}.
146 Luzar and coworkers have done significant work modeling these transitions on
147 nano-patterned surfaces\cite{Daub07,Daub10,Daub11,Ritchie12}, and have found
148 the change in contact angle to be due to the external field perturbing the
149 hydrogen bonding of the liquid/vapor interface\cite{Daub07}.
150
151 The resulting solid/liquid kinetic friction coefficients were
152 reported, and displayed a factor of two difference between the
153 basal and prismatic facets.
154 In this paper we present the same analysis for the pyramidal and secondary
155 prismatic facets, and show that the differential interfacial friction
156 coefficients for the four facets of ice-I$_\mathrm{h}$ are determined by their
157 relative hydrophilicity by means of dynamics water contact angle simulations.
158
159 \section{Methodology}
160
161 \subsection{Water Model}
162 Understanding ice/water interfaces inherently begins with the isolated
163 systems. There has been extensive work parameterizing models for liquid water,
164 such as the SPC\cite{Berendsen81}, SPC/E\cite{Berendsen87},
165 TIP4P\cite{Jorgensen85}, TIP4P/2005\cite{Abascal05},
166 ($\dots$), and more recently, models for simulating
167 the solid phases of water, such as the TIP4P/Ice\cite{Abascal05b} model. The
168 melting point of various crystal structures of ice have been calculated for
169 many of these models
170 (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}),
171 and the partial or complete phase diagram for the model has been determined
172 (SPC/E\cite{Baez95,Bryk04b,Sanz04b}, TIP4P\cite{Sanz04,Sanz04b,Koyama04}, TIP5P\cite{Sanz04,Koyama04}).
173
174 Haymet et al. have studied the quiescent Ice-I$_\mathrm{h}$/water interface
175 using the rigid SPC, SPC/E, TIP4P, and the flexible CF1 water models, and has seen good
176 agreement for structural and dynamic measurements of the interfacial
177 width. Given the expansive size of our systems of interest, and the
178 apparent independence of water model on interfacial width, we have chosen to use the rigid SPC/E
179 water model in this study.
180
181 \subsection{Pyramidal and secondary prismatic system construction}
182 To construct the pyramidal and secondary prismatic ice/water systems,
183 first a proton-ordered zero dipole crystal of ice-I$_\mathrm{h}$ with exposed strips
184 of H-atoms and lone pairs was constructed from Structure 6 of Hirsch
185 and Ojam\"{a}e's recent paper\cite{Hirsch04}. The crystal was then cut
186 along the plane of the desired facet, and reoriented so that the
187 $z$-axis was perpdicular to the exposed face. Two orthoganol cuts were
188 then made to the crystal such that perfect periodic replication could
189 be perfromed in the $x$ and $y$ dimensions. The slab was then
190 replicated along the $x$ and $y$ axes until the desired crystal size
191 was obtained. Liquid water boxes were created having identical
192 dimensions (in $x$ and$y$) as the ice blocks, and a $z$ dimension of
193 three times that of the ice block. Each of the ice slabs and water
194 boxes were independently equilibrated to 50K, and the resulting
195 systems were merged by carving out any liquid water molecules within 3
196 \AA\ of any atoms in the ice slabs. For a more detailed explanation of
197 the ice/water systems construction, please refer to a previous
198 paper\cite{Louden13}. The resulting dimensions, number of ice, and liquid water molecules
199 contained in each of these systems can be seen in Table \ref{tab:method}.
200 \subsection{Shearing simulations}
201 % Do we need to justify the sims at 225K?
202 % No crystal growth or shrinkage over 2 successive 1 ns NVT simulations for
203 % either the pyramidal or sec. prismatic ice/water systems.
204
205 The computational details performed here were equivalent to those reported
206 in our previous publication\cite{Louden13}. The only changes made to the
207 previously reported procedure were the following. VSS-RNEMD moves were
208 attempted every 2 fs instead of every 50 fs. This was done to minimize
209 the magnitude of each individual VSS-RNEMD perturbation to the system.
210
211 All pyramidal simulations were performed under the canonical (NVT) ensamble
212 except those
213 during which statistics were accumulated for the orientational correlation
214 function, which were performed under the microcanonical (NVE) ensamble. All
215 secondary prismatic
216 simulations were performed under the NVE ensamble.
217
218 \subsection{Droplet simulations}
219 Here, we will calculate the contact angle of a water droplet as it spreads
220 across each of the four ice I$_\mathrm{h}$ crystal facets in order to
221 determine the surface's relative hydrophilicites. The ice surfaces were
222 oriented so that the desired facet was exposed to the positive z dimension.
223 The sizes and number of molecules in each of the surfaces is given in Table
224 \ref{tab:ice_sheets}. Molecular restraints were applied to the center of mass
225 of the rigid bodies to prevent surface melting, however the molecules were
226 allowed to reorient themselves freely. The water doplet to be placed on the
227 surface contained 2048 SPC/E molecules, which has been found to produce
228 agreement for the Young contact angle extrapolated to an infinite drop
229 size\cite{Daub10}. The surfaces and droplet were equilibrated to 225 K, at
230 which time the droplet was placed 3-5 \AA\ above the surface at 5 unique
231 locations. Each simulation was 5 ns in length and conducted in the NVE
232 ensemble.
233
234
235 \section{Results and discussion}
236 \subsection{Interfacial width}
237 In the literature there is good agreement that between the solid ice and
238 the bulk water, there exists a region of 'slush-like' water molecules.
239 In this region, the water molecules are structurely distinguishable and
240 behave differently than those of the solid ice or the bulk water.
241 The characteristics of this region have been defined by both structural
242 and dynamic properties; and its width has been measured by the change of these
243 properties from their bulk liquid values to those of the solid ice.
244 Examples of these properties include the density, the diffusion constant, and
245 the translational order profile. \cite{Bryk02,Karim90,Gay02,Hayward01,Hayward02,Karim88}
246
247 Since the VSS-RNEMD moves used to impose the thermal and velocity gradients
248 perturb the momenta of the water molecules in
249 the systems, parameters that depend on translational motion may give
250 faulty results. A stuructural parameter will be less effected by the
251 VSS-RNEMD perturbations to the system. Due to this, we have used the
252 local tetrahedral order parameter (Eq 5\cite{Louden13} to quantify the width of the interface,
253 which was originally described by Kumar\cite{Kumar09} and
254 Errington\cite{Errington01}, and used by Bryk and Haymet in a previous study
255 of ice/water interfaces.\cite{Bryk04b}
256
257 To determine the width of the interfaces, each of the systems were
258 divided into 100 artificial bins along the
259 $z$-dimension, and the local tetrahedral order parameter, $q(z)$, was
260 time-averaged for each of the bins, resulting in a tetrahedrality profile of
261 the system. These profiles are shown across the $z$-dimension of the systems
262 in panel $a$ of Figures \ref{fig:pyrComic}
263 and \ref{fig:spComic} (black circles). The $q(z)$ function has a range of
264 (0,1), where a larger value indicates a more tetrahedral environment.
265 The $q(z)$ for the bulk liquid was found to be $\approx $ 0.77, while values of
266 $\approx $ 0.92 were more common for the ice. The tetrahedrality profiles were
267 fit using a hyperbolic tangent\cite{Louden13} designed to smoothly fit the
268 bulk to ice
269 transition, while accounting for the thermal influence on the profile by the
270 kinetic energy exchanges of the VSS-RNEMD moves. In panels $b$ and $c$, the
271 resulting thermal and velocity gradients from an imposed kinetic energy and
272 momentum fluxes can be seen. The verticle dotted
273 lines traversing all three panels indicate the midpoints of the interface
274 as determined by the hyperbolic tangent fit of the tetrahedrality profiles.
275
276 From fitting the tetrahedrality profiles for each of the 0.5 nanosecond
277 simulations (panel c of Figures \ref{fig:pyrComic} and \ref{fig:spComic})
278 by eq. 6\cite{Louden13},we find the interfacial width to be
279 3.2 $\pm$ 0.2 and 3.2 $\pm$ 0.2 \AA\ for the control system with no applied
280 momentum flux for both the pyramidal and secondary prismatic systems.
281 Over the range of shear rates investigated,
282 0.6 $\pm$ 0.2 $\mathrm{ms}^{-1} \rightarrow$ 5.6 $\pm$ 0.4 $\mathrm{ms}^{-1}$
283 for the pyramidal system and 0.9 $\pm$ 0.3 $\mathrm{ms}^{-1} \rightarrow$ 5.4
284 $\pm$ 0.1 $\mathrm{ms}^{-1}$ for the secondary prismatic, we found no
285 significant change in the interfacial width. This follows our previous
286 findings of the basal and
287 prismatic systems, in which the interfacial width was invarient of the
288 shear rate of the ice. The interfacial width of the quiescent basal and
289 prismatic systems was found to be 3.2 $\pm$ 0.4 \AA\ and 3.6 $\pm$ 0.2 \AA\
290 respectively, over the range of shear rates investigated, 0.6 $\pm$ 0.3
291 $\mathrm{ms}^{-1} \rightarrow$ 5.3 $\pm$ 0.5 $\mathrm{ms}^{-1}$ for the basal
292 system and 0.9 $\pm$ 0.2 $\mathrm{ms}^{-1} \rightarrow$ 4.5 $\pm$ 0.1
293 $\mathrm{ms}^{-1}$ for the prismatic.
294
295 These results indicate that the surface structure of the exposed ice crystal
296 has little to no effect on how far into the bulk the ice-like structural
297 ordering is. Also, it appears that the interface is not structurally effected
298 by shearing the ice through water.
299
300
301 \subsection{Orientational dynamics}
302 %Should we include the math here?
303 The orientational time correlation function,
304 \begin{equation}\label{C(t)1}
305 C_{2}(t)=\langle P_{2}(\mathbf{u}(0)\cdot \mathbf{u}(t))\rangle,
306 \end{equation}
307 helps indicate the local environment around the water molecules. The function
308 begins with an initial value of unity, and decays to zero as the water molecule
309 loses memory of its former orientation. Observing the rate at which this decay
310 occurs can provide insight to the mechanism and timescales for the relaxation.
311 In eq. \eqref{C(t)1}, $P_{2}$ is the second-order Legendre polynomial, and
312 $\mathbf{u}$ is the bisecting HOH vector. The angle brackets indicate
313 an ensemble average over all the water molecules in a given spatial region.
314
315 To investigate the dynamics of the water molecules across the interface, the
316 systems were divided in the $z$-dimension into bins, each $\approx$ 3 \AA\
317 wide, and eq. \eqref{C(t)1} was computed for each of the bins. A water
318 molecule was allocated to a particular bin if it was initially in the bin
319 at time zero. To compute eq. \eqref{C(t)1}, each 0.5 ns simulation was
320 followed by an additional 200 ps NVE simulation during which the
321 position and orientations of each molecule were recorded every 0.1 ps.
322
323 The data obtained for each bin was then fit to a triexponential decay
324 with the three decay constants
325 $\tau_{short}$ corresponding to the librational motion of the water
326 molecules, $\tau_{middle}$ corresponding to jumps between the breaking and
327 making of hydrogen bonds, and $\tau_{long}$ corresponding to the translational
328 motion of the water molecules. An additive constant in the fit accounts
329 for the water molecules trapped in the ice which do not experience any
330 long-time orientational decay.
331
332 In Figures \ref{fig:PyrOrient} and \ref{fig:SPorient} we see the $z$-coordinate
333 profiles for the three decay constants, $\tau_{short}$ (panel a),
334 $\tau_{middle}$ (panel b),
335 and $\tau_{long}$ (panel c) for the pyramidal and secondary prismatic systems
336 respectively. The control experiments (no shear) are shown in black, and
337 an experiment with an imposed momentum flux is shown in red. The vertical
338 dotted line traversing all three panels denotes the midpoint of the
339 interface as determined by the local tetrahedral order parameter fitting.
340 In the liquid regions of both systems, we see that $\tau_{middle}$ and
341 $\tau_{long}$ have approximately consistent values of $3-6$ ps and $30-40$ ps,
342 resepctively, and increase in value as we approach the interface. Conversely,
343 in panel a, we see that $\tau_{short}$ decreases from the liquid value
344 of $72-76$ fs as we approach the interface. We believe this speed up is due to
345 the constrained motion of librations closer to the interface. Both the
346 approximate values for the decays and trends approaching the interface match
347 those reported previously for the basal and prismatic interfaces.
348
349 As done previously, we have attempted to quantify the distance, $d_{pyramidal}$
350 and $d_{secondary prismatic}$, from the
351 interface that the deviations from the bulk liquid values begin. This was done
352 by fitting the orientational decay constant $z$-profiles by
353 \begin{equation}\label{tauFit}
354 \tau(z)\approx\tau_{liquid}+(\tau_{wall}-\tau_{liquid})e^{-(z-z_{wall})/d}
355 \end{equation}
356 where $\tau_{liquid}$ and $\tau_{wall}$ are the liquid and projected wall
357 values of the decay constants, $z_{wall}$ is the location of the interface,
358 and $d$ is the displacement from the interface at which these deviations
359 occur. The values for $d_{pyramidal}$ and $d_{secondary prismatic}$ were
360 determined
361 for each of the decay constants, and then averaged for better statistics
362 ($\tau_{middle}$ was ommitted for secondary prismatic). For the pyramidal
363 system,
364 $d_{pyramidal}$ was found to be 2.7 \AA\ for both the control and the sheared
365 system. We found $d_{secondary prismatic}$ to be slightly larger than
366 $d_{pyramidal}$ for both the control and with an applied shear, with
367 displacements of $4$ \AA\ for the control system and $3$ \AA\ for the
368 experiment with the imposed momentum flux. These values are consistent with
369 those found for the basal ($d_{basal}\approx2.9$ \AA\ ) and prismatic
370 ($d_{prismatic}\approx3.5$ \AA\ ) systems.
371
372 \subsection{Coefficient of friction of the interfaces}
373 While investigating the kinetic coefficient of friction, there was found
374 to be a dependence for $\mu_k$
375 on the temperature of the liquid water in the system. We believe this
376 dependence
377 arrises from the sharp discontinuity of the viscosity for the SPC/E model
378 at temperatures approaching 200 K\cite{Kuang12}. Due to this, we propose
379 a weighting to the interfacial friction coefficient, $\kappa$ by the
380 shear viscosity of the fluid at 225 K. The interfacial friction coefficient
381 relates the shear stress with the relative velocity of the fluid normal to the
382 interface:
383 \begin{equation}\label{Shenyu-13}
384 j_{z}(p_{x}) = \kappa[v_{x}(fluid)-v_{x}(solid)]
385 \end{equation}
386 where $j_{z}(p_{x})$ is the applied momentum flux (shear stress) across $z$
387 in the
388 $x$-dimension, and $v_{x}$(fluid) and $v_{x}$(solid) are the velocities
389 directly adjacent to the interface. The shear viscosity, $\eta(T)$, of the
390 fluid can be determined under a linear response of the momentum
391 gradient to the applied shear stress by
392 \begin{equation}\label{Shenyu-11}
393 j_{z}(p_{x}) = \eta(T) \frac{\partial v_{x}}{\partial z}.
394 \end{equation}
395 Using eqs \eqref{Shenyu-13} and \eqref{Shenyu-11}, we can find the following
396 expression for $\kappa$,
397 \begin{equation}\label{kappa-1}
398 \kappa = \eta(T) \frac{\partial v_{x}}{\partial z}\frac{1}{[v_{x}(fluid)-v_{x}(solid)]}.
399 \end{equation}
400 Here is where we will introduce the weighting term of $\eta(225)/\eta(T)$
401 giving us
402 \begin{equation}\label{kappa-2}
403 \kappa = \frac{\eta(225)}{[v_{x}(fluid)-v_{x}(solid)]}\frac{\partial v_{x}}{\partial z}.
404 \end{equation}
405
406 To obtain the value of $\eta(225)$ for the SPC/E model, a $31.09 \times 29.38
407 \times 124.39$ \AA\ box with 3744 SPC/E liquid water molecules was
408 equilibrated to 225K,
409 and 5 unique shearing experiments were performed. Each experiment was
410 conducted in the NVE and were 5 ns in
411 length. The VSS were attempted every timestep, which was set to 2 fs.
412 For our SPC/E systems, we found $\eta(225)$ to be 0.0148 $\pm$ 0.0007 Pa s,
413 roughly ten times larger than the value found for 280 K SPC/E bulk water by
414 Kuang\cite{Kuang12}.
415
416 The interfacial friction coefficient, $\kappa$, can equivalently be expressed
417 as the ratio of the viscosity of the fluid to the slip length, $\delta$, which
418 is an indication of how 'slippery' the interface is.
419 \begin{equation}\label{kappa-3}
420 \kappa = \frac{\eta}{\delta}
421 \end{equation}
422 In each of the systems, the interfacial temperature was kept fixed to 225K,
423 which ensured the viscosity of the fluid at the
424 interace was approximately the same. Thus, any significant variation in
425 $\kappa$ between
426 the systems indicates differences in the 'slipperiness' of the interfaces.
427 As each of the ice systems are sheared relative to liquid water, the
428 'slipperiness' of the interface can be taken as an indication of how
429 hydrophobic or hydrophilic the interface is. The calculated $\kappa$ values
430 found for the four crystal facets of Ice-I$_\mathrm{h}$ investigated are shown
431 in Table \ref{tab:kappa}. The basal and pyramidal facets were found to have
432 similar values of $\kappa \approx$ 0.0006
433 (amu \AA\textsuperscript{-2} fs\textsuperscript{-1}), while values of
434 $\kappa \approx$ 0.0003 (amu \AA\textsuperscript{-2} fs\textsuperscript{-1})
435 were found for the prismatic and secondary prismatic systems.
436 These results indicate that the basal and pyramidal facets are
437 more hydrophilic than the prismatic and secondary prismatic facets.
438
439 \subsection{Dynamic water contact angle}
440
441
442
443
444 \section{Conclusion}
445 We present the results of molecular dynamics simulations of the pyrmaidal
446 and secondary prismatic facets of an SPC/E model of the
447 Ice-I$_\mathrm{h}$/water interface. The ice was sheared through the liquid
448 water while being exposed to a thermal gradient to maintain a stable
449 interface by using the minimally perturbing VSS RNEMD method. In agreement
450 with our previous findings for the basal and prismatic facets, the interfacial
451 width was found to be independent of shear rate as measured by the local
452 tetrahedral order parameter. This width was found to be
453 3.2~$\pm$ 0.2~\AA\ for both the pyramidal and the secondary prismatic systems.
454 These values are in good agreement with our previously calculated interfacial
455 widths for the basal (3.2~$\pm$ 0.4~\AA\ ) and prismatic (3.6~$\pm$ 0.2~\AA\ )
456 systems.
457
458 Orientational dynamics of the Ice-I$_\mathrm{h}$/water interfaces were studied
459 by calculation of the orientational time correlation function at varying
460 displacements normal to the interface. The decays were fit
461 to a tri-exponential decay, where the three decay constants correspond to
462 the librational motion of the molecules driven by the restoring forces of
463 existing hydrogen bonds ($\tau_{short}$ $\mathcal{O}$(10 fs)), jumps between
464 two different hydrogen bonds ($\tau_{middle}$ $\mathcal{O}$(1 ps)), and
465 translational motion of the molecules ($\tau_{long}$ $\mathcal{O}$(100 ps)).
466 $\tau_{short}$ was found to decrease approaching the interface due to the
467 constrained motion of the molecules as the local environment becomes more
468 ice-like. Conversely, the two longer-time decay constants were found to
469 increase at small displacements from the interface. As seen in our previous
470 work on the basal and prismatic facets, there appears to be a dynamic
471 interface width at which deviations from the bulk liquid values occur.
472 We had previously found $d_{basal}$ and $d_{prismatic}$ to be approximately
473 2.8~\AA\ and 3.5~\AA. We found good agreement of these values for the
474 pyramidal and secondary prismatic systems with $d_{pyramidal}$ and
475 $d_{secondary prismatic}$ to be 2.7~\AA\ and 3~\AA. For all four of the
476 facets, no apparent dependence of the dynamic width on the shear rate was
477 found.
478
479 %Paragraph summarizing the Kappa values
480 The interfacial friction coefficient, $\kappa$, was determined for each facet
481 interface. We were able to reach an expression for $\kappa$ as a function of
482 the velocity profile of the system which is scaled by the viscosity of the liquid
483 at 225 K. In doing so, we have obtained an expression for $\kappa$ which is
484 independent of temperature differences of the liquid water at far displacements
485 from the interface. We found the basal and pyramidal facets to have
486 similar $\kappa$ values, of $\kappa \approx$ 0.0006
487 (amu \AA\textsuperscript{-2} fs\textsuperscript{-1}). However, the
488 prismatic and secondary prismatic facets were found to have $\kappa$ values of
489 $\kappa \approx$ 0.0003 (amu \AA\textsuperscript{-2} fs\textsuperscript{-1}).
490 As these ice facets are being sheared relative to liquid water, with the
491 structural and dynamic width of all four
492 interfaces being approximately the same, the difference in the coefficient of
493 friction indicates the hydrophilicity of the crystal facets are not
494 equivalent. Namely, that the basal and pyramidal facets of Ice-I$_\mathrm{h}$
495 are more hydrophilic than the prismatic and secondary prismatic facets.
496
497
498 \begin{acknowledgments}
499 Support for this project was provided by the National
500 Science Foundation under grant CHE-1362211. Computational time was
501 provided by the Center for Research Computing (CRC) at the
502 University of Notre Dame.
503 \end{acknowledgments}
504
505 \newpage
506
507 \bibliography{iceWater.bib}
508
509 There is significant interest in the properties of ice/ice and ice/water
510 interfaces in the geophysics community. Most commonly, the results of shearing
511 two ice blocks past one
512 another\cite{Casassa91, Sukhorukov13, Pritchard12, Lishman13} or the shearing
513 of ice through water\cite{Cuffey99, Bell08}. Using molecular dynamics
514 simulations, Samadashvili has recently shown that when two smooth ice slabs
515 slide past one another, a stable liquid-like layer develops between
516 them\cite{Samadashvili13}. To fundamentally understand these processes, a
517 molecular understanding of the ice/water interfaces is needed.
518
519 Investigation of the ice/water interface is also crucial in understanding
520 processes such as nucleation, crystal
521 growth,\cite{Han92, Granasy95, Vanfleet95} and crystal
522 melting\cite{Weber83, Han92, Sakai96, Sakai96B}. Insight gained to these
523 properties can also be applied to biological systems of interest, such as
524 the behavior of the antifreeze protein found in winter
525 flounder,\cite{Wierzbicki07,Chapsky97} and certain terrestial
526 arthropods.\cite{Duman:2001qy,Meister29012013} Elucidating the properties which
527 give rise to these processes through experimental techniques can be expensive,
528 complicated, and sometimes infeasible. However, through the use of molecular
529 dynamics simulations much of the problems of investigating these properties
530 are alleviated.
531
532 Understanding ice/water interfaces inherently begins with the isolated
533 systems. There has been extensive work parameterizing models for liquid water,
534 such as the SPC\cite{Berendsen81}, SPC/E\cite{Berendsen87},
535 TIP4P\cite{Jorgensen85}, TIP4P/2005\cite{Abascal05},
536 ($\dots$), and more recently, models for simulating
537 the solid phases of water, such as the TIP4P/Ice\cite{Abascal05b} model. The
538 melting point of various crystal structures of ice have been calculated for
539 many of these models
540 (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}),
541 and the partial or complete phase diagram for the model has been determined
542 (SPC/E\cite{Baez95,Bryk04b,Sanz04b}, TIP4P\cite{Sanz04,Sanz04b,Koyama04}, TIP5P\cite{Sanz04,Koyama04}).
543 Knowing the behavior and melting point for these models has enabled an initial
544 investigation of ice/water interfaces.
545
546 The Ice-I$_\mathrm{h}$/water quiescent interface has been extensively studied
547 over the past 30 years by theory and experiment. Haymet \emph{et al.} have
548 done significant work characterizing and quantifying the width of these
549 interfaces for the SPC,\cite{Karim90} SPC/E,\cite{Gay02,Bryk02},
550 CF1,\cite{Hayward01,Hayward02} and TIP4P\cite{Karim88} models for water. In
551 recent years, Haymet has focused on investigating the effects cations and
552 anions have on crystal nucleaion and
553 melting.\cite{Bryk04,Smith05,Wilson08,Wilson10} Nada and Furukawa have studied
554 the the basal- and prismatic-water interface width\cite{Nada95}, crystal
555 surface restructuring at temperatures approaching the melting
556 point\cite{Nada00}, and the mechanism of ice growth inhibition by antifreeze
557 proteins\cite{Nada08,Nada11,Nada12}. Nada has developed a six-site water model
558 for ice/water interfaces near the melting point\cite{Nada03}, and studied the
559 dependence of crystal growth shape on applied pressure\cite{Nada11b}. Using
560 this model, Nada and Furukawa have established differential
561 growth rates for the basal, prismatic, and secondary prismatic facets of
562 Ice-I$_\mathrm{h}$ and found their origins due to a reordering of the hydrogen
563 bond network in water near the interface\cite{Nada05}. While the work
564 described so far has mainly focused on bulk water on ice, there is significant
565 interest in thin films of water on ice surfaces as well.
566
567 It is well known that the surface of ice exhibits a premelting layer at
568 temperatures near the melting point, often called a quasi-liquid layer (QLL).
569 Molecular dynamics simulations of the facets of ice-I$_\mathrm{h}$ exposed
570 to vacuum performed by Conde, Vega and Patrykiejew have found QLL widths of
571 approximately 10 \AA\ at 3 K below the melting point\cite{Conde08}.
572 Similarly, Limmer and Chandler have used course grain simulations and
573 statistical field theory to estimated QLL widths at the same temperature to
574 be about 3 nm\cite{Limmer14}.
575 Recently, Sazaki and Furukawa have developed an experimental technique with
576 sufficient spatial and temporal resolution to visulaize and quantitatively
577 analyze QLLs on ice crystals at temperatures near melting\cite{Sazaki10}. They
578 have found the width of the QLLs perpindicular to the surface at -2.2$^{o}$C
579 to be $\mathcal{O}$(\AA). They have also seen the formation of two immiscible
580 QLLs, which displayed different stabilities and dynamics on the crystal
581 surface\cite{Sazaki12}. Knowledge of the hydrophilicities of each
582 of the crystal facets would help further our understanding of the properties
583 and dynamics of the QLLs.
584
585 Presented here is the follow up to our previous paper\cite{Louden13}, in which
586 the basal and prismatic facets of an ice-I$_\mathrm{h}$/water interface were
587 investigated where the ice was sheared relative to the liquid. By using a
588 recently developed velocity shearing and scaling approach to reverse
589 non-equilibrium molecular dynamics (VSS-RNEMD), simultaneous temperature and
590 velocity gradients can be applied to the system, which allows for measurment
591 of friction and thermal transport properties while maintaining a stable
592 interfacial temperature\cite{Kuang12}. Structural analysis and dynamic
593 correlation functions were used to probe the interfacial response to a shear,
594 and the resulting solid/liquid kinetic friction coefficients were reported.
595 In this paper we present the same analysis for the pyramidal and secondary
596 prismatic facets, and show that the differential interfacial friction
597 coefficients for the four facets of ice-I$_\mathrm{h}$ are determined by their
598 relative hydrophilicity by means of dynamics water contact angle
599 simulations.
600
601 The local tetrahedral order parameter, $q(z)$, is given by
602 \begin{equation}
603 q(z) = \int_0^L \sum_{k=1}^{N} \Bigg(1 -\frac{3}{8}\sum_{i=1}^{3}
604 \sum_{j=i+1}^{4} \bigg(\cos\psi_{ikj}+\frac{1}{3}\bigg)^2\Bigg)
605 \delta(z_{k}-z)\mathrm{d}z \Bigg/ N_z
606 \label{eq:qz}
607 \end{equation}
608 where $\psi_{ikj}$ is the angle formed between the oxygen sites of molecules
609 $i$,$k$, and $j$, where the centeral oxygen is located within molecule $k$ and
610 molecules $i$ and $j$ are two of the closest four water molecules
611 around molecule $k$. All four closest neighbors of molecule $k$ are also
612 required to reside within the first peak of the pair distribution function
613 for molecule $k$ (typically $<$ 3.41 \AA\ for water).
614 $N_z = \int\delta(z_k - z) \mathrm{d}z$ is a normalization factor to account
615 for the varying population of molecules within each finite-width bin.
616
617
618 The hydrophobicity or hydrophilicity of a surface can be described by the
619 extent a droplet of water wets the surface. The contact angle formed between
620 the solid and the liquid, $\theta$, which relates the free energies of the
621 three interfaces involved, is given by Young's equation.
622 \begin{equation}\label{young}
623 \cos\theta = (\gamma_{sv} - \gamma_{sl})/\gamma_{lv}
624 \end{equation}
625 Here $\gamma_{sv}$, $\gamma_{sl}$, and $\gamma_{lv}$ are the free energies
626 of the solid/vapor, solid/liquid, and liquid/vapor interfaces respectively.
627 Large contact angles ($\theta$ $\gg$ 90\textsuperscript{o}) correspond to low
628 wettability and hydrophobic surfaces, while small contact angles
629 ($\theta$ $\ll$ 90\textsuperscript{o}) correspond to high wettability and
630 hydrophilic surfaces. Experimentally, measurements of the contact angle
631 of sessile drops has been used to quantify the extent of wetting on surfaces
632 with thermally selective wetting charactaristics\cite{Tadanaga00,Liu04,Sun04},
633 as well as nano-pillared surfaces with electrically tunable Cassie-Baxter and
634 Wenzel states\cite{Herbertson06,Dhindsa06,Verplanck07,Ahuja08,Manukyan11}.
635 Luzar and coworkers have done significant work modeling these transitions on
636 nano-patterned surfaces\cite{Daub07,Daub10,Daub11,Ritchie12}, and have found
637 the change in contact angle to be due to the external field perturbing the
638 hydrogen bonding of the liquid/vapor interface\cite{Daub07}.
639
640
641
642 \end{article}
643
644 \begin{figure}
645 \includegraphics[width=\linewidth]{Pyr_comic_strip}
646 \caption{\label{fig:pyrComic} The pyramidal interface with a shear
647 rate of 3.8 ms\textsuperscript{-1}. Lower panel: the local tetrahedral order
648 parameter, $q(z)$, (black circles) and the hyperbolic tangent fit (red line).
649 Middle panel: the imposed thermal gradient required to maintain a fixed
650 interfacial temperature. Upper panel: the transverse velocity gradient that
651 develops in response to an imposed momentum flux. The vertical dotted lines
652 indicate the locations of the midpoints of the two interfaces.}
653 \end{figure}
654
655 \begin{figure}
656 \includegraphics[width=\linewidth]{SP_comic_strip}
657 \caption{\label{fig:spComic} The secondary prismatic interface with a shear
658 rate of 3.5 \
659 ms\textsuperscript{-1}. Panel descriptions match those in figure \ref{fig:pyrComic}.}
660 \end{figure}
661
662 \begin{figure}
663 \includegraphics[width=\linewidth]{Pyr-orient}
664 \caption{\label{fig:PyrOrient} The three decay constants of the
665 orientational time correlation function, $C_2(t)$, for water as a function
666 of distance from the center of the ice slab. The vertical dashed line
667 indicates the edge of the pyramidal ice slab determined by the local order
668 tetrahedral parameter. The control (black circles) and sheared (red squares)
669 experiments were fit by a shifted exponential decay (Eq. 9\cite{Louden13})
670 shown by the black and red lines respectively. The upper two panels show that
671 translational and hydrogen bond making and breaking events slow down
672 through the interface while approaching the ice slab. The bottom most panel
673 shows the librational motion of the water molecules speeding up approaching
674 the ice block due to the confined region of space allowed for the molecules
675 to move in.}
676 \end{figure}
677
678 \begin{figure}
679 \includegraphics[width=\linewidth]{SP-orient-less}
680 \caption{\label{fig:SPorient} Decay constants for $C_2(t)$ at the secondary
681 prismatic face. Panel descriptions match those in \ref{fig:PyrOrient}.}
682 \end{figure}
683
684
685 \begin{table}[h]
686 \centering
687 \caption{Phyiscal properties of the basal, prismatic, pyramidal, and secondary prismatic facets of Ice-I$_\mathrm{h}$}
688 \label{tab:kappa}
689 \begin{tabular}{|ccccc|} \hline
690 & \multicolumn{2}{c}{$\kappa_{Drag direction}$
691 (x10\textsuperscript{-4} amu \AA\textsuperscript{-2} fs\textsuperscript{-1})} & & \\
692 Interface & $\kappa_{x}$ & $\kappa_{y}$ & $\theta_{\infty}$ & $K_{spread} (ns^{-1})$ \\ \hline
693 basal & $5.9 \pm 0.3$ & $6.5 \pm 0.8$ & $34.1 \pm 0.9$ & $0.60 \pm 0.07$ \\
694 pyramidal & $5.8 \pm 0.4$ & $6.1 \pm 0.5$ & $35 \pm 3$ & $0.7 \pm 0.1$ \\
695 prismatic & $3.0 \pm 0.2$ & $3.0 \pm 0.1$ & $45 \pm 3$ & $0.75 \pm 0.09$ \\
696 secondary prismatic & $3.5 \pm 0.1$ & $3.3 \pm 0.2$ & $42 \pm 2$ & $0.69 \pm 0.03$ \\ \hline
697 \end{tabular}
698 \end{table}
699
700
701 \begin{table}[h]
702 \centering
703 \caption{Shearing and Droplet simulation parameters}
704 \label{tab:method}
705 \begin{tabular}{|cccc|ccc|} \hline
706 & \multicolumn{3}{c}{Shearing} & \multicolumn{3}{c}{Droplet}\\
707 Interface & $N_{ice}$ & $N_{liquid}$ & Lx, Ly, Lz (\AA) &
708 $N_{ice}$ & $N_{droplet}$ & Lx, Ly (\AA) \\ \hline
709 Basal & 900 & 1846 & (23.87, 35.83, 98.64) & 12960 & 2048 & (134.70, 140.04)\\
710 Prismatic & 3000 & 5464 & (35.95, 35.65, 205.77) & 9900 & 2048 &
711 (110.04, 115.00)\\
712 Pyramidal & 1216 & 2203& (37.47, 29.50, 93.02) & 11136 & 2048 &
713 (143.75, 121.41)\\
714 Secondary Prismatic & 3840 & 8176 & (71.87, 31.66, 161.55) & 11520 &
715 2048 & (146.72, 124.48)\\
716 \hline
717 \end{tabular}
718 \end{table}
719
720 \end{document}