ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/iceWater2/iceWater3.tex
Revision: 4243
Committed: Tue Dec 9 20:15:55 2014 UTC (9 years, 9 months ago) by gezelter
Content type: application/x-tex
File size: 44237 byte(s)
Log Message:
Latest changes

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