ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/iceWater/iceWater.tex
Revision: 3950
Committed: Mon Sep 9 12:45:25 2013 UTC (10 years, 11 months ago) by plouden
Content type: application/x-tex
File size: 34353 byte(s)
Log Message:
making sure revisions sent

File Contents

# Content
1 \documentclass[journal = jpccck, manuscript = article]{achemso}
2 \setkeys{acs}{usetitle = true}
3 \usepackage{achemso}
4 \usepackage{natbib}
5 \usepackage{multirow}
6 \usepackage{wrapfig}
7 \usepackage{fixltx2e}
8 %\mciteErrorOnUnknownfalse
9
10 \usepackage[version=3]{mhchem} % this is a great package for formatting chemical reactions
11 \usepackage{url}
12
13 \title{Simulations of solid-liquid friction at ice-I$_\mathrm{h}$ /
14 water interfaces}
15
16 \author{P. B. Louden}
17 \author{J. Daniel Gezelter}
18 \email{gezelter@nd.edu}
19 \affiliation[University of Notre Dame]{251 Nieuwland Science Hall\\
20 Department of Chemistry and Biochemistry\\ University of Notre
21 Dame\\ Notre Dame, Indiana 46556}
22
23 \keywords{}
24
25 \begin{document}
26
27 \begin{abstract}
28 We have investigated the structural and dynamic properties of the
29 basal and prismatic facets of an ice I$_\mathrm{h}$ / water
30 interface when the solid phase is being drawn through the liquid
31 (i.e. sheared relative to the fluid phase). To impose the shear, we
32 utilized a velocity-shearing and scaling (VSS) approach to reverse
33 non-equilibrium molecular dynamics (RNEMD). This method can create
34 simultaneous temperature and velocity gradients and allow the
35 measurement of transport properties at interfaces. The interfacial
36 width was found to be independent of relative velocity of the ice
37 and liquid layers over a wide range of shear rates. Decays of
38 molecular orientational time correlation functions for gave very
39 similar estimates for the width of the interfaces, although the
40 short- and longer-time decay components of the orientational
41 correlation functions behave differently closer to the interface.
42 Although both facets of ice are in ``stick'' boundary conditions in
43 liquid water, the solid-liquid friction coefficient was found to be
44 different for the basal and prismatic facets of ice.
45 \end{abstract}
46
47 \newpage
48
49 \section{Introduction}
50 %-----Outline of Intro---------------
51 % in general, ice/water interface is important b/c ....
52 % here are some people who have worked on ice/water, trying to understand the processes above ....
53 % with the recent development of VSS-RNEMD, we can now look at the shearing problem
54 % talk about what we will present in this paper
55 % -------End Intro------------------
56
57 %Gay02: cites many other ice/water papers, make sure to cite them.
58
59 Understanding the ice/water interface is essential for explaining
60 complex processes such as nucleation and crystal
61 growth,\cite{Han92,Granasy95,Vanfleet95} crystal
62 melting,\cite{Weber83,Han92,Sakai96,Sakai96B} and some fascinating
63 biological processes, such as the behavior of the antifreeze proteins
64 found in winter flounder,\cite{Wierzbicki07, Chapsky97} and certain
65 terrestrial arthropods.\cite{Duman:2001qy,Meister29012013} There has
66 been significant progress on understanding the structure and dynamics
67 of quiescent ice/water interfaces utilizing both theory and
68 experiment. Haymet \emph{et al.} have done extensive work on ice I$_\mathrm{h}$,
69 including characterizing and determining the width of the ice/water
70 interface for the SPC\cite{Karim90}, SPC/E\cite{Gay02,Bryk02}, CF1\cite{Hayward01,Hayward02}, and TIP4P\cite{Karim88} models for
71 water.
72 More recently, Haymet \emph{et al.} have investigated the effects
73 cations and anions have on crystal
74 nucleation\cite{Bryk04,Smith05,Wilson08,Wilson10}. Nada \emph{et al.}
75 have also studied ice/water
76 interfaces,\cite{Nada95,Nada00,Nada03,Nada12} and have found that the
77 differential growth rates of the facets of ice I$_\mathrm{h}$ are due to the the
78 reordering of the hydrogen bonding network\cite{Nada05}.
79
80 The movement of liquid water over the facets of ice has been less
81 thoroughly studied than the quiescent surfaces. This process is
82 potentially important in understanding transport of large blocks of
83 ice in water (which has important implications in the earth sciences),
84 as well as the relative motion of crystal-crystal interfaces that have
85 been separated by nanometer-scale fluid domains. In addition to
86 understanding both the structure and thickness of the interfacial
87 regions, it is important to understand the molecular origin of
88 friction, drag, and other changes in dynamical properties of the
89 liquid in the regions close to the surface that are altered by the
90 presence of a shearing of the bulk fluid relative to the solid phase.
91
92 In this work, we apply a recently-developed velocity shearing and
93 scaling approach to reverse non-equilibrium molecular dynamics
94 (VSS-RNEMD). This method makes it possible to calculate transport
95 properties like the interfacial thermal conductance across
96 heterogeneous interfaces,\cite{Kuang12} and can create simultaneous
97 temperature and velocity gradients and allow the measurement of
98 friction and thermal transport properties at interfaces. This has
99 allowed us to investigate the width of the ice/water interface as the
100 ice is sheared through the liquid, while simultaneously imposing a
101 weak thermal gradient to prevent frictional heating of the interface.
102 In the sections that follow, we discuss the methodology for creating
103 and simulating ice/water interfaces under shear and provide results
104 from both structural and dynamical correlation functions. We also
105 show that the solid-liquid interfacial friction coefficient depends
106 sensitively on the details of the surface morphology.
107
108 \section{Methodology}
109
110 \subsection{Stable ice I$_\mathrm{h}$ / water interfaces under shear}
111
112 The structure of ice I$_\mathrm{h}$ is well understood; it
113 crystallizes in a hexagonal space group P$6_3/mmc$, and the hexagonal
114 crystals of ice have two faces that are commonly exposed, the basal
115 face $\{0~0~0~1\}$, which forms the top and bottom of each hexagonal
116 plate, and the prismatic $\{1~0~\bar{1}~0\}$ face which forms the
117 sides of the plate. Other less-common, but still important, faces of
118 ice I$_\mathrm{h}$ are the secondary prism, $\{1~1~\bar{2}~0\}$, and
119 pyramidal, $\{2~0~\bar{2}~1\}$, faces. Ice I$_\mathrm{h}$ is normally
120 proton disordered in bulk crystals, although the surfaces probably
121 have a preference for proton ordering along strips of dangling H-atoms
122 and Oxygen lone pairs.\cite{Buch:2008fk}
123
124 \begin{wraptable}{r}{3.5in}
125 \caption{Mapping between the Miller indices of four facets of ice in
126 the $P6_3/mmc$ crystal system to the orthorhombic $P2_12_12_1$
127 system in reference \protect\cite{Hirsch04}}
128 \label{tab:equiv}
129 \begin{tabular}{|ccc|} \hline
130 & hexagonal & orthorhombic \\
131 & ($P6_3/mmc$) & ($P2_12_12_1$) \\
132 crystal face & Miller indices & equivalent \\ \hline
133 basal & $\{0~0~0~1\}$ & $\{0~0~1\}$ \\
134 prism & $\{1~0~\bar{1}~0\}$ & $\{1~0~0\}$ \\
135 secondary prism & $\{1~1~\bar{2}~0\}$ & $\{1~3~0\}$ \\
136 pyramidal & $\{2~0~\bar{2}~1\}$ & $\{2~0~1\}$ \\ \hline
137 \end{tabular}
138 \end{wraptable}
139
140 For small simulated ice interfaces, it is useful to work with
141 proton-ordered, but zero-dipole crystal that exposes these strips of
142 dangling H-atoms and lone pairs. When placing another material in
143 contact with one of the ice crystalline planes, it is also quite
144 useful to have an orthorhombic (rectangular) box. Recent work by
145 Hirsch and Ojam\"{a}e describes a number of alternative crystal
146 systems for proton-ordered bulk ice I$_\mathrm{h}$ using orthorhombic
147 cells.\cite{Hirsch04}
148
149 In this work, we are using Hirsch and Ojam\"{a}e's structure 6 which
150 is an orthorhombic cell ($P2_12_12_1$) that produces a proton-ordered
151 version of ice I$_\mathrm{h}$. Table \ref{tab:equiv} contains a
152 mapping between the Miller indices of common ice facets in the
153 P$6_3/mmc$ crystal system and those in the Hirsch and Ojam\"{a}e
154 $P2_12_12_1$ system.
155
156 Structure 6 from the Hirsch and Ojam\"{a}e paper has lattice
157 parameters $a = 4.49225$, $b = 7.78080$, $c = 7.33581$ and two water
158 molecules whose atoms reside at fractional coordinates given in table
159 \ref{tab:p212121}. To construct the basal and prismatic interfaces,
160 these crystallographic coordinates were used to construct an
161 orthorhombic unit cell which was then replicated in all three
162 dimensions yielding a proton-ordered block of ice I$_{h}$. To expose
163 the desired face, the orthorhombic representation was then cut along
164 the ($001$) or ($100$) planes for the basal and prismatic faces
165 respectively. The resulting block was rotated so that the exposed
166 faces were aligned with the $z$-axis normal to the exposed face. The
167 block was then cut along two perpendicular directions in a way that
168 allowed for perfect periodic replication in the $x$ and $y$ axes,
169 creating a slab with either the basal or prismatic faces exposed along
170 the $z$ axis. The slab was then replicated in the $x$ and $y$
171 dimensions until a desired sample size was obtained.
172
173 \begin{wraptable}{r}{2.85in}
174 \caption{Fractional coordinates for water in the orthorhombic
175 $P2_12_12_1$ system for ice I$_\mathrm{h}$ in reference
176 \protect\cite{Hirsch04}}
177 \label{tab:p212121}
178 \begin{tabular}{|cccc|} \hline
179 atom type & x & y & z \\ \hline
180 O & 0.75 & 0.1667 & 0.4375 \\
181 H & 0.5735 & 0.2202 & 0.4836 \\
182 H & 0.7420 & 0.0517 & 0.4836 \\
183 O & 0.25 & 0.6667 & 0.4375 \\
184 H & 0.2580 & 0.6693 & 0.3071 \\
185 H & 0.4265 & 0.7255 & 0.4756 \\ \hline
186 \end{tabular}
187 \end{wraptable}
188
189 Our ice / water interfaces were created using a box of liquid water
190 that had the same dimensions (in $x$ and $y$) as the ice block.
191 Although the experimental solid/liquid coexistence temperature under
192 atmospheric pressure is close to 273K, Haymet \emph{et al.} have done
193 extensive work on characterizing the ice/water interface, and find
194 that the coexistence temperature for simulated water is often quite a
195 bit different.\cite{Karim88,Karim90,Hayward01,Bryk02,Hayward02} They
196 have found that for the SPC/E water model,\cite{Berendsen87} which is
197 also used in this study, the ice/water interface is most stable at
198 225$\pm$5K.\cite{Bryk02} This liquid box was therefore equilibrated at
199 225 K and 1 atm of pressure in the NPAT ensemble (with the $z$ axis
200 allowed to fluctuate to equilibrate to the correct pressure). The
201 liquid and solid systems were combined by carving out any water
202 molecule from the liquid simulation cell that was within 3 \AA\ of any
203 atom in the ice slab.
204
205 Molecular translation and orientational restraints were applied in the
206 early stages of equilibration to prevent melting of the ice slab.
207 These restraints were removed during NVT equilibration, well before
208 data collection was carried out.
209
210 \subsection{Shearing ice / water interfaces without bulk melting}
211
212 As a solid is dragged through a liquid, there is frictional heating
213 that will act to melt the interface. To study the behavior of the
214 interface under a shear stress without causing the interface to melt,
215 it is necessary to apply a weak thermal gradient in combination with
216 the momentum gradient. This can be accomplished using the velocity
217 shearing and scaling (VSS) variant of reverse non-equilibrium
218 molecular dynamics (RNEMD), which utilizes a series of simultaneous
219 velocity exchanges between two regions within the simulation
220 cell.\cite{Kuang12} One of these regions is centered within the ice
221 slab, while the other is centrally located in the liquid
222 region. VSS-RNEMD provides a set of conservation constraints for
223 creating either a momentum flux or a thermal flux (or both
224 simultaneously) between the two slabs. Satisfying the constraint
225 equations ensures that the new configurations are sampled from the
226 same NVE ensemble as before the VSS move.
227
228 The VSS moves are applied periodically to scale and shift the particle
229 velocities ($\mathbf{v}_i$ and $\mathbf{v}_j$) in two slabs ($H$ and
230 $C$) which are separated by half of the simulation box,
231 \begin{displaymath}
232 \begin{array}{rclcl}
233
234 & \underline{\mathrm{shearing}} & &
235 \underline{~~~~~~~~~~~~\mathrm{scaling}~~~~~~~~~~~~} \\
236 \mathbf{v}_i \leftarrow &
237 \mathbf{a}_c\ & + & c\cdot\left(\mathbf{v}_i - \langle\mathbf{v}_c
238 \rangle\right) + \langle\mathbf{v}_c\rangle \\
239 \mathbf{v}_j \leftarrow &
240 \mathbf{a}_h & + & h\cdot\left(\mathbf{v}_j - \langle\mathbf{v}_h
241 \rangle\right) + \langle\mathbf{v}_h\rangle .
242
243 \end{array}
244 \end{displaymath}
245 Here $\langle\mathbf{v}_c\rangle$ and $\langle\mathbf{v}_h\rangle$ are
246 the center of mass velocities in the $C$ and $H$ slabs, respectively.
247 Within the two slabs, particles receive incremental changes or a
248 ``shear'' to their velocities. The amount of shear is governed by the
249 imposed momentum flux, $\mathbf{j}_z(\mathbf{p})$
250 \begin{eqnarray}
251 \mathbf{a}_c & = & - \mathbf{j}_z(\mathbf{p}) \Delta t / M_c \label{vss1}\\
252 \mathbf{a}_h & = & + \mathbf{j}_z(\mathbf{p}) \Delta t / M_h \label{vss2}
253 \end{eqnarray}
254 where $M_{\{c,h\}}$ is the total mass of particles within each of the
255 slabs and $\Delta t$ is the interval between two separate operations.
256
257 To simultaneously impose a thermal flux ($J_z$) between the slabs we
258 use energy conservation constraints,
259 \begin{eqnarray}
260 K_c - J_z\Delta t & = & c^2 (K_c - \frac{1}{2}M_c \langle\mathbf{v}_c
261 \rangle^2) + \frac{1}{2}M_c (\langle \mathbf{v}_c \rangle + \mathbf{a}_c)^2 \label{vss3}\\
262 K_h + J_z\Delta t & = & h^2 (K_h - \frac{1}{2}M_h \langle\mathbf{v}_h
263 \rangle^2) + \frac{1}{2}M_h (\langle \mathbf{v}_h \rangle +
264 \mathbf{a}_h)^2 \label{vss4}.
265 \label{constraint}
266 \end{eqnarray}
267 Simultaneous solution of these quadratic formulae for the scaling
268 coefficients, $c$ and $h$, will ensure that the simulation samples from
269 the original microcanonical (NVE) ensemble. Here $K_{\{c,h\}}$ is the
270 instantaneous translational kinetic energy of each slab. At each time
271 interval, it is a simple matter to solve for $c$, $h$, $\mathbf{a}_c$,
272 and $\mathbf{a}_h$, subject to the imposed momentum flux,
273 $j_z(\mathbf{p})$, and thermal flux, $J_z$, values. Since the VSS
274 operations do not change the kinetic energy due to orientational
275 degrees of freedom or the potential energy of a system, configurations
276 after the VSS move have exactly the same energy (and linear
277 momentum) as before the move.
278
279 As the simulation progresses, the VSS moves are performed on a regular
280 basis, and the system develops a thermal and/or velocity gradient in
281 response to the applied flux. In a bulk material, it is quite simple
282 to use the slope of the temperature or velocity gradients to obtain
283 either the thermal conductivity or shear viscosity.
284
285 The VSS-RNEMD approach is versatile in that it may be used to
286 implement thermal and shear transport simultaneously. Perturbations
287 of velocities away from the ideal Maxwell-Boltzmann distributions are
288 minimal, as is thermal anisotropy. This ability to generate
289 simultaneous thermal and shear fluxes has been previously utilized to
290 map out the shear viscosity of SPC/E water over a wide range of
291 temperatures (90~K) with a single 1 ns simulation.\cite{Kuang12}
292
293 For this work, we are using the VSS-RNEMD method primarily to generate
294 a shear between the ice slab and the liquid phase, while using a weak
295 thermal gradient to maintain the interface at the 225K target
296 value. This ensures minimal melting of the bulk ice phase and allows
297 us to control the exact temperature of the interface.
298
299 \subsection{Computational Details}
300 All simulations were performed using OpenMD,\cite{OOPSE,openmd} with a
301 time step of 2 fs and periodic boundary conditions in all three
302 dimensions. Electrostatics were handled using the damped-shifted
303 force real-space electrostatic kernel.\cite{Ewald} The systems were
304 divided into 100 bins along the $z$-axis for the VSS-RNEMD moves,
305 which were attempted every 50 fs.
306
307 The interfaces were equilibrated for a total of 10 ns at equilibrium
308 conditions before being exposed to either a shear or thermal gradient.
309 This consisted of 5 ns under a constant temperature (NVT) integrator
310 set to 225K followed by 5 ns under a microcanonical integrator. Weak
311 thermal gradients were allowed to develop using the VSS-RNEMD (NVE)
312 integrator using a a small thermal flux ($-2.0\times 10^{-6}$
313 kcal/mol/\AA$^2$/fs) for a duration of 5 ns to allow the gradient to
314 stabilize. The resulting temperature gradient was $\approx$ 10K over
315 the entire 100 \AA\ box length, which was sufficient to keep the
316 temperature at the interface within $\pm 1$ K of the 225K target.
317
318 Velocity gradients were then imposed using the VSS-RNEMD (NVE)
319 integrator with a range of momentum fluxes. These gradients were
320 allowed to stabilize for 1 ns before data collection began. Once
321 established, four successive 0.5 ns runs were performed for each shear
322 rate. During these simulations, snapshots of the system were taken
323 every 1 ps, and statistics on the structure and dynamics in each bin
324 were accumulated throughout the simulations.
325
326 \section{Results and discussion}
327
328 \subsection{Interfacial width}
329 Any order parameter or time correlation function that changes as one
330 crosses an interface from a bulk liquid to a solid can be used to
331 measure the width of the interface. In previous work on the ice/water
332 interface, Haymet {\it et al.}\cite{} have utilized structural
333 features (including the density) as well as dynamic properties
334 (including the diffusion constant) to estimate the width of the
335 interfaces for a number of facets of the ice crystals. Because
336 VSS-RNEMD imposes a lateral flow, parameters that depend on
337 translational motion of the molecules (e.g. diffusion) may be
338 artifically skewed by the RNEMD moves. A structural parameter is not
339 influenced by the RNEMD perturbations to the same degree. Here, we
340 have used the local tetraherdal order parameter as described by
341 Kumar\cite{Kumar09} and Errington\cite{Errington01} as our principal
342 measure of the interfacial width.
343
344 The local tetrahedral order parameter, $q(z)$, is given by
345 \begin{equation}
346 q(z) = \int_0^L \sum_{k=1}^{N} \Bigg(1 -\frac{3}{8}\sum_{i=1}^{3}
347 \sum_{j=i+1}^{4} \bigg(\cos\psi_{ikj}+\frac{1}{3}\bigg)^2\Bigg)
348 \delta(z_{k}-z)\mathrm{d}z \Bigg/ N_z
349 \label{eq:qz}
350 \end{equation}
351 where $\psi_{ikj}$ is the angle formed between the oxygen site on
352 central molecule $k$, and the oxygen sites on two of the four closest
353 molecules, $i$ and $j$. Molecules $i$ and $j$ are further restricted
354 to lie within the first solvation shell of molecule $k$. $N_z = \int
355 \delta(z_k - z) \mathrm{d}z$ is a normalization factor to account for
356 the varying population of molecules within each finite-width bin. The
357 local tetrahedral order parameter has a range of $(0,1)$, where the
358 larger values of $q$ indicate a larger degree of tetrahedral ordering
359 of the local environment. In perfect ice I$_\mathrm{h}$ structures,
360 the parameter can approach 1 at low temperatures, while in liquid
361 water, the ordering is significantly less tetrahedral, and values of
362 $q(z) \approx 0.75$ are more common.
363
364 To estimate the interfacial width, the system was divided into 100
365 bins along the $z$-dimension, and a cutoff radius for the first
366 solvation shell was set to 3.41 \AA\ . The $q_{z}$ function was
367 time-averaged to give yield a tetrahedrality profile of the
368 system. The profile was then fit to a hyperbolic tangent that smoothly
369 links the liquid and solid states,
370 \begin{equation}\label{tet_fit}
371 q(z) \approx
372 q_{liq}+\frac{q_{ice}-q_{liq}}{2}\left[\tanh\left(\frac{z-l}{w}\right)-\tanh\left(\frac{z-r}{w}\right)\right]+\beta\left|z-
373 \frac{r+l}{2}\right|.
374 \end{equation}
375 Here $q_{liq}$ and $q_{ice}$ are the local tetrahedral order parameter
376 for the bulk liquid and ice domains, respectively, $w$ is the width of
377 the interface. $l$ and $r$ are the midpoints of the left and right
378 interfaces, respectively. The last term in equation \eqref{tet_fit}
379 accounts for the influence that the weak thermal gradient has on the
380 tetrahedrality profile in the liquid region. To estimate the
381 10\%-90\% widths commonly used in previous studies,\cite{} it is a
382 simple matter to scale the widths obtained from the hyberbolic tangent
383 fits to obtain $w_{10-90} = 2.9 w$.\cite{}
384
385 In Figures \ref{fig:bComic} and \ref{fig:pComic} we see the
386 $z$-coordinate profiles for tetrahedrality, temperature, and the
387 $x$-component of the velocity for the basal and prismatic interfaces.
388 The lower panels show the $q(z)$ (black circles) along with the
389 hyperbolic tangent fits (red lines). In the liquid region, the local
390 tetrahedral order parameter, $q(z) \approx 0.75$ while in the
391 crystalline region, $q(z) \approx 0.94$, indicating a more tetrahedral
392 environment. The vertical dotted lines denote the midpoint of the
393 interfaces ($r$ and $l$ in equation \eqref{tet_fit}). The weak thermal
394 gradient applied to the systems in order to keep the interface at
395 225$\pm$5K, can be seen in middle panels. The tranverse velocity
396 profile is shown in the upper panels. It is clear from the upper
397 panels that water molecules in close proximity to the surface (i.e.
398 within 10 \AA\ to 15 \AA\ of the interfaces) have transverse
399 velocities quite close to the velocities within the ice block. There
400 is no velocity discontinuity at the interface, which indicates that
401 the shearing of ice/water interfaces occurs in the ``stick'' or
402 no-slip boundary conditions.
403
404 \begin{figure}
405 \includegraphics[width=\linewidth]{bComicStrip.pdf}
406 \caption{\label{fig:bComic} The basal interfaces. Lower panel: the
407 local tetrahedral order parameter, $q(z)$, (black circles) and the
408 hyperbolic tangent fit (red line). Middle panel: the imposed
409 thermal gradient required to maintain a fixed interfacial
410 temperature. Upper panel: the transverse velocity gradient that
411 develops in response to an imposed momentum flux. The vertical
412 dotted lines indicate the locations of the midpoints of the two
413 interfaces.}
414 \end{figure}
415
416 \begin{figure}
417 \includegraphics[width=\linewidth]{pComicStrip.pdf}
418 \caption{\label{fig:pComic} The prismatic interfaces. Panel
419 descriptions match those in figure \ref{fig:bComic}}
420 \end{figure}
421
422 From the fits using equation \eqref{tet_fit}, we find the interfacial
423 width for the basal and prismatic systems to be 3.2$\pm$0.4 \AA\ and
424 3.6$\pm$0.2 \AA\ , respectively, with no applied momentum flux. Over
425 the range of shear rates investigated, $0.6 \pm 0.3 \mathrm{ms}^{-1}
426 \rightarrow 5.3 \pm 0.5 \mathrm{ms}^{-1}$ for the basal system and
427 $0.9 \pm 0.2 \mathrm{ms}^{-1} \rightarrow 4.5 \pm 0.1
428 \mathrm{ms}^{-1}$ for the prismatic, we found no appreciable change in
429 the interface width. The fit values for the interfacial width ($w$)
430 over all shear rates contained the values reported above within their
431 error bars.
432
433 \subsubsection{Orientational Time Correlation Function}
434 The orientational time correlation function,
435 \begin{equation}\label{C(t)1}
436 C_{2}(t)=\langle P_{2}(\mathbf{u}(0) \cdot \mathbf{u}(t)) \rangle,
437 \end{equation}
438 gives insight into the local dynamic environment around the water
439 molecules. The rate at which the function decays provides information
440 about hindered motions and the timescales for relaxation. In
441 eq. \eqref{C(t)1}, $P_{2}$ is the second-order Legendre polynomial,
442 the vector $\mathbf{u}$ is often taken as HOH bisector, although
443 slightly different behavior can be observed when $\mathbf{u}$ is the
444 vector along one of the OH bonds. The angle brackets denote an
445 ensemble average over all water molecules in a given spatial region.
446
447 To investigate the dynamic behavior of water at the ice interfaces, we
448 have computed $C_{2}(z,t)$ for molecules that are present within a
449 particular slab along the $z$- axis at the initial time. The change
450 in the decay behavior as a function of the $z$ coordinate is another
451 measure of the change of how the local environment changes across the
452 ice/water interface. To compute these correlation functions, each of
453 the 0.5 ns simulations was followed by a shorter 200 ps microcanonical
454 (NVE) simulation in which the positions and orientations of every
455 molecule in the system were recorded every 0.1 ps. The systems were
456 then divided into 30 bins and $C_2(t)$ was evaluated for each bin.
457
458 In simulations of water at biological interfaces, Furse {\em et al.}
459 fit $C_2(t)$ functions for water with triexponential
460 functions,\cite{Furse08} where the three components of the decay
461 correspond to a fast (<200 fs) reorientational piece driven by the
462 restoring forces of existing hydrogen bonds, a middle (on the order of
463 several ps) piece describing the large angle jumps that occur during
464 the breaking and formation of new hydrogen bonds,and a slow (on the
465 order of tens of ps) contribution describing the translational motion
466 of the molecules. The model for orientational decay presented
467 recently by Laage and Hynes {\em et al.}\cite{Laage08,Laage11} also
468 includes three similar decay constants, although two of the time
469 constants are linked, and the resulting decay curve has two parameters
470 governing the dynamics of decay.
471
472 In our ice/water interfaces, we are at substantially lower
473 temperatures, and the water molecules are further perturbed by the
474 presence of the ice phase nearby. We have obtained the most
475 reasonable fits using triexponential functions with three distinct
476 time domains, as well as a constant piece that accounts for the water
477 stuck in the ice phase that does not experience any long-time
478 orientational decay,
479 \begin{equation}
480 C_{2}(t) \approx a e^{-t/\tau_\mathrm{short}} + b e^{-t/\tau_\mathrm{middle}} + c
481 e^{-t/\tau_\mathrm{long}} + (1-a-b-c)
482 \end{equation}
483 Average values for the three decay constants (and error estimates)
484 were obtained for each bin. In figures \ref{fig:basal_Tau_comic_strip}
485 and \ref{fig:prismatic_Tau_comic_strip}, the three orientational decay
486 times are shown as a function of distance from the center of the ice
487 slab.
488
489 \begin{figure}
490 \includegraphics[width=\linewidth]{basal_Tau_comic_strip.pdf}
491 \caption{\label{fig:basal_Tau_comic_strip} The three decay constants
492 of the orientational time correlation function, $C_2(t)$, for water
493 as a function of distance from the center of the ice slab. The
494 dashed line indicates the location of the basal face (as determined
495 from the tetrahedrality order parameter). The moderate and long
496 time contributions slow down close to the interface which would be
497 expected under reorganizations that inolve large motions of the
498 molecules (e.g. frame-reorientations and jumps). The observed
499 speed-up in the short time contribution is surprising, but appears
500 to reflect the restricted motion of librations closer to the
501 interface.}
502 \end{figure}
503
504 \begin{figure}
505 \includegraphics[width=\linewidth]{prismatic_Tau_comic_strip.pdf}
506 \caption{\label{fig:prismatic_Tau_comic_strip}
507 Decay constants for $C_2(t)$ at the prismatic interface. Panel
508 descriptions match those in figure \ref{fig:basal_Tau_comic_strip}.}
509 \end{figure}
510
511 Figures \ref{fig:basal_Tau_comic_strip} and
512 \ref{fig:prismatic_Tau_comic_strip} show the three decay constants for
513 the orientational time correlation function for water at varying
514 displacements from the center of the ice slab for both the basal and
515 prismatic interfaces. The vertical dotted lines indicate the
516 locations of the midpoints of the interfaces as determined by the
517 tetrahedrality fits. In the liquid regions, $\tau_{middle}$ and
518 $\tau_{long}$ have consistent values around 3-4 ps and 20-40 ps,
519 respectively, and increase in value approaching the interface.
520 According to the jump model of Laage and Hynes {\em et
521 al.},\cite{Laage08,Laage11} $\tau_{middle}$ corresponds to the
522 breaking and making of hydrogen bonds and $\tau_{long}$ is explained
523 with translational motion of the molecules (i.e. frame reorientation).
524 The shortest of the three decay constants, the librational time
525 $\tau_\mathrm{short}$ has a value of about 70 fs in the liquid region,
526 and decreases in value approaching the interface. The observed
527 speed-up in the short time contribution is surprising, but appears to
528 reflect the restricted motion of librations closer to the interface.
529
530 The control systems (with no applied momentum flux) are shown with
531 black symbols in figs. \ref{fig:basal_Tau_comic_strip} and
532 \ref{fig:prismatic_Tau_comic_strip}, while those obtained while a
533 shear was active are shown in red.
534
535 One remarkable feature, is that for each of the interfaces, there is
536 an apparent fixed liquid-state value for $\tau_{short}$,
537 $\tau_{middle}$, and $\tau_{long}$ at large displacements from the
538 interface. There also appears to be a single distance, $d_{basal}$ or
539 $d_{prismatic}$, from the interface at which all three decay times
540 begin to deviate from their bulk liquid values. We find $d_{basal}$
541 and $d_{prismatic}$ to be roughly 15 \AA\ and 8 \AA\ respectively.
542 These two results indicate that the dynamics of the water molecules
543 within $d_{basal}$ and $d_{prismatic}$ are being significantly
544 perturbed by the interface, even though the structural width of the
545 interface via analysis of the tetrahedrality profile indicates that
546 bulk liquid structure of water is recovered after about 4 \AA\ from
547 the edge of the ice.
548
549 Beaglehole and Wilson have measured the ice/water interface to have a
550 thickness approximately 10 \AA\ for both the basal and prismatic face
551 of ice by ellipticity measurements \cite{Beaglehole93}. Structurally,
552 we have found the basal and prismatic interfacial width to be
553 3.2$\pm$0.4 \AA\ and 3.6$\pm$0.2 \AA\ . However, decomposition of the
554 spatial dependence of the decay times of $C_2(t)$ appears to indicate
555 that a somewhat thicker interfacial region exists in which the
556 orientational dynamics of the water molecules begin to resemble the
557 trapped interfacial water more than the surrounding bulk.
558
559 \subsection{Coefficient of Friction of the Interface}
560 As the ice is sheared through the liquid, there will be a friction
561 between the solid and the liquid. Pit has shown how to calculate the
562 coefficient of friction $\lambda$ for a solid-liquid interface for a
563 Newtonian fluid of viscosity $\eta$ and has a slip length of
564 $\delta$. \cite{Pit99}
565 \begin{equation}\label{Pit}
566 \lambda=\eta/\delta
567 \end{equation}
568 From linear response theory, $\eta$ can be obtained from the imposed
569 momentum flux and the slope of the velocity about the dimension of the
570 imposed flux.\cite{Kuang12}
571 \begin{equation}\label{Kuang}
572 j_{z}(p_{x})=-\eta\frac{\partial v_{x}}{\partial z}
573 \end{equation}
574 Solving eq. \eqref{Kuang} for $\eta$ and substituting the result into
575 eq. \eqref{Pit}, we obtain an alternate expression for the coefficient
576 of friction.
577 \begin{equation}
578 \lambda=-\frac{j_{z}(p_{x})}{\delta \frac{\partial v_{x}}{\partial z}}
579 \end{equation}
580
581 For our simulations, we obtain $\delta$ from the difference between the structural edge of the ice block determined by the tetrahedrality profile fit, and the intersection of the linear regression of the $v_{x}$ profiles about the $z$-dimension for the ice and liquid. (See Figure \ref{fig:delta_example}) The coefficient of friction for the basal and the prismatic facets were found to be (0.07$\pm$0.01) amu fs\textsuperscript{-1} and (0.032$\pm$0.007) amu fs\textsuperscript{-1}. It is known that the basal and prismatic faces have different surface structures. The basal face is smoother than the prismatic with small alternating valleys and crests, while the prismatic surface has deep corrugating channels. We believe the reason that the prismatic face's coefficient of friction was found to be smaller than the basal's is due to the direction of the shear. The shear of the ice/water was in the same direction of the corrugating channels, allowing water molecules to pass through the channels during the shear.
582
583 \begin{figure}
584 \includegraphics[width=\linewidth]{delta_example.pdf}
585 \caption{\label{fig:delta_example} A schematic of determining the slip length ($\delta$). The slip length is the difference of the structural starting point of the ice and the point of intersection of the linear regressions of the liquid phase velocity profile (red) and of the solid ice velocity profile (black). The dotted line indicates the location of the ice as determined by the tetrahedrality profile.}
586 \end{figure}
587
588
589 \section{Conclusion}
590 Here we have simulated the basal and prismatic facets of an SPC/E model of the ice I$_\mathrm{h}$ / water interface. Using VSS-RNEMD, the ice was sheared relative to the liquid while imposed thermal gradients kept the interface at a stable temperature. Caculation of the local tetrahedrality order parameter has shown an apparent independence of the shear rate on the interfacial width, which was found to be 3.2$\pm$0.4 \AA\ and 3.6$\pm$0.2 \AA\ for the basal and prismatic systems. The orientational time correlation function was calculated from varying displacements from the interface. Decomposition by a triexponential decay also showed an apparent independence of the shear rate. The short time decay due to the restoring forces of existing hydrogen bonds decreased at close displacements from the interface, while the middle and long time decays were found to increase. There is also an apparent displacement, $d_{basal}$ and $d_{prismatic}$, from the interface at which these deviations from bulk liquid values occurs. We found $d_{basal}$ and $d_{prismatic}$ to be approximately 15 \AA\ and 8 \AA\ . This implies that the dynamics of water molecules which are structurally equivalent to bulk phase molecules are being perturbed by the presence of the ice and/or the interface. The coefficient of friction of each of the facets was also determined. They were found to be (0.07$\pm$0.01) amu fs\textsuperscript{-1} and (0.032$\pm$0.007) amu fs\textsuperscript{-1} for the basal and prismatic facets respectively. We believe the large difference between the two friction coefficients is due to the direction of the shear and the surface structure of the crystal facets.
591
592 \begin{acknowledgement}
593 Support for this project was provided by the National Science
594 Foundation under grant CHE-0848243. Computational time was provided
595 by the Center for Research Computing (CRC) at the University of
596 Notre Dame.
597 \end{acknowledgement}
598
599 \newpage
600 \bibstyle{achemso}
601 \bibliography{iceWater}
602
603 \begin{tocentry}
604 \begin{wrapfigure}{l}{0.5\textwidth}
605 \begin{center}
606 \includegraphics[width=\linewidth]{SystemImage.png}
607 \end{center}
608 \end{wrapfigure}
609 An image of our system.
610 \end{tocentry}
611
612 \end{document}
613
614 %**************************************************************
615 %Non-mass weighted slopes (amu*Angstroms^-2 * fs^-1)
616 % basal: slope=0.090677616, error in slope = 0.003691743
617 %prismatic: slope = 0.050101506, error in slope = 0.001348181
618 %Mass weighted slopes (Angstroms^-2 * fs^-1)
619 %basal slope = 4.76598E-06, error in slope = 1.94037E-07
620 %prismatic slope = 3.23131E-06, error in slope = 8.69514E-08
621 %**************************************************************