ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/iceWater/iceWater.tex
Revision: 3954
Committed: Mon Sep 9 18:48:22 2013 UTC (11 years ago) by gezelter
Content type: application/x-tex
File size: 35909 byte(s)
Log Message:
Nearing the end

File Contents

# User Rev Content
1 gezelter 3914 \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 gezelter 3897
10 gezelter 3914 \usepackage[version=3]{mhchem} % this is a great package for formatting chemical reactions
11     \usepackage{url}
12 gezelter 3897
13 gezelter 3949 \title{Simulations of solid-liquid friction at ice-I$_\mathrm{h}$ /
14     water interfaces}
15 gezelter 3897
16     \author{P. B. Louden}
17 gezelter 3914 \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 gezelter 3897
23 gezelter 3914 \keywords{}
24 gezelter 3897
25 gezelter 3914 \begin{document}
26 gezelter 3897
27 gezelter 3914 \begin{abstract}
28 gezelter 3945 We have investigated the structural and dynamic properties of the
29 gezelter 3954 basal and prismatic facets of the ice I$_\mathrm{h}$ / water
30     interface when the solid phase is drawn through the liquid
31 gezelter 3945 (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 gezelter 3954 width was found to be independent of the relative velocity of the
37     ice and liquid layers over a wide range of shear rates. Decays of
38     molecular orientational time correlation functions gave similar
39     estimates for the width of the interfaces, although the short- and
40     longer-time decay components behave differently closer to the
41     interface. Although both facets of ice are in ``stick'' boundary
42     conditions in liquid water, the solid-liquid friction coefficients
43     were found to be significantly different for the basal and prismatic
44     facets of ice.
45 gezelter 3914 \end{abstract}
46 gezelter 3897
47 gezelter 3914 \newpage
48 gezelter 3897
49     \section{Introduction}
50 plouden 3919 %-----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 gezelter 3897
57 plouden 3919 %Gay02: cites many other ice/water papers, make sure to cite them.
58    
59 gezelter 3945 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 plouden 3950 interface for the SPC\cite{Karim90}, SPC/E\cite{Gay02,Bryk02}, CF1\cite{Hayward01,Hayward02}, and TIP4P\cite{Karim88} models for
71     water.
72 gezelter 3945 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 plouden 3919
80 gezelter 3945 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 plouden 3919
92 gezelter 3945 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 gezelter 3897 \section{Methodology}
109    
110 gezelter 3945 \subsection{Stable ice I$_\mathrm{h}$ / water interfaces under shear}
111 gezelter 3914
112 gezelter 3946 The structure of ice I$_\mathrm{h}$ is well understood; it
113 gezelter 3914 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 gezelter 3945 ice I$_\mathrm{h}$ are the secondary prism, $\{1~1~\bar{2}~0\}$, and
119 gezelter 3946 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 gezelter 3914
124     \begin{wraptable}{r}{3.5in}
125 gezelter 3945 \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 gezelter 3914 \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 gezelter 3946 pyramidal & $\{2~0~\bar{2}~1\}$ & $\{2~0~1\}$ \\ \hline
137 gezelter 3914 \end{tabular}
138     \end{wraptable}
139    
140 gezelter 3946 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 gezelter 3914 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 gezelter 3945 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 gezelter 3946 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 gezelter 3945 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 gezelter 3914
173 gezelter 3946 \begin{wraptable}{r}{2.85in}
174 gezelter 3945 \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 gezelter 3914 \end{tabular}
187     \end{wraptable}
188    
189 gezelter 3946 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 gezelter 3914
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 gezelter 3945 data collection was carried out.
209 gezelter 3914
210 gezelter 3918 \subsection{Shearing ice / water interfaces without bulk melting}
211    
212 gezelter 3945 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 gezelter 3946 it is necessary to apply a weak thermal gradient in combination with
216     the momentum gradient. This can be accomplished using the velocity
217 gezelter 3945 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 gezelter 3946 slab, while the other is centrally located in the liquid
222 gezelter 3918 region. VSS-RNEMD provides a set of conservation constraints for
223 gezelter 3946 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 gezelter 3918
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 gezelter 3945 response to the applied flux. In a bulk material, it is quite simple
282 gezelter 3918 to use the slope of the temperature or velocity gradients to obtain
283 gezelter 3945 either the thermal conductivity or shear viscosity.
284 gezelter 3918
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 gezelter 3945 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 gezelter 3946 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 gezelter 3918
299 gezelter 3897 \subsection{Computational Details}
300 gezelter 3945 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 gezelter 3897
307 gezelter 3945 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 gezelter 3946 stabilize. The resulting temperature gradient was $\approx$ 10K over
315     the entire 100 \AA\ box length, which was sufficient to keep the
316 gezelter 3945 temperature at the interface within $\pm 1$ K of the 225K target.
317 gezelter 3897
318 gezelter 3945 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 plouden 3941 \section{Results and discussion}
327    
328 gezelter 3946 \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 gezelter 3954 interface, Haymet {\it et al.}\cite{Bryk02} have utilized structural
333 gezelter 3946 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 gezelter 3954 artificially skewed by the RNEMD moves. A structural parameter is not
339 gezelter 3946 influenced by the RNEMD perturbations to the same degree. Here, we
340 gezelter 3954 have used the local tetrahedral order parameter as described by
341 gezelter 3946 Kumar\cite{Kumar09} and Errington\cite{Errington01} as our principal
342 gezelter 3945 measure of the interfacial width.
343 plouden 3936
344 gezelter 3945 The local tetrahedral order parameter, $q(z)$, is given by
345 gezelter 3897 \begin{equation}
346 gezelter 3946 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 gezelter 3945 \label{eq:qz}
350 gezelter 3897 \end{equation}
351 gezelter 3945 where $\psi_{ikj}$ is the angle formed between the oxygen site on
352 gezelter 3946 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 plouden 3909
364 gezelter 3946 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 plouden 3941 \begin{equation}\label{tet_fit}
371 gezelter 3946 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 gezelter 3897 \end{equation}
375 gezelter 3946 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 gezelter 3954 10\%-90\% widths commonly used in previous studies,\cite{Bryk02} it is
382     a simple matter to scale the widths obtained from the hyperbolic
383     tangent fits to obtain $w_{10-90} = 2.1971 \times w$.\cite{Bryk02}
384 gezelter 3897
385 gezelter 3946 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 gezelter 3954 225$\pm$5K, can be seen in middle panels. The transverse velocity
396 gezelter 3946 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 gezelter 3897
404 gezelter 3914 \begin{figure}
405 gezelter 3946 \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 gezelter 3914 \end{figure}
415 plouden 3904
416 gezelter 3914 \begin{figure}
417 gezelter 3946 \includegraphics[width=\linewidth]{pComicStrip.pdf}
418     \caption{\label{fig:pComic} The prismatic interfaces. Panel
419     descriptions match those in figure \ref{fig:bComic}}
420 gezelter 3914 \end{figure}
421    
422 gezelter 3946 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 gezelter 3914
433 gezelter 3954 \subsubsection{Orientational Dynamics}
434 gezelter 3946 The orientational time correlation function,
435 plouden 3941 \begin{equation}\label{C(t)1}
436 gezelter 3946 C_{2}(t)=\langle P_{2}(\mathbf{u}(0) \cdot \mathbf{u}(t)) \rangle,
437 plouden 3941 \end{equation}
438 gezelter 3946 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 gezelter 3948 the vector $\mathbf{u}$ is often taken as HOH bisector, although
443     slightly different behavior can be observed when $\mathbf{u}$ is the
444 gezelter 3946 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 gezelter 3897
447 gezelter 3946 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 gezelter 3948 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 gezelter 3954 then divided into 30 bins along the $z$-axis and $C_2(t)$ was
457     evaluated for each bin.
458 plouden 3919
459 gezelter 3948 In simulations of water at biological interfaces, Furse {\em et al.}
460     fit $C_2(t)$ functions for water with triexponential
461     functions,\cite{Furse08} where the three components of the decay
462     correspond to a fast (<200 fs) reorientational piece driven by the
463     restoring forces of existing hydrogen bonds, a middle (on the order of
464     several ps) piece describing the large angle jumps that occur during
465     the breaking and formation of new hydrogen bonds,and a slow (on the
466     order of tens of ps) contribution describing the translational motion
467     of the molecules. The model for orientational decay presented
468     recently by Laage and Hynes {\em et al.}\cite{Laage08,Laage11} also
469     includes three similar decay constants, although two of the time
470     constants are linked, and the resulting decay curve has two parameters
471     governing the dynamics of decay.
472    
473     In our ice/water interfaces, we are at substantially lower
474     temperatures, and the water molecules are further perturbed by the
475     presence of the ice phase nearby. We have obtained the most
476     reasonable fits using triexponential functions with three distinct
477 gezelter 3954 time domains, as well as a constant piece to account for the water
478 gezelter 3948 stuck in the ice phase that does not experience any long-time
479     orientational decay,
480 gezelter 3897 \begin{equation}
481 gezelter 3946 C_{2}(t) \approx a e^{-t/\tau_\mathrm{short}} + b e^{-t/\tau_\mathrm{middle}} + c
482     e^{-t/\tau_\mathrm{long}} + (1-a-b-c)
483 gezelter 3897 \end{equation}
484 gezelter 3948 Average values for the three decay constants (and error estimates)
485     were obtained for each bin. In figures \ref{fig:basal_Tau_comic_strip}
486     and \ref{fig:prismatic_Tau_comic_strip}, the three orientational decay
487     times are shown as a function of distance from the center of the ice
488     slab.
489 gezelter 3897
490 plouden 3941 \begin{figure}
491 gezelter 3946 \includegraphics[width=\linewidth]{basal_Tau_comic_strip.pdf}
492 gezelter 3949 \caption{\label{fig:basal_Tau_comic_strip} The three decay constants
493     of the orientational time correlation function, $C_2(t)$, for water
494     as a function of distance from the center of the ice slab. The
495     dashed line indicates the location of the basal face (as determined
496     from the tetrahedrality order parameter). The moderate and long
497     time contributions slow down close to the interface which would be
498 gezelter 3954 expected under reorganizations that involve large motions of the
499 gezelter 3949 molecules (e.g. frame-reorientations and jumps). The observed
500     speed-up in the short time contribution is surprising, but appears
501     to reflect the restricted motion of librations closer to the
502     interface.}
503 plouden 3941 \end{figure}
504 gezelter 3897
505 gezelter 3914 \begin{figure}
506 gezelter 3946 \includegraphics[width=\linewidth]{prismatic_Tau_comic_strip.pdf}
507 gezelter 3949 \caption{\label{fig:prismatic_Tau_comic_strip}
508     Decay constants for $C_2(t)$ at the prismatic interface. Panel
509     descriptions match those in figure \ref{fig:basal_Tau_comic_strip}.}
510 gezelter 3914 \end{figure}
511 plouden 3904
512 gezelter 3949 Figures \ref{fig:basal_Tau_comic_strip} and
513     \ref{fig:prismatic_Tau_comic_strip} show the three decay constants for
514     the orientational time correlation function for water at varying
515     displacements from the center of the ice slab for both the basal and
516     prismatic interfaces. The vertical dotted lines indicate the
517     locations of the midpoints of the interfaces as determined by the
518     tetrahedrality fits. In the liquid regions, $\tau_{middle}$ and
519     $\tau_{long}$ have consistent values around 3-4 ps and 20-40 ps,
520     respectively, and increase in value approaching the interface.
521     According to the jump model of Laage and Hynes {\em et
522     al.},\cite{Laage08,Laage11} $\tau_{middle}$ corresponds to the
523     breaking and making of hydrogen bonds and $\tau_{long}$ is explained
524     with translational motion of the molecules (i.e. frame reorientation).
525     The shortest of the three decay constants, the librational time
526     $\tau_\mathrm{short}$ has a value of about 70 fs in the liquid region,
527     and decreases in value approaching the interface. The observed
528     speed-up in the short time contribution is surprising, but appears to
529     reflect the restricted motion of librations closer to the interface.
530 plouden 3937
531 gezelter 3949 The control systems (with no applied momentum flux) are shown with
532     black symbols in figs. \ref{fig:basal_Tau_comic_strip} and
533     \ref{fig:prismatic_Tau_comic_strip}, while those obtained while a
534     shear was active are shown in red.
535 plouden 3937
536 gezelter 3952 Two notable features deserve clarification. First, there are
537     nearly-constant liquid-state values for $\tau_{short}$,
538 gezelter 3949 $\tau_{middle}$, and $\tau_{long}$ at large displacements from the
539 gezelter 3952 interface. Second, there appears to be a single distance, $d_{basal}$
540     or $d_{prismatic}$, from the interface at which all three decay times
541     begin to deviate from their bulk liquid values. We find these
542     distances to be approximately 15 \AA\ and 8 \AA\ , respectively,
543     although significantly finer binning of the $C_2(t)$ data would be
544     necessary to provide better estimates of a ``dynamic'' interfacial
545     thickness.
546 plouden 3937
547 gezelter 3952 Beaglehole and Wilson have measured the ice/water interface using
548     ellipsometry and find a thickness of approximately 10 \AA\ for both
549     the basal and prismatic faces.\cite{Beaglehole93} Structurally, we
550     have found the basal and prismatic interfacial width to be 3.2$\pm$0.4
551     \AA\ and 3.6$\pm$0.2 \AA\ . However, decomposition of the spatial
552     dependence of the decay times of $C_2(t)$ indicates that a somewhat
553     thicker interfacial region exists in which the orientational dynamics
554     of the water molecules begin to resemble the trapped interfacial water
555     more than the surrounding liquid.
556 plouden 3937
557 gezelter 3952 Our results indicate that the dynamics of the water molecules within
558     $d_{basal}$ and $d_{prismatic}$ are being significantly perturbed by
559     the interface, even though the structural width of the interface via
560     analysis of the tetrahedrality profile indicates that bulk liquid
561     structure of water is recovered after about 4 \AA\ from the edge of
562     the ice.
563    
564 plouden 3941 \subsection{Coefficient of Friction of the Interface}
565 gezelter 3954 As liquid water flows over an ice interface, there is a distance from
566     the structural interface where bulk-like hydrodynamics are recovered.
567     Bocquet and Barrat constructed a theory for the hydrodynamic boundary
568     parameters, which include the slipping length
569 gezelter 3952 $\left(\delta_\mathrm{wall}\right)$ of this boundary layer and the
570     ``hydrodynamic position'' of the boundary
571 gezelter 3954 $\left(z_\mathrm{wall}\right)$.\cite{PhysRevLett.70.2726,PhysRevE.49.3079}
572     This last parameter is the location (relative to a solid surface)
573 gezelter 3952 where the bulk-like behavior is recovered. Work by Mundy {\it et al.}
574 gezelter 3954 has helped to combine these parameters into a liquid-solid friction
575 gezelter 3952 coefficient, which quantifies the resistance to pulling the solid
576 gezelter 3954 interface through a liquid,\cite{Mundy1997305}
577 gezelter 3952 \begin{equation}
578     \lambda_\mathrm{wall} = \frac{\eta}{\delta_\mathrm{wall}}.
579     \end{equation}
580 gezelter 3954 This expression is nearly identical to one provided by Pit {\it et
581     al.} for the solid-liquid friction of an interface,\cite{Pit99}
582 plouden 3942 \begin{equation}\label{Pit}
583 gezelter 3952 \lambda=\frac{\eta}{\delta}
584 plouden 3942 \end{equation}
585 gezelter 3952 where $\delta$ is the slip length for the liquid measured at the
586     location of the interface itself.
587    
588     In both of these expressions, $\eta$ is the shear viscosity of the
589     bulk-like region of the liquid, a quantity which is easily obtained in
590     VSS-RNEMD simulations by fitting the velocity profile in the region
591 gezelter 3954 far from the surface.\cite{Kuang12} Assuming linear response in the
592 gezelter 3952 bulk-like region,
593 plouden 3942 \begin{equation}\label{Kuang}
594 gezelter 3952 j_{z}(p_{x})=-\eta \left(\frac{\partial v_{x}}{\partial z}\right)
595 plouden 3942 \end{equation}
596 gezelter 3952 Substituting this result into eq. \eqref{Pit}, we can estimate the
597     solid-liquid coefficient using the slip length,
598 plouden 3941 \begin{equation}
599 gezelter 3954 \lambda=-\frac{j_{z}(p_{x})} {\left(\frac{\partial v_{x}}{\partial
600     z}\right) \delta}
601 plouden 3941 \end{equation}
602 plouden 3937
603 gezelter 3954 For ice / water interfaces, the boundary conditions are markedly
604     no-slip, so projecting the bulk liquid state velocity profile yields a
605     negative slip length. This length is the difference between the
606     structural edge of the ice (determined by the tetrahedrality profile)
607     and the location where the projected velocity of the bulk liquid
608     intersects the solid phase velocity (see Figure
609     \ref{fig:delta_example}). The coefficients of friction for the basal
610     and the prismatic facets are found to be (0.07$\pm$0.01) amu
611 gezelter 3952 fs\textsuperscript{-1} and (0.032$\pm$0.007) amu
612 gezelter 3954 fs\textsuperscript{-1}, respectively. These results may seem
613     surprising as the basal face is smoother than the prismatic with only
614     small undulations of the oxygen positions, while the prismatic surface
615     has deep corrugated channels. The applied momentum flux used in our
616     simulations is parallel to these channels, however, and this results
617     in a flow of water in the same direction as the corrugations, allowing
618     water molecules to pass through the channels during the shear.
619 plouden 3939
620 plouden 3942 \begin{figure}
621 gezelter 3946 \includegraphics[width=\linewidth]{delta_example.pdf}
622 gezelter 3952 \caption{\label{fig:delta_example} Determining the (negative) slip
623     length ($\delta$) for the ice-water interfaces (which have decidedly
624     non-slip behavior). This length is the difference between the
625     structural edge of the ice (determined by the tetrahedrality
626     profile) and the location where the projected velocity of the bulk
627     liquid (dashed red line) intersects the solid phase velocity (solid
628     black line). The dotted line indicates the location of the ice as
629     determined by the tetrahedrality profile.}
630 plouden 3942 \end{figure}
631    
632    
633 gezelter 3897 \section{Conclusion}
634 gezelter 3952 We have simulated the basal and prismatic facets of an SPC/E model of
635     the ice I$_\mathrm{h}$ / water interface. Using VSS-RNEMD, the ice
636     was sheared relative to the liquid while simultaneously being exposed
637     to a weak thermal gradient which kept the interface at a stable
638     temperature. Calculation of the local tetrahedrality order parameter
639     has shown an apparent independence of the interfacial width on the
640     shear rate. This width was found to be 3.2$\pm$0.4 \AA\ and
641     3.6$\pm$0.2 \AA\ for the basal and prismatic systems, respectively.
642 gezelter 3897
643 gezelter 3952 Orientational time correlation functions were calculated at varying
644     displacements from the interface, and were found to be similarly
645     independent of the applied momentum flux. The short decay due to the
646     restoring forces of existing hydrogen bonds decreased close to the
647     interface, while the longer-time decay constants increased in close
648     proximity to the interface. There is also an apparent dynamic
649     interface width, $d_{basal}$ and $d_{prismatic}$, at which these
650 gezelter 3954 deviations from bulk liquid values begin. We found $d_{basal}$ and
651     $d_{prismatic}$ to be approximately 15 \AA\ and 8 \AA\ . This implies
652     that the dynamics of water molecules which have similar structural
653     environments to liquid phase molecules are dynamically perturbed by
654     the presence of the ice interface.
655 gezelter 3952
656 gezelter 3954 The coefficient of liquid-solid friction for each of the facets was
657     also determined. They were found to be (0.07$\pm$0.01) amu
658 gezelter 3952 fs\textsuperscript{-1} and (0.032$\pm$0.007) amu
659     fs\textsuperscript{-1} for the basal and prismatic facets
660     respectively. We attribute the large difference between the two
661     friction coefficients to the direction of the shear and to the surface
662     structure of the crystal facets.
663    
664 gezelter 3914 \begin{acknowledgement}
665     Support for this project was provided by the National Science
666     Foundation under grant CHE-0848243. Computational time was provided
667     by the Center for Research Computing (CRC) at the University of
668     Notre Dame.
669     \end{acknowledgement}
670 gezelter 3897
671 gezelter 3914 \newpage
672     \bibstyle{achemso}
673 plouden 3909 \bibliography{iceWater}
674 gezelter 3897
675 gezelter 3914 \begin{tocentry}
676     \begin{wrapfigure}{l}{0.5\textwidth}
677     \begin{center}
678     \includegraphics[width=\linewidth]{SystemImage.png}
679     \end{center}
680     \end{wrapfigure}
681 gezelter 3954 The cell used to simulate liquid-solid shear in ice I$_\mathrm{h}$ /
682     water interfaces. Velocity gradients were applied using the velocity
683     shearing and scaling variant of reverse non-equilibrium molecular
684     dynamics (VSS-RNEMD) with a weak thermal gradient to prevent melting.
685     The interface width is relatively robust in both structual and dynamic
686     measures as a function of the applied shear.
687 gezelter 3914 \end{tocentry}
688 gezelter 3897
689 gezelter 3914 \end{document}
690 gezelter 3897
691 plouden 3924 %**************************************************************
692     %Non-mass weighted slopes (amu*Angstroms^-2 * fs^-1)
693     % basal: slope=0.090677616, error in slope = 0.003691743
694     %prismatic: slope = 0.050101506, error in slope = 0.001348181
695     %Mass weighted slopes (Angstroms^-2 * fs^-1)
696     %basal slope = 4.76598E-06, error in slope = 1.94037E-07
697     %prismatic slope = 3.23131E-06, error in slope = 8.69514E-08
698     %**************************************************************