ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/iceWater/iceWater.tex
Revision: 3945
Committed: Tue Sep 3 21:16:58 2013 UTC (11 years ago) by gezelter
Content type: application/x-tex
File size: 34602 byte(s)
Log Message:
Some editing

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    
14 gezelter 3945 \title{Solid-liquid friction at ice-I$_\mathrm{h}$ / 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     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 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     interface for the SPC, SPC/E, CF1, and TIP4P models for
71     water.\cite{Karim87,Karim88,Karim90,Hayward01,Bryk02,Hayward02,Gay02}
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 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 3945 The structure of ice I$_\mathrm{h}$ is very 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     the prismatic, $\{2~0~\bar{2}~1\}$, faces. Ice I$_\mathrm{h}$ is
120     normally proton disordered in bulk crystals, although the surfaces
121     probably have a preference for proton ordering along strips of
122     dangling H-atoms and Oxygen lone pairs.\cite{Buch:2008fk}
123 gezelter 3914
124     For small simulated ice interfaces, it is useful to have a
125     proton-ordered, but zero-dipole crystal that exposes these strips of
126 gezelter 3945 dangling H-atoms and lone pairs. When placing another material in
127     contact with one of the ice crystalline planes, it is useful to have
128     an orthorhombic (rectangular) box. A recent paper by Hirsch and
129     Ojam\"{a}e describes how to create proton-ordered bulk ice
130     I$_\mathrm{h}$ in alternative orthorhombic cells.\cite{Hirsch04}
131 gezelter 3914
132 plouden 3941 We are using Hirsch and Ojam\"{a}e's structure 6 which is an
133 gezelter 3914 orthorhombic cell ($P2_12_12_1$) that produces a proton-ordered
134 gezelter 3945 version of ice I$_\mathrm{h}$. Table \ref{tab:equiv} contains a mapping between
135 gezelter 3914 the Miller indices in the P$6_3/mmc$ crystal system and those in the
136     Hirsch and Ojam\"{a}e $P2_12_12_1$ system.
137    
138     \begin{wraptable}{r}{3.5in}
139 gezelter 3945 \caption{Mapping between the Miller indices of four facets of ice in
140     the $P6_3/mmc$ crystal system to the orthorhombic $P2_12_12_1$
141     system in reference \protect\cite{Hirsch04}}
142     \label{tab:equiv}
143 gezelter 3914 \begin{tabular}{|ccc|} \hline
144     & hexagonal & orthorhombic \\
145     & ($P6_3/mmc$) & ($P2_12_12_1$) \\
146     crystal face & Miller indices & equivalent \\ \hline
147     basal & $\{0~0~0~1\}$ & $\{0~0~1\}$ \\
148     prism & $\{1~0~\bar{1}~0\}$ & $\{1~0~0\}$ \\
149     secondary prism & $\{1~1~\bar{2}~0\}$ & $\{1~3~0\}$ \\
150     pryamidal & $\{2~0~\bar{2}~1\}$ & $\{2~0~1\}$ \\ \hline
151     \end{tabular}
152     \end{wraptable}
153    
154     Structure 6 from the Hirsch and Ojam\"{a}e paper has lattice
155     parameters $a = 4.49225$, $b = 7.78080$, $c = 7.33581$ and two water
156 gezelter 3945 molecules whose atoms reside at fractional coordinates given in table
157     \ref{tab:p212121}. To construct the basal and prismatic interfaces,
158     these crystallographic coordinates were used to construct an
159     orthorhombic unit cell which was then replicated in all three
160     dimensions yielding a proton-ordered block of ice I$_{h}$. To expose
161     the desired face, the orthorhombic representation was then cut along
162     the ($001$) or ($100$) planes for the basal and prismatic faces
163     respectively. The resulting block was rotated so that the exposed
164     faces were aligned with the $z$-dimension normal to the exposed face.
165     The block was then cut along two perpendicular directions in a way
166     that allowed for perfect periodic replication in the $x$ and $y$ axes,
167     creating a slab with either the basal or prismatic faces exposed along
168     the $z$ axis. The slab was then replicated in the $x$ and $y$
169     dimensions until a desired sample size was obtained.
170 gezelter 3914
171     \begin{wraptable}{r}{3.25in}
172 gezelter 3945 \caption{Fractional coordinates for water in the orthorhombic
173     $P2_12_12_1$ system for ice I$_\mathrm{h}$ in reference
174     \protect\cite{Hirsch04}}
175     \label{tab:p212121}
176     \begin{tabular}{|cccc|} \hline
177     atom type & x & y & z \\ \hline
178     O & 0.75 & 0.1667 & 0.4375 \\
179     H & 0.5735 & 0.2202 & 0.4836 \\
180     H & 0.7420 & 0.0517 & 0.4836 \\
181     O & 0.25 & 0.6667 & 0.4375 \\
182     H & 0.2580 & 0.6693 & 0.3071 \\
183     H & 0.4265 & 0.7255 & 0.4756 \\ \hline
184 gezelter 3914 \end{tabular}
185     \end{wraptable}
186    
187     Although experimental solid/liquid coexistant temperature under normal
188     pressure are close to 273K, Haymet \emph{et al.} have done extensive
189     work on characterizing the ice/water
190     interface.\cite{Karim88,Karim90,Hayward01,Bryk02,Hayward02} They have
191 gezelter 3945 found that for the SPC/E water model,\cite{Berendsen87} which is also
192     used in this study, the ice/water interface is most stable at
193     225$\pm$5K.\cite{Bryk02} Therefore, we created our ice / water
194     interfaces, utilizing a box of liquid water that had the same
195     dimensions in $x$ and $y$ was equilibrated at 225 K and 1 atm of
196     pressure in the NPAT ensemble (with the $z$ axis allowed to fluctuate
197     to equilibrate to the correct pressure). The liquid and solid systems
198     were combined by carving out any water molecule from the liquid
199     simulation cell that was within 3 \AA\ of any atom in the ice slab.
200 gezelter 3914
201     Molecular translation and orientational restraints were applied in the
202     early stages of equilibration to prevent melting of the ice slab.
203     These restraints were removed during NVT equilibration, well before
204 gezelter 3945 data collection was carried out.
205 gezelter 3914
206 gezelter 3918 \subsection{Shearing ice / water interfaces without bulk melting}
207    
208 gezelter 3945 As a solid is dragged through a liquid, there is frictional heating
209     that will act to melt the interface. To study the behavior of the
210     interface under a shear stress without causing the interface to melt,
211     it is necessary to apply a weak thermal gradient along with the
212     momentum gradient. This can be accomplished using he velocity
213     shearing and scaling (VSS) variant of reverse non-equilibrium
214     molecular dynamics (RNEMD), which utilizes a series of simultaneous
215     velocity exchanges between two regions within the simulation
216     cell.\cite{Kuang12} One of these regions is centered within the ice
217     slab, while the other is centrally located in the liquid phase
218 gezelter 3918 region. VSS-RNEMD provides a set of conservation constraints for
219     simultaneously creating either a momentum flux or a thermal flux (or
220     both) between the two slabs. Satisfying the constraint equations
221     ensures that the new configurations are sampled from the same NVE
222 gezelter 3945 ensemble as before the VSS move.
223 gezelter 3918
224     The VSS moves are applied periodically to scale and shift the particle
225     velocities ($\mathbf{v}_i$ and $\mathbf{v}_j$) in two slabs ($H$ and
226     $C$) which are separated by half of the simulation box,
227     \begin{displaymath}
228     \begin{array}{rclcl}
229    
230     & \underline{\mathrm{shearing}} & &
231     \underline{~~~~~~~~~~~~\mathrm{scaling}~~~~~~~~~~~~} \\
232     \mathbf{v}_i \leftarrow &
233     \mathbf{a}_c\ & + & c\cdot\left(\mathbf{v}_i - \langle\mathbf{v}_c
234     \rangle\right) + \langle\mathbf{v}_c\rangle \\
235     \mathbf{v}_j \leftarrow &
236     \mathbf{a}_h & + & h\cdot\left(\mathbf{v}_j - \langle\mathbf{v}_h
237     \rangle\right) + \langle\mathbf{v}_h\rangle .
238    
239     \end{array}
240     \end{displaymath}
241     Here $\langle\mathbf{v}_c\rangle$ and $\langle\mathbf{v}_h\rangle$ are
242     the center of mass velocities in the $C$ and $H$ slabs, respectively.
243     Within the two slabs, particles receive incremental changes or a
244     ``shear'' to their velocities. The amount of shear is governed by the
245     imposed momentum flux, $\mathbf{j}_z(\mathbf{p})$
246     \begin{eqnarray}
247     \mathbf{a}_c & = & - \mathbf{j}_z(\mathbf{p}) \Delta t / M_c \label{vss1}\\
248     \mathbf{a}_h & = & + \mathbf{j}_z(\mathbf{p}) \Delta t / M_h \label{vss2}
249     \end{eqnarray}
250     where $M_{\{c,h\}}$ is the total mass of particles within each of the
251     slabs and $\Delta t$ is the interval between two separate operations.
252    
253     To simultaneously impose a thermal flux ($J_z$) between the slabs we
254     use energy conservation constraints,
255     \begin{eqnarray}
256     K_c - J_z\Delta t & = & c^2 (K_c - \frac{1}{2}M_c \langle\mathbf{v}_c
257     \rangle^2) + \frac{1}{2}M_c (\langle \mathbf{v}_c \rangle + \mathbf{a}_c)^2 \label{vss3}\\
258     K_h + J_z\Delta t & = & h^2 (K_h - \frac{1}{2}M_h \langle\mathbf{v}_h
259     \rangle^2) + \frac{1}{2}M_h (\langle \mathbf{v}_h \rangle +
260     \mathbf{a}_h)^2 \label{vss4}.
261     \label{constraint}
262     \end{eqnarray}
263     Simultaneous solution of these quadratic formulae for the scaling
264     coefficients, $c$ and $h$, will ensure that the simulation samples from
265     the original microcanonical (NVE) ensemble. Here $K_{\{c,h\}}$ is the
266     instantaneous translational kinetic energy of each slab. At each time
267     interval, it is a simple matter to solve for $c$, $h$, $\mathbf{a}_c$,
268     and $\mathbf{a}_h$, subject to the imposed momentum flux,
269     $j_z(\mathbf{p})$, and thermal flux, $J_z$, values. Since the VSS
270     operations do not change the kinetic energy due to orientational
271     degrees of freedom or the potential energy of a system, configurations
272     after the VSS move have exactly the same energy (and linear
273     momentum) as before the move.
274    
275     As the simulation progresses, the VSS moves are performed on a regular
276     basis, and the system develops a thermal and/or velocity gradient in
277 gezelter 3945 response to the applied flux. In a bulk material, it is quite simple
278 gezelter 3918 to use the slope of the temperature or velocity gradients to obtain
279 gezelter 3945 either the thermal conductivity or shear viscosity.
280 gezelter 3918
281     The VSS-RNEMD approach is versatile in that it may be used to
282     implement thermal and shear transport simultaneously. Perturbations
283     of velocities away from the ideal Maxwell-Boltzmann distributions are
284     minimal, as is thermal anisotropy. This ability to generate
285     simultaneous thermal and shear fluxes has been previously utilized to
286     map out the shear viscosity of SPC/E water over a wide range of
287     temperatures (90~K) with a single 1 ns simulation.\cite{Kuang12}
288    
289 gezelter 3945 For this work, we are using the VSS-RNEMD method primarily to generate
290     a shear between the ice slab and the liquid phase, while using a weak
291     thermal gradient to maintaining the interface at the 225K target
292     value. This will insure minimal melting of the bulk ice phase and
293     allows us to control the exact temperature of the interface.
294 gezelter 3918
295 gezelter 3897 \subsection{Computational Details}
296 gezelter 3945 All simulations were performed using OpenMD,\cite{OOPSE,openmd} with a
297     time step of 2 fs and periodic boundary conditions in all three
298     dimensions. Electrostatics were handled using the damped-shifted
299     force real-space electrostatic kernel.\cite{Ewald} The systems were
300     divided into 100 bins along the $z$-axis for the VSS-RNEMD moves,
301     which were attempted every 50 fs.
302 gezelter 3897
303 gezelter 3945 The interfaces were equilibrated for a total of 10 ns at equilibrium
304     conditions before being exposed to either a shear or thermal gradient.
305     This consisted of 5 ns under a constant temperature (NVT) integrator
306     set to 225K followed by 5 ns under a microcanonical integrator. Weak
307     thermal gradients were allowed to develop using the VSS-RNEMD (NVE)
308     integrator using a a small thermal flux ($-2.0\times 10^{-6}$
309     kcal/mol/\AA$^2$/fs) for a duration of 5 ns to allow the gradient to
310     stabilize. The resulting temperature gradient was less than 5K over
311     the entire 1 nm box length, which was sufficient to keep the
312     temperature at the interface within $\pm 1$ K of the 225K target.
313 gezelter 3897
314 gezelter 3945 Velocity gradients were then imposed using the VSS-RNEMD (NVE)
315     integrator with a range of momentum fluxes. These gradients were
316     allowed to stabilize for 1 ns before data collection began. Once
317     established, four successive 0.5 ns runs were performed for each shear
318     rate. During these simulations, snapshots of the system were taken
319     every 1 ps, and statistics on the structure and dynamics in each bin
320     were accumulated throughout the simulations.
321    
322 plouden 3941 \section{Results and discussion}
323    
324 gezelter 3897 \subsection{Measuring the Width of the Interface}
325 gezelter 3945 Any order parameter or correlation function that varies across the
326     interface from a bulk liquid to a solid can be used as a measure of
327     the width of the interface. However, because VSS-RNEMD imposes a
328     lateral flow, parameters that depend on translational motion of the
329     molecules (e.g. the diffusion constant) may be artifically skewed by
330     the RNEMD moves. A structural parameter like a radial distribution
331     function is not influenced by the RNEMD perturbations to the same
332     degree. Here, we have used the local tetraherdal order parameter as
333     described by Kumar\cite{Kumar09} and Errington\cite{Errington01} as a
334     measure of the interfacial width.
335 plouden 3936
336 gezelter 3945 The local tetrahedral order parameter, $q(z)$, is given by
337 gezelter 3897 \begin{equation}
338 gezelter 3945 q(z) \equiv \int_0^L \sum_{k=1}^{N} \Bigg(1 -\frac{3}{8}\sum_{i=1}^{3}
339     \sum_{j=i+1}^{4} \bigg[\cos\psi_{ikj}+\frac{1}{3}\bigg]^2\Bigg)
340     \delta(z_{k}-z)\mathrm{d}z
341     \label{eq:qz}
342 gezelter 3897 \end{equation}
343 gezelter 3945 where $\psi_{ikj}$ is the angle formed between the oxygen site on
344     molecule $k$, and the oxygen sites on its two closest neighbors,
345     molecules $i$ and $j$. The local tetrahedral order parameter function
346     has a range of (0,1), where the larger the value $q$ has the more
347     tetrahedral the ordering of the local environment is. A $q$ value of
348     one describes a perfectly tetrahedral environment relative to it and
349     its four nearest neighbors, and the parameter's value decreases as the
350     local ordering becomes less tetrahedral. Equation \ref{eq:qz}
351     describes a $z$-binned tetrahedral order parameter in which the $z$
352     coordinate of the central molecule is used to give a spatial
353     description of the local orientational ordering.
354 plouden 3909
355 gezelter 3945 The system was divided into 100 bins along the $z$-dimension, and a
356     cutoff radius for the neighboring molecules was set to 3.41 \AA\ .
357     The $q_{z}$ values for each snapshot were then averaged to give a
358     tetrahedrality profile of the system about the $z$-dimension. The
359     profile was then fit with a hyperbolic tangent function given by
360 plouden 3909
361 plouden 3941 \begin{equation}\label{tet_fit}
362 plouden 3942 q_{z} \approx q_{liq}+\frac{q_{ice}-q_{liq}}{2}\Bigg[\tanh\bigg(\frac{z-I_{L,m}}{w}\bigg)-\tanh\bigg(\frac{z-I_{R,m}}{w}\bigg)\Bigg]+\beta|(z-z_{mid})|
363 gezelter 3897 \end{equation}
364    
365 plouden 3942 where $q_{liq}$ and $q_{ice}$ are fitting parameters for the local tetrahedral order parameter for the liquid and ice, $w$ is the width of the interface, and $I_{L,m}$ and $I_{R,m}$ are the midpoints of the left and right interface. The last term in \eqref{tet_fit} accounts for the influence the thermal gradient has on the tetrahedrality profile in the liquid region; where $\beta$ is a fitting parameter and $z_{mid}$ is the midpoint of the $z$-dimension of the simulation box.
366 gezelter 3897
367 plouden 3942 In Figures \ref{fig:bComic} and \ref{fig:pComic} we see the $z$-dimensional profiles for several components of the basal and prismatic systems. In panel (a) of the figures we see the tetrahedrality profile of the systems (black circles). In the liquid region of the system, the local tetrahedral order parameter is approximately 0.75 while in the solid region the parameter is approximately 0.94, indicating a more tetrahedral structure of the water molecules. The hyperbolic tangent function used to fit the tetrahedrality profiles is in red and the verticle dotted lines denote the midpoint of the interfaces. The weak thermal gradient applied to the systems in order to keep the interface at a stable temperature, 225$\pm$5K, can be seen in panel (b). Lastly, the velocity gradient across the systems can be seen in panel (c). From panel (c), we can see liquid phase water molecules 10 \AA\ to 15 \AA\ from the interfaces are being dragged along with the ice block, indicating that the shearing of ice water is in the stick boundary condition.
368 plouden 3898
369 gezelter 3914 \begin{figure}
370     \includegraphics[width=\linewidth]{bComicStrip}
371     \caption{\label{fig:bComic} The basal system: (a) The local tetrahedral order parameter,$q$, (black circles) fit by a hyperbolic tangent (red line), (b) the thermal gradient imposed on the system to maintain a stable interfacial temperature, and (c) the velocity gradient imposed on the system. The verticle dotted lines indicate the midpoint of the interfaces.}
372     \end{figure}
373 plouden 3904
374 gezelter 3914 \begin{figure}
375     \includegraphics[width=\linewidth]{pComicStrip}
376     \caption{\label{fig:pComic} The prismatic system: (a) The local tetrahedral order parameter,$q$, (black circles) fit by a hyperbolic tangent (red line), (b) the thermal gradient imposed on the system to maintain a stable interfacial temperature, and (c) the velocity gradient imposed on the system.The verticle dotted lines indicate the midpoint of the interfaces.}
377     \end{figure}
378    
379 plouden 3942 From the tetrahedrality fits, we found the interfacial width for the basal and prismatic systems to be 3.2$\pm$0.4 \AA\ and 3.6$\pm$0.2 \AA\ with no applied momentum flux. Over the range of shear rates investigated, 0.6$\pm$0.3 ms\textsuperscript{-1} to 5.3$\pm$0.5 ms\textsuperscript{-1} for the basal system and 0.9$\pm$0.2 ms\textsuperscript{-1} to 4.5$\pm$0.1 ms\textsuperscript{-1} for the prismatic, there was no appreciable change in the interface width found. The calculated values for the interfacial width over all shear rates investigated contained the control values within their error bars.
380 gezelter 3914
381 plouden 3941 \subsubsection{Orientational Time Correlation Function}
382     The orientational time correlation function (OTCF) gives insight of the local environment of molecules. The rate at which the function decays corresponds to how hindered the motions of a molecule are. The more hindered a molecules motion is the slower the function will decay, and the function decays more rapidly for molecules with less constrained motions.
383     \begin{equation}\label{C(t)1}
384     C_{2}(t)=\langle P_{2}(\mathbf{v}_{i}(t)\mathbf{v}_{i}(t=0))\rangle
385     \end{equation}
386 plouden 3942 In eq. \eqref{C(t)1}, $P_{2}$ is the Legendre polynomial of the second order and $\mathbf{v}_{i}$ is the bisecting unit vector of the $i$th water molecule in the lab frame.
387 gezelter 3897
388 plouden 3942 Here, we are evaluating this function across the $z$-dimension of the system as another measure of the change in the local environment and behavior of water molecules from the liquid region to the slushy interfacial region. After each of the 0.5 ns simulations, the systems were run for an additional 200 ps where the positions of every molecule in the system were recorded every 0.1 ps. The systems were then divided into 30 bins and the OTCF was evaluated for each bin.
389 plouden 3919
390 plouden 3942 It has been shown that the OTCF for water can be fit by a triexponential decay\cite{Furse08}, where the three components of the decay correspond to a fast (<200 fs) reorientational piece driven by the restoring forces of existing hydrogen bonds, a middle (on the order of several ps) piece describing the large angle jumps that occur during the breaking and formation of new hydrogen bonds\cite{Laage08,Laage11}, and a slow (on the order of tens of ps) contribution describing the translational motion of the molecules. The OTCF data for each bin were truncated at 100 ps, and fit to the triexponential decay
391 gezelter 3897 \begin{equation}
392 plouden 3942 C_{2}(t) \approx a_{1}e^{-t/\tau_{short}}+a_{2}e^{-t/\tau_{middle}}+a_{3}e^{-t/\tau_{long}}+a_{4}
393 gezelter 3897 \end{equation}
394 plouden 3941 where $a_{1}+a_{2}+a_{3}+a_{4}=1$. An average value and standard deviation for each $\tau$ was obtained for each bin from the four runs. Lastly, the means and standard deviations were averaged about the center of the system.
395 gezelter 3897
396 plouden 3941 \begin{figure}
397     \includegraphics[width=\linewidth]{basal_Tau_comic_strip}
398     \caption{\label{fig:basal_Tau_comic_strip} The orientational time correlation function for the basal system fit by the triexponential decay $a_{1}e^{-t/\tau_{short}}+a_{2}e^{-t/\tau_{middle}}+a_{3}e^{-t/\tau_{long}}+a_{4}$ where $a_{1}+a_{2}+a_{3}+a_{4}=1$. The verticle dotted line indicates the average midpoint of the interface as determined by the tetrahedrality fit. In (b) and (c) $\tau_{middle}$ and $\tau_{long}$ have a consistent value of about 5.5 ps and 50 ps in the liquid region, and increase in value approaching the interface. $\tau_{middle}$ corresponds to the breaking and making of hydrogen bonds as explained by extended jump model proposed by Laage and Hynes\cite{Laage08,Laage11} and $\tau_{long}$ relates to the translational motion of the molecules. In (a), we see that $\tau_{short}$ has a value of about 71 fs in the liquid region, and decreases in value approaching the interface. This component of the decay corresponds to the reorientational forces of existing hydrogen bonds on the inertial rotational of the molecules. Thus $\tau_{short}$ decreases approaching the interface due to the hindered range of motion of the molecules. }
399     \end{figure}
400 gezelter 3897
401 gezelter 3914 \begin{figure}
402 plouden 3941 \includegraphics[width=\linewidth]{prismatic_Tau_comic_strip}
403     \caption{\label{fig:prismatic_Tau_comic_strip} The orientational time correlation function for the prismatic system fit by the triexponential decay $a_{1}e^{-t/\tau_{short}}+a_{2}e^{-t/\tau_{middle}}+a_{3}e^{-t/\tau_{long}}+a_{4}$ where $a_{1}+a_{2}+a_{3}+a_{4}=1$. The verticle dotted line indicates the average midpoint of the interface as determined by the tetrahedrality fit. In (b) and (c) $\tau_{middle}$ and $\tau_{long}$ have a consistent value of about 3.5 ps and 30 ps in the liquid region, and increase in value approaching the interface. $\tau_{middle}$ corresponds to the breaking and making of hydrogen bonds as explained by extended jump model proposed by Laage and Hynes\cite{Laage08,Laage11} and $\tau_{long}$ relates to the translational motion of the molecules. In (a), we see that $\tau_{short}$ has a value of about 73 fs in the liquid region, and decreases in value approaching the interface. This component of the decay corresponds to the reorientational forces of existing hydrogen bonds on the inertial rotational of the molecules. Thus $\tau_{short}$ decreases approaching the interface due to the hindered range of motion of the molecules.}
404 gezelter 3914 \end{figure}
405 plouden 3904
406 plouden 3942 Figures \ref{fig:basal_Tau_comic_strip} and \ref{fig:prismatic_Tau_comic_strip} plot the decomposition of the OTCF at varying displacements from the center of the ice for the basal and prismatic systems. We see in (a) $\tau_{short}$, (b) $\tau_{middle}$, and (c) $\tau_{long}$ for the control system (no applied momentum flux) in black, and a system with a large shear rate in red. The verticle dotted lines at a displacement of about 17 \AA\ and 9 \AA\ denote the midpoints of the interfaces as determined by the hyperbolic tangent fit of the tetrahedrality profile.
407 plouden 3937
408 plouden 3942 In panels (a), we see at large displacements from the center of the ice $\tau_{short}$ for the basal system has a value of about 71 fs and 72 fs for the prismatic. Decreasing in displacement from about 26 \AA\ to about 19 \AA\ in the basal system, the value of $\tau_{short}$ decreases to about 63 fs. Likewise, $\tau_{short}$ decreases to about 63 fs from roughly 20 \AA\ to 12 \AA\ . This is due to the increasingly constrained motion of the water molecules as we approach the interface. In panels (b), $\tau_{middle}$ at large displacements from the ice has a value of about 5.5 ps and 3 ps for the basal and prismatic systems. We find $\tau_{middle}$ increases in value as we approach the interface in both cases. This component of the decay corresponds to the rearrangement of the hydrogen bonding network, which takes longer as the molecules motion becomes more constrained. In panels (c), $\tau_{long}$ has a value of about 50 ps for the basal system and roughly 30 ps for the prismatic system at large displacements from the interface. Similar to $\tau_{middle}$, $\tau_{long}$ also increases in value as we approach the interface for both systems. It is also apparent from Figures \ref{fig:basal_Tau_comic_strip} and \ref{fig:prismatic_Tau_comic_strip} that shearing the ice water has no effect on the orientational decay time, or on any of its decomposed components.
409 plouden 3937
410 plouden 3942 For each system, there is an apparent approximate value for $\tau_{short}$, $\tau_{middle}$, and $\tau_{long}$ at large displacements from the interface. There also appears to be a single displacement, $d_{basal}$ and $d_{prismatic}$, from the interface at which all three decay times begin to deviate from their bulk liquid values. We found $d_{basal}$ and $d_{prismatic}$ to be roughly 15 \AA\ and 8 \AA\ respectively. These two results indicate that the dynamics of the water molecules within $d_{basal}$ and $d_{prismatic}$ are being significantly perturbed by the ice and/or the interface, even though the structural width of the interface by analysis of the tetrahedrality profile indicates that bulk liquid structure of water is recovered after about 4 \AA\ from the edge of the ice.
411 plouden 3937
412 plouden 3942 Beaglehole and Wilson have measured the ice/water interface to have a thickness approximately 10 \AA\ for both the basal and prismatic face of ice by ellipticity measurements \cite{Beaglehole93}. Structurally, we have found the basal and prismatic interfacial width to be 3.2$\pm$0.4 \AA\ and 3.6$\pm$0.2 \AA\ . However, we have shown through decomposition of the OTCF a much larger interfacial region exists in which the dynamics of the water molecules behave differently than those of the bulk liquid.
413 plouden 3937
414 plouden 3941 \subsection{Coefficient of Friction of the Interface}
415 plouden 3942 As the ice is sheared through the liquid, there will be a friction between the solid and the liquid. Pit has shown how to calculate the coefficient of friction $\lambda$ for a solid-liquid interface for a Newtonian fluid of viscosity $\eta$ and has a slip length of $\delta$. \cite{Pit99}
416     \begin{equation}\label{Pit}
417     \lambda=\eta/\delta
418     \end{equation}
419     From linear response theory, $\eta$ can be obtained from the imposed momentum flux and the slope of the velocity about the dimension of the imposed flux.\cite{Kuang12}
420     \begin{equation}\label{Kuang}
421     j_{z}(p_{x})=-\eta\frac{\partial v_{x}}{\partial z}
422     \end{equation}
423     Solving eq. \eqref{Kuang} for $\eta$ and substituting the result into eq. \eqref{Pit}, we obtain an alternate expression for the coefficient of friction.
424 plouden 3941 \begin{equation}
425 plouden 3942 \lambda=-\frac{j_{z}(p_{x})}{\delta \frac{\partial v_{x}}{\partial z}}
426 plouden 3941 \end{equation}
427 plouden 3937
428 plouden 3942 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$0.007) amu fs\textsuperscript{-1}. It is known that the basal and prismatic faces have different surface structures. The basal face is smoother than the prismatic with small alternating valleys and crests, while the prismatic surface has deep corrugating channels. We believe the reason that the prismatic face's coefficient of friction was found to be smaller than the basal's is due to the direction of the shear. The shear of the ice/water was in the same direction of the corrugating channels, allowing water molecules to pass through the channels during the shear.
429 plouden 3939
430 plouden 3942 \begin{figure}
431     \includegraphics[width=\linewidth]{delta_example}
432     \caption{\label{fig:delta_example} A schematic of determining the slip length ($\delta$). The slip length is the difference of the structural starting point of the ice and the point of intersection of the linear regressions of the liquid phase velocity profile (red) and of the solid ice velocity profile (black). The dotted line indicates the location of the ice as determined by the tetrahedrality profile.}
433     \end{figure}
434    
435    
436 gezelter 3897 \section{Conclusion}
437 gezelter 3945 Here we have simulated the basal and prismatic facets of an SPC/E model of the ice I$_\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.
438 gezelter 3897
439 gezelter 3914 \begin{acknowledgement}
440     Support for this project was provided by the National Science
441     Foundation under grant CHE-0848243. Computational time was provided
442     by the Center for Research Computing (CRC) at the University of
443     Notre Dame.
444     \end{acknowledgement}
445 gezelter 3897
446 gezelter 3914 \newpage
447     \bibstyle{achemso}
448 plouden 3909 \bibliography{iceWater}
449 gezelter 3897
450 gezelter 3914 \begin{tocentry}
451     \begin{wrapfigure}{l}{0.5\textwidth}
452     \begin{center}
453     \includegraphics[width=\linewidth]{SystemImage.png}
454     \end{center}
455     \end{wrapfigure}
456     An image of our system.
457     \end{tocentry}
458 gezelter 3897
459 gezelter 3914 \end{document}
460 gezelter 3897
461 plouden 3924 %**************************************************************
462     %Non-mass weighted slopes (amu*Angstroms^-2 * fs^-1)
463     % basal: slope=0.090677616, error in slope = 0.003691743
464     %prismatic: slope = 0.050101506, error in slope = 0.001348181
465     %Mass weighted slopes (Angstroms^-2 * fs^-1)
466     %basal slope = 4.76598E-06, error in slope = 1.94037E-07
467     %prismatic slope = 3.23131E-06, error in slope = 8.69514E-08
468     %**************************************************************