ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/iceWater/iceWater.tex
Revision: 3942
Committed: Fri Aug 30 18:13:50 2013 UTC (11 years ago) by plouden
Content type: application/x-tex
File size: 35155 byte(s)
Log Message:
revised again, handing this copy to Dan at group meeting

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 3914 \title{Do the facets of ice $I_\mathrm{h}$ crystals have different
15     friction coefficients? Simulating shear in ice/water interfaces}
16 gezelter 3897
17     \author{P. B. Louden}
18 gezelter 3914 \author{J. Daniel Gezelter}
19     \email{gezelter@nd.edu}
20     \affiliation[University of Notre Dame]{251 Nieuwland Science Hall\\
21     Department of Chemistry and Biochemistry\\ University of Notre
22     Dame\\ Notre Dame, Indiana 46556}
23 gezelter 3897
24 gezelter 3914 \keywords{}
25 gezelter 3897
26 gezelter 3914 \begin{document}
27 gezelter 3897
28 gezelter 3914 \begin{abstract}
29     We have investigated the structural properties of the basal and
30     prismatic facets of an SPC/E model of the ice Ih / water interface
31     when the solid phase is being drawn through liquid water (i.e. sheared
32     relative to the fluid phase). To impose the shear, we utilized a
33     reverse non-equilibrium molecular dynamics (RNEMD) method that creates
34     non-equilibrium conditions using velocity shearing and scaling (VSS)
35     moves of the molecules in two physically separated slabs in the
36     simulation cell. This method can create simultaneous temperature and
37     velocity gradients and allow the measurement of friction transport
38 plouden 3942 properties at interfaces. We found the interfacial width for the basal and prismatic facets to be 3.2$\pm$0.4 \AA\ and 3.6$\pm$0.2 \AA\ with no momentum gradient present. The interfacial width was found to be independent of shear rate over the range of shear rates investigated, 0.6$\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. The orientational time correlation function was evaluated at varying displacements from the interface, and fit to a triexponential decay with decay times $\tau_{short}$, $\tau_{middle}$, and $\tau_{long}$. We found that each decay time was independent of the shear rate, and that they each deviated from a bulk liquid value at the same displacement $d_{basal}$ and $d_{prismatic}$. We determined $d_{basal}$ and $d_{prismatic}$ to be approximately 15 \AA\ and 8 \AA\ . Both $\tau_{middle}$ and $\tau_{long}$ increased in value in decreasing displacements from $d_{basal}$ and $d_{prismatic}$, while we found that $\tau_{short}$ had the inverse correlation. Lastly we present calculations of the friction coefficients ($\lambda$) for the basal and prismatic facets. We have found $\lambda_{basal}$ to be (0.07$\pm$0.01) amu fs\textsuperscript{-1} and $\lambda_{prismatic}$ to be (0.032$\pm$0.007) amu fs\textsuperscript{-1}.
39 gezelter 3914 \end{abstract}
40 gezelter 3897
41 gezelter 3914 \newpage
42 gezelter 3897
43     \section{Introduction}
44 plouden 3919 %-----Outline of Intro---------------
45     % in general, ice/water interface is important b/c ....
46     % here are some people who have worked on ice/water, trying to understand the processes above ....
47     % with the recent development of VSS-RNEMD, we can now look at the shearing problem
48     % talk about what we will present in this paper
49     % -------End Intro------------------
50 gezelter 3897
51 plouden 3919 %Gay02: cites many other ice/water papers, make sure to cite them.
52    
53 plouden 3920 Understanding the ice/water interface is essential for explaining complex processes such as nucleartion and crystal growth\cite{Han92,Granasy95,Vanfleet95}, crystal melting\cite{Weber83,Han92,Sakai96,Sakai96B}, and biological interfacial processes, such as the antifreeze protein found in winter flounder\cite{Wierzbicki07, Chapsky97}. These processes have been studied at the fundamental level of the ice/water interface by several groups, including studying the structure and width of the interface. Haymet \emph{et al.} have done extensive work on ice Ih, the most common form of ice on Earth, including characterizing and determining the width of the ice/water interface for the SPC, SPC/E, CF1, and TIP4P models for water. \cite{Karim87,Karim88,Karim90,Hayward01,Bryk02,Hayward02,Gay02} More recently, Haymet \emph{et al.} have been investigating the effects cations and anions have on crystal nucleation\cite{Bryk04,Smith05,Wilson08,Wilson10}. Nada \emph{et al.} have also studied the ice/water interface\cite{Nada95,Nada00,Nada03,Nada12}. They have found that the different facets of ice Ih have different growth rates, primarily, that the prismatic facet grows faster than the basal facet due to the mechanism of the crystal growth being the reordering of the hydrogen bonding network\cite{Nada05}.
54 plouden 3919
55 plouden 3920 Another complex process which requires investigation at the ice/water interface is the movement of water over ice, such as icebergs floating in the ocean. In addition to understanding the structure and width of the interface, it is pertinent to understand the friction caused by the shearing of water across the ice to understand this process. However, until recently, simulations of this nature were not possible.
56     With the recent development of velocity shearing and scaling reverse non-equilibrium molecular dynamics (VSS-RNEMD), it is now possible to calculate transport properties from heterogeneous systems.\cite{Kuang12} This method can create simultaneous temperature and velocity gradients and allow the measurement of friction and thermal transport properties at interfaces. This allows for the study of the width of the ice/water interface as the ice is sheared through the liquid, while imposing a thermal gradient to prevent frictional heating of the interface. In this paper, we investigate the width and the friction coefficient of the ice/water interface as the ice is sheared through the liquid under a weak thermal gradient.
57 plouden 3919
58 gezelter 3897 \section{Methodology}
59    
60 gezelter 3914 \subsection{Stable ice I$_\mathrm{h}$ / water interfaces}
61    
62     The structure of ice I$_\mathrm{h}$ is well understood; it
63     crystallizes in a hexagonal space group P$6_3/mmc$, and the hexagonal
64     crystals of ice have two faces that are commonly exposed, the basal
65     face $\{0~0~0~1\}$, which forms the top and bottom of each hexagonal
66     plate, and the prismatic $\{1~0~\bar{1}~0\}$ face which forms the
67     sides of the plate. Other less-common, but still important, faces of
68     ice I$_\mathrm{h}$ are the secondary prism face, $\{1~1~\bar{2}~0\}$,
69     and the prismatic face, $\{2~0~\bar{2}~1\}$.
70    
71     Ice I$_\mathrm{h}$ is normally proton disordered in bulk crystals,
72     although the surfaces probably have a preference for proton ordering
73     along strips of dangling H-atoms and Oxygen lone
74     pairs.\cite{Buch:2008fk}
75    
76     For small simulated ice interfaces, it is useful to have a
77     proton-ordered, but zero-dipole crystal that exposes these strips of
78     dangling H-atoms and lone pairs. Also, if we're going to place
79     another material in contact with one of the ice crystalline planes, it
80     is useful to have an orthorhombic (rectangular) box to work with. A
81     recent paper by Hirsch and Ojam\"{a}e describes how to create
82     proton-ordered bulk ice I$_\mathrm{h}$ in alternative orthorhombic
83 gezelter 3915 cells.\cite{Hirsch04}
84 gezelter 3914
85 plouden 3941 We are using Hirsch and Ojam\"{a}e's structure 6 which is an
86 gezelter 3914 orthorhombic cell ($P2_12_12_1$) that produces a proton-ordered
87     version of ice Ih. Table \ref{tab:equiv} contains a mapping between
88     the Miller indices in the P$6_3/mmc$ crystal system and those in the
89     Hirsch and Ojam\"{a}e $P2_12_12_1$ system.
90    
91     \begin{wraptable}{r}{3.5in}
92     \begin{tabular}{|ccc|} \hline
93     & hexagonal & orthorhombic \\
94     & ($P6_3/mmc$) & ($P2_12_12_1$) \\
95     crystal face & Miller indices & equivalent \\ \hline
96     basal & $\{0~0~0~1\}$ & $\{0~0~1\}$ \\
97     prism & $\{1~0~\bar{1}~0\}$ & $\{1~0~0\}$ \\
98     secondary prism & $\{1~1~\bar{2}~0\}$ & $\{1~3~0\}$ \\
99     pryamidal & $\{2~0~\bar{2}~1\}$ & $\{2~0~1\}$ \\ \hline
100     \end{tabular}
101     \end{wraptable}
102    
103     Structure 6 from the Hirsch and Ojam\"{a}e paper has lattice
104     parameters $a = 4.49225$, $b = 7.78080$, $c = 7.33581$ and two water
105     molecules whose atoms reside at the following fractional coordinates:
106    
107     \begin{wraptable}{r}{3.25in}
108     \begin{tabular}{|ccccc|} \hline
109     atom label & type & x & y & z \\ \hline
110     O$_{a}$ & O & 0.75 & 0.1667 & 0.4375 \\
111     H$_{1a}$ & H & 0.5735 & 0.2202 & 0.4836 \\
112     H$_{2a}$ & H & 0.7420 & 0.0517 & 0.4836 \\
113     O$_{b}$ & O & 0.25 & 0.6667 & 0.4375 \\
114     H$_{1b}$ & H & 0.2580 & 0.6693 & 0.3071 \\
115     H$_{2b}$ & H & 0.4265 & 0.7255 & 0.4756 \\ \hline
116     \end{tabular}
117     \end{wraptable}
118    
119     To construct the basal and prismatic interfaces, the crystallographic
120     coordinates above were used to construct an orthorhombic unit cell
121     which was then replicated in all three dimensions yielding a
122     proton-ordered block of ice I$_{h}$. To expose the desired face, the
123     orthorhombic representation was then cut along the ($001$) or ($100$)
124     planes for the basal and prismatic faces respectively. The resulting
125 plouden 3942 block was rotated so that the exposed faces were aligned with the $z$-dimension normal to the exposed face. The block was then cut along two
126 gezelter 3914 perpendicular directions in a way that allowed for perfect periodic
127     replication in the $x$ and $y$ axes, creating a slab with either the
128     basal or prismatic faces exposed along the $z$ axis. The slab was
129     then replicated in the $x$ and $y$ dimensions until a desired sample
130     size was obtained.
131    
132     Although experimental solid/liquid coexistant temperature under normal
133     pressure are close to 273K, Haymet \emph{et al.} have done extensive
134     work on characterizing the ice/water
135     interface.\cite{Karim88,Karim90,Hayward01,Bryk02,Hayward02} They have
136     found for the SPC/E water model,\cite{Berendsen87} which is also used
137     in this study, the ice/water interface is most stable at
138     225$\pm$5K.\cite{Bryk02} To create a ice / water interface, a box of
139     liquid water that had the same dimensions in $x$ and $y$ was
140     equilibrated at 225 K and 1 atm of pressure in the NPAT ensemble (with
141     the $z$ axis allowed to fluctuate to equilibrate to the correct
142     pressure). The liquid and solid systems were combined by carving out
143     any water molecule from the liquid simulation cell that was within 3
144     \AA\ of any atom in the ice slab.
145    
146     Molecular translation and orientational restraints were applied in the
147     early stages of equilibration to prevent melting of the ice slab.
148     These restraints were removed during NVT equilibration, well before
149     data collection was carried out.
150    
151 gezelter 3918 \subsection{Shearing ice / water interfaces without bulk melting}
152    
153     As one drags a solid through a liquid, there will be frictional
154     heating that will act to melt the interface. To study the frictional
155     behavior of the interface without causing the interface to melt, it is
156     necessary to apply a weak thermal gradient along with the momentum
157     gradient. This can be accomplished with of the newly-developed
158     approaches to reverse non-equilibrium molecular dynamics (RNEMD). The
159     velocity shearing and scaling (VSS) variant of RNEMD utilizes a series
160     of simultaneous velocity exchanges between two regions within the
161     simulation cell.\cite{Kuang12} One of these regions is centered within
162     the ice slab, while the other is centrally located in the liquid phase
163     region. VSS-RNEMD provides a set of conservation constraints for
164     simultaneously creating either a momentum flux or a thermal flux (or
165     both) between the two slabs. Satisfying the constraint equations
166     ensures that the new configurations are sampled from the same NVE
167     ensemble as previously.
168    
169     The VSS moves are applied periodically to scale and shift the particle
170     velocities ($\mathbf{v}_i$ and $\mathbf{v}_j$) in two slabs ($H$ and
171     $C$) which are separated by half of the simulation box,
172     \begin{displaymath}
173     \begin{array}{rclcl}
174    
175     & \underline{\mathrm{shearing}} & &
176     \underline{~~~~~~~~~~~~\mathrm{scaling}~~~~~~~~~~~~} \\
177     \mathbf{v}_i \leftarrow &
178     \mathbf{a}_c\ & + & c\cdot\left(\mathbf{v}_i - \langle\mathbf{v}_c
179     \rangle\right) + \langle\mathbf{v}_c\rangle \\
180     \mathbf{v}_j \leftarrow &
181     \mathbf{a}_h & + & h\cdot\left(\mathbf{v}_j - \langle\mathbf{v}_h
182     \rangle\right) + \langle\mathbf{v}_h\rangle .
183    
184     \end{array}
185     \end{displaymath}
186     Here $\langle\mathbf{v}_c\rangle$ and $\langle\mathbf{v}_h\rangle$ are
187     the center of mass velocities in the $C$ and $H$ slabs, respectively.
188     Within the two slabs, particles receive incremental changes or a
189     ``shear'' to their velocities. The amount of shear is governed by the
190     imposed momentum flux, $\mathbf{j}_z(\mathbf{p})$
191     \begin{eqnarray}
192     \mathbf{a}_c & = & - \mathbf{j}_z(\mathbf{p}) \Delta t / M_c \label{vss1}\\
193     \mathbf{a}_h & = & + \mathbf{j}_z(\mathbf{p}) \Delta t / M_h \label{vss2}
194     \end{eqnarray}
195     where $M_{\{c,h\}}$ is the total mass of particles within each of the
196     slabs and $\Delta t$ is the interval between two separate operations.
197    
198     To simultaneously impose a thermal flux ($J_z$) between the slabs we
199     use energy conservation constraints,
200     \begin{eqnarray}
201     K_c - J_z\Delta t & = & c^2 (K_c - \frac{1}{2}M_c \langle\mathbf{v}_c
202     \rangle^2) + \frac{1}{2}M_c (\langle \mathbf{v}_c \rangle + \mathbf{a}_c)^2 \label{vss3}\\
203     K_h + J_z\Delta t & = & h^2 (K_h - \frac{1}{2}M_h \langle\mathbf{v}_h
204     \rangle^2) + \frac{1}{2}M_h (\langle \mathbf{v}_h \rangle +
205     \mathbf{a}_h)^2 \label{vss4}.
206     \label{constraint}
207     \end{eqnarray}
208     Simultaneous solution of these quadratic formulae for the scaling
209     coefficients, $c$ and $h$, will ensure that the simulation samples from
210     the original microcanonical (NVE) ensemble. Here $K_{\{c,h\}}$ is the
211     instantaneous translational kinetic energy of each slab. At each time
212     interval, it is a simple matter to solve for $c$, $h$, $\mathbf{a}_c$,
213     and $\mathbf{a}_h$, subject to the imposed momentum flux,
214     $j_z(\mathbf{p})$, and thermal flux, $J_z$, values. Since the VSS
215     operations do not change the kinetic energy due to orientational
216     degrees of freedom or the potential energy of a system, configurations
217     after the VSS move have exactly the same energy (and linear
218     momentum) as before the move.
219    
220     As the simulation progresses, the VSS moves are performed on a regular
221     basis, and the system develops a thermal and/or velocity gradient in
222     response to the applied flux. In a bulk material it is quite simple
223     to use the slope of the temperature or velocity gradients to obtain
224     the thermal conductivity or shear viscosity.
225    
226     The VSS-RNEMD approach is versatile in that it may be used to
227     implement thermal and shear transport simultaneously. Perturbations
228     of velocities away from the ideal Maxwell-Boltzmann distributions are
229     minimal, as is thermal anisotropy. This ability to generate
230     simultaneous thermal and shear fluxes has been previously utilized to
231     map out the shear viscosity of SPC/E water over a wide range of
232     temperatures (90~K) with a single 1 ns simulation.\cite{Kuang12}
233    
234     Here we are using this method primarily to generate a shear between
235     the ice slab and the liquid phase, while using a weak thermal gradient
236     to maintaining the interface at the 225K target value. This will
237     insure minimal melting of the bulk ice phase and allows us to control
238     the exact temperature of the interface.
239    
240 gezelter 3897 \subsection{Computational Details}
241 gezelter 3918 All simulations were performed using OpenMD with a time step of 2 fs,
242     and periodic boundary conditions in all three dimensions. The systems
243     were divided into 100 artificial bins along the $z$-axis for the
244     VSS-RNEMD moves, which were attempted every 50 fs. The gradients were
245     allowed to develop for 1 ns before data collection was began. Once
246 plouden 3941 established, four successive 0.5 ns runs were performed for each shear rate. During these simulations, snapshots of the system were taken every 1 ps, and the
247 gezelter 3918 average velocities and densities of each bin were accumulated every
248     attempted VSS-RNEMD move.
249 gezelter 3897
250     %A paragraph on the equilibration procedure of the system? Shenyu did some amount of equilibration to the files and then I was handed them. I performed 5 ns of NVT at 225K for both systems, then 5 ns of NVE at 225K for both systems, with no gradients imposed.
251     %For the basal, once the thermal gradient was found which gave me the interfacial temperature I wanted (-2.0E-6 kcal/mol/A^2/fs), I equilibrated the file for 5 ns letting this gradient stabilize. Then I continued to use this thermal gradient as I imposed momentum gradients and watched the response of the interface.
252     %For the prismatic, a gradient was not found that would give me the interfacial temperature I desired, so while imposing a thermal gradient that had the interface at 220K, I raised the temperature of the system to 230K. This resulted in a thermal gradient which gave my interface at 225K, equilibrated for ins NVT, then ins NVE while this gradient was still imposed, then I began dragging.
253     %I have run each system for 1 ns under PTgrads to allow them to develop, then ran each system for an additional 2 ns in segments of 0.5 ns in order to calculate statistics of the calculated values.
254    
255 plouden 3941 \section{Results and discussion}
256    
257 gezelter 3897 \subsection{Measuring the Width of the Interface}
258 plouden 3941 \subsubsection{Tetrahedrality Order Parameter}
259     Any parameter or function that varies across the interface from a bulk liquid value to a solid value can be used as a measure of the width of the interface. However, due to the VSS-RNEMD moves pertrurbing the momentum of the molecules, parameters such as the translational order parameter and the diffusion order parameter may be artifically skewed. A structural parameter such as the pairwise correlation function would not be influenced by the perturbations. Here, the local order tetraherdal parameter as described by Kumar\cite{Kumar09} and
260     Errington\cite{Errington01} was used as a measure of the interfacial width.
261 plouden 3936
262     The local tetrahedral order parameter, $q$, is given by
263 gezelter 3897 \begin{equation}
264     q_{k} \equiv 1 -\frac{3}{8}\sum_{i=1}^{3} \sum_{j=i+1}^{4} \Bigg[\cos\psi_{ikj}+\frac{1}{3}\Bigg]^2
265     \end{equation}
266 plouden 3909 where $\psi_{ikj}$ is the angle formed by the oxygen sites on molecule $k$, and the oxygen site on its two closest neighbors, molecules $i$ and $j$. The local tetrahedral order parameter function has a range of (0,1), where the larger the value $q$ has the more tetrahedral the ordering of the local environment is. A $q$ value of one describes a perfectly tetrahedral environment relative to it and its four nearest neighbors, and the parameter's value decreases as the local ordering becomes less tetrahedral.
267    
268     %If the central water molecule has a perfect tetrahedral geometry with its four nearest neighbors, the parameter goes to one, and decreases to zero as the geometry deviates from the ideal configuration.
269    
270 plouden 3942 The system was divided into 100 bins of length $L$ along the $z$-dimension, and a cutoff radius for the neighboring molecules was set to 3.41 \AA\ . A $q_{z}$ value was then determined by averaging the $q$ values for each molecule in the bin.
271 gezelter 3897 \begin{equation}
272 plouden 3921 q_{z} \equiv \int_0^L \Bigg[1 -\frac{3}{8}\sum_{i=1}^{3} \sum_{j=i+1}^{4} \bigg[\cos\psi_{ikj}+\frac{1}{3}\bigg]^2\Bigg]\delta(z_{k}-z)\mathrm{d}z
273 gezelter 3897 \end{equation}
274 plouden 3942 The $q_{z}$ values for each snapshot were then averaged to give a tetrahedrality profile of the system about the $z$-dimension. The profile was then fit with a hyperbolic tangent function given by
275 plouden 3941
276     \begin{equation}\label{tet_fit}
277 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})|
278 gezelter 3897 \end{equation}
279    
280 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.
281 gezelter 3897
282 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$5K, can be seen in panel (b). Lastly, the velocity gradient across the systems can be seen in panel (c). From panel (c), we can see liquid phase water molecules 10 \AA\ to 15 \AA\ from the interfaces are being dragged along with the ice block, indicating that the shearing of ice water is in the stick boundary condition.
283 plouden 3898
284 gezelter 3914 \begin{figure}
285     \includegraphics[width=\linewidth]{bComicStrip}
286     \caption{\label{fig:bComic} The basal system: (a) The local tetrahedral order parameter,$q$, (black circles) fit by a hyperbolic tangent (red line), (b) the thermal gradient imposed on the system to maintain a stable interfacial temperature, and (c) the velocity gradient imposed on the system. The verticle dotted lines indicate the midpoint of the interfaces.}
287     \end{figure}
288 plouden 3904
289 gezelter 3914 \begin{figure}
290     \includegraphics[width=\linewidth]{pComicStrip}
291     \caption{\label{fig:pComic} The prismatic system: (a) The local tetrahedral order parameter,$q$, (black circles) fit by a hyperbolic tangent (red line), (b) the thermal gradient imposed on the system to maintain a stable interfacial temperature, and (c) the velocity gradient imposed on the system.The verticle dotted lines indicate the midpoint of the interfaces.}
292     \end{figure}
293    
294 plouden 3942 From the tetrahedrality fits, we found the interfacial width for the basal and prismatic systems to be 3.2$\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.
295 gezelter 3914
296 plouden 3941 \subsubsection{Orientational Time Correlation Function}
297     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.
298     \begin{equation}\label{C(t)1}
299     C_{2}(t)=\langle P_{2}(\mathbf{v}_{i}(t)\mathbf{v}_{i}(t=0))\rangle
300     \end{equation}
301 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.
302 gezelter 3897
303 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.
304 plouden 3919
305 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
306 gezelter 3897 \begin{equation}
307 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}
308 gezelter 3897 \end{equation}
309 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.
310 gezelter 3897
311 plouden 3941 \begin{figure}
312     \includegraphics[width=\linewidth]{basal_Tau_comic_strip}
313     \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. }
314     \end{figure}
315 gezelter 3897
316 gezelter 3914 \begin{figure}
317 plouden 3941 \includegraphics[width=\linewidth]{prismatic_Tau_comic_strip}
318     \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.}
319 gezelter 3914 \end{figure}
320 plouden 3904
321 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.
322 plouden 3937
323 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.
324 plouden 3937
325 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.
326 plouden 3937
327 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.
328 plouden 3937
329 plouden 3941 \subsection{Coefficient of Friction of the Interface}
330 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}
331     \begin{equation}\label{Pit}
332     \lambda=\eta/\delta
333     \end{equation}
334     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}
335     \begin{equation}\label{Kuang}
336     j_{z}(p_{x})=-\eta\frac{\partial v_{x}}{\partial z}
337     \end{equation}
338     Solving eq. \eqref{Kuang} for $\eta$ and substituting the result into eq. \eqref{Pit}, we obtain an alternate expression for the coefficient of friction.
339 plouden 3941 \begin{equation}
340 plouden 3942 \lambda=-\frac{j_{z}(p_{x})}{\delta \frac{\partial v_{x}}{\partial z}}
341 plouden 3941 \end{equation}
342 plouden 3937
343 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$0.007) amu fs\textsuperscript{-1}. It is known that the basal and prismatic faces have different surface structures. The basal face is smoother than the prismatic with small alternating valleys and crests, while the prismatic surface has deep corrugating channels. We believe the reason that the prismatic face's coefficient of friction was found to be smaller than the basal's is due to the direction of the shear. The shear of the ice/water was in the same direction of the corrugating channels, allowing water molecules to pass through the channels during the shear.
344 plouden 3939
345 plouden 3942 \begin{figure}
346     \includegraphics[width=\linewidth]{delta_example}
347     \caption{\label{fig:delta_example} A schematic of determining the slip length ($\delta$). The slip length is the difference of the structural starting point of the ice and the point of intersection of the linear regressions of the liquid phase velocity profile (red) and of the solid ice velocity profile (black). The dotted line indicates the location of the ice as determined by the tetrahedrality profile.}
348     \end{figure}
349    
350    
351 gezelter 3897 \section{Conclusion}
352 plouden 3942 Here we have simulated the basal and prismatic facets of an SPC/E model of the ice Ih / water interface. Using VSS-RNEMD, the ice was sheared relative to the liquid while imposed thermal gradients kept the interface at a stable temperature. Caculation of the local tetrahedrality order parameter has shown an apparent independence of the shear rate on the interfacial width, which was found to be 3.2$\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.
353 gezelter 3897
354 gezelter 3914 \begin{acknowledgement}
355     Support for this project was provided by the National Science
356     Foundation under grant CHE-0848243. Computational time was provided
357     by the Center for Research Computing (CRC) at the University of
358     Notre Dame.
359     \end{acknowledgement}
360 gezelter 3897
361 gezelter 3914 \newpage
362     \bibstyle{achemso}
363 plouden 3909 \bibliography{iceWater}
364 gezelter 3897
365 gezelter 3914 \begin{tocentry}
366     \begin{wrapfigure}{l}{0.5\textwidth}
367     \begin{center}
368     \includegraphics[width=\linewidth]{SystemImage.png}
369     \end{center}
370     \end{wrapfigure}
371     An image of our system.
372     \end{tocentry}
373 gezelter 3897
374 gezelter 3914 \end{document}
375 gezelter 3897
376 plouden 3924 %**************************************************************
377     %Non-mass weighted slopes (amu*Angstroms^-2 * fs^-1)
378     % basal: slope=0.090677616, error in slope = 0.003691743
379     %prismatic: slope = 0.050101506, error in slope = 0.001348181
380     %Mass weighted slopes (Angstroms^-2 * fs^-1)
381     %basal slope = 4.76598E-06, error in slope = 1.94037E-07
382     %prismatic slope = 3.23131E-06, error in slope = 8.69514E-08
383     %**************************************************************