ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/iceWater/iceWater.tex
Revision: 3919
Committed: Mon Jul 15 20:48:02 2013 UTC (11 years, 2 months ago) by plouden
Content type: application/x-tex
File size: 25265 byte(s)
Log Message:
intro

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     properties at interfaces. We present calculations of the interfacial
39     friction coefficients and the apparent independence of shear rate on
40     interfacial width and show that water moving over a flat ice/water
41     interface is close to the no-slip boundary condition.
42     \end{abstract}
43 gezelter 3897
44 gezelter 3914 \newpage
45 gezelter 3897
46     \section{Introduction}
47 plouden 3919 %-----Outline of Intro---------------
48     % in general, ice/water interface is important b/c ....
49     % here are some people who have worked on ice/water, trying to understand the processes above ....
50     % with the recent development of VSS-RNEMD, we can now look at the shearing problem
51     % talk about what we will present in this paper
52     % -------End Intro------------------
53 gezelter 3897
54 plouden 3919
55     %Gay02: cites many other ice/water papers, make sure to cite them.
56    
57     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}.
58    
59     Haymet \emph{et al.} have done extensive work on ice Ih, 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}. Other groups \cite{Baez95,Arbuckle02,}.
60    
61    
62    
63     %Other people looking at the ice/water interface, bio, crystal nucleation, kinetics, etc.
64 gezelter 3897 %Geologists are concerned with the flow of water over ice
65     %Antifreeze protein in fish--Haymet's group has cited this before
66    
67 plouden 3899 %Paragraph explaining why the ice/water interface is important
68     %Paragraph on what other people have done / lead into what hasn't been done
69     %Paragraph on what I'm going to do
70 plouden 3898
71    
72 gezelter 3897
73    
74 plouden 3902 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.
75 plouden 3898
76 plouden 3899 In this paper, we investigate the width and the friction coefficient of the ice/water interface as the ice is sheared through the liquid.
77    
78 gezelter 3897 \section{Methodology}
79    
80 gezelter 3914 \subsection{Stable ice I$_\mathrm{h}$ / water interfaces}
81    
82     The structure of ice I$_\mathrm{h}$ is well understood; it
83     crystallizes in a hexagonal space group P$6_3/mmc$, and the hexagonal
84     crystals of ice have two faces that are commonly exposed, the basal
85     face $\{0~0~0~1\}$, which forms the top and bottom of each hexagonal
86     plate, and the prismatic $\{1~0~\bar{1}~0\}$ face which forms the
87     sides of the plate. Other less-common, but still important, faces of
88     ice I$_\mathrm{h}$ are the secondary prism face, $\{1~1~\bar{2}~0\}$,
89     and the prismatic face, $\{2~0~\bar{2}~1\}$.
90    
91     Ice I$_\mathrm{h}$ is normally proton disordered in bulk crystals,
92     although the surfaces probably have a preference for proton ordering
93     along strips of dangling H-atoms and Oxygen lone
94     pairs.\cite{Buch:2008fk}
95    
96     For small simulated ice interfaces, it is useful to have a
97     proton-ordered, but zero-dipole crystal that exposes these strips of
98     dangling H-atoms and lone pairs. Also, if we're going to place
99     another material in contact with one of the ice crystalline planes, it
100     is useful to have an orthorhombic (rectangular) box to work with. A
101     recent paper by Hirsch and Ojam\"{a}e describes how to create
102     proton-ordered bulk ice I$_\mathrm{h}$ in alternative orthorhombic
103 gezelter 3915 cells.\cite{Hirsch04}
104 gezelter 3914
105     We have using Hirsch and Ojam\"{a}e's structure 6 which is an
106     orthorhombic cell ($P2_12_12_1$) that produces a proton-ordered
107     version of ice Ih. Table \ref{tab:equiv} contains a mapping between
108     the Miller indices in the P$6_3/mmc$ crystal system and those in the
109     Hirsch and Ojam\"{a}e $P2_12_12_1$ system.
110    
111     \begin{wraptable}{r}{3.5in}
112     \begin{tabular}{|ccc|} \hline
113     & hexagonal & orthorhombic \\
114     & ($P6_3/mmc$) & ($P2_12_12_1$) \\
115     crystal face & Miller indices & equivalent \\ \hline
116     basal & $\{0~0~0~1\}$ & $\{0~0~1\}$ \\
117     prism & $\{1~0~\bar{1}~0\}$ & $\{1~0~0\}$ \\
118     secondary prism & $\{1~1~\bar{2}~0\}$ & $\{1~3~0\}$ \\
119     pryamidal & $\{2~0~\bar{2}~1\}$ & $\{2~0~1\}$ \\ \hline
120     \end{tabular}
121     \end{wraptable}
122    
123     Structure 6 from the Hirsch and Ojam\"{a}e paper has lattice
124     parameters $a = 4.49225$, $b = 7.78080$, $c = 7.33581$ and two water
125     molecules whose atoms reside at the following fractional coordinates:
126    
127     \begin{wraptable}{r}{3.25in}
128     \begin{tabular}{|ccccc|} \hline
129     atom label & type & x & y & z \\ \hline
130     O$_{a}$ & O & 0.75 & 0.1667 & 0.4375 \\
131     H$_{1a}$ & H & 0.5735 & 0.2202 & 0.4836 \\
132     H$_{2a}$ & H & 0.7420 & 0.0517 & 0.4836 \\
133     O$_{b}$ & O & 0.25 & 0.6667 & 0.4375 \\
134     H$_{1b}$ & H & 0.2580 & 0.6693 & 0.3071 \\
135     H$_{2b}$ & H & 0.4265 & 0.7255 & 0.4756 \\ \hline
136     \end{tabular}
137     \end{wraptable}
138    
139     To construct the basal and prismatic interfaces, the crystallographic
140     coordinates above were used to construct an orthorhombic unit cell
141     which was then replicated in all three dimensions yielding a
142     proton-ordered block of ice I$_{h}$. To expose the desired face, the
143     orthorhombic representation was then cut along the ($001$) or ($100$)
144     planes for the basal and prismatic faces respectively. The resulting
145     block was rotated so that the exposed faces were aligned with the $z$
146     axis normal to the exposed face. The block was then cut along two
147     perpendicular directions in a way that allowed for perfect periodic
148     replication in the $x$ and $y$ axes, creating a slab with either the
149     basal or prismatic faces exposed along the $z$ axis. The slab was
150     then replicated in the $x$ and $y$ dimensions until a desired sample
151     size was obtained.
152    
153     Although experimental solid/liquid coexistant temperature under normal
154     pressure are close to 273K, Haymet \emph{et al.} have done extensive
155     work on characterizing the ice/water
156     interface.\cite{Karim88,Karim90,Hayward01,Bryk02,Hayward02} They have
157     found for the SPC/E water model,\cite{Berendsen87} which is also used
158     in this study, the ice/water interface is most stable at
159     225$\pm$5K.\cite{Bryk02} To create a ice / water interface, a box of
160     liquid water that had the same dimensions in $x$ and $y$ was
161     equilibrated at 225 K and 1 atm of pressure in the NPAT ensemble (with
162     the $z$ axis allowed to fluctuate to equilibrate to the correct
163     pressure). The liquid and solid systems were combined by carving out
164     any water molecule from the liquid simulation cell that was within 3
165     \AA\ of any atom in the ice slab.
166    
167     Molecular translation and orientational restraints were applied in the
168     early stages of equilibration to prevent melting of the ice slab.
169     These restraints were removed during NVT equilibration, well before
170     data collection was carried out.
171    
172 gezelter 3918 \subsection{Shearing ice / water interfaces without bulk melting}
173    
174     As one drags a solid through a liquid, there will be frictional
175     heating that will act to melt the interface. To study the frictional
176     behavior of the interface without causing the interface to melt, it is
177     necessary to apply a weak thermal gradient along with the momentum
178     gradient. This can be accomplished with of the newly-developed
179     approaches to reverse non-equilibrium molecular dynamics (RNEMD). The
180     velocity shearing and scaling (VSS) variant of RNEMD utilizes a series
181     of simultaneous velocity exchanges between two regions within the
182     simulation cell.\cite{Kuang12} One of these regions is centered within
183     the ice slab, while the other is centrally located in the liquid phase
184     region. VSS-RNEMD provides a set of conservation constraints for
185     simultaneously creating either a momentum flux or a thermal flux (or
186     both) between the two slabs. Satisfying the constraint equations
187     ensures that the new configurations are sampled from the same NVE
188     ensemble as previously.
189    
190     The VSS moves are applied periodically to scale and shift the particle
191     velocities ($\mathbf{v}_i$ and $\mathbf{v}_j$) in two slabs ($H$ and
192     $C$) which are separated by half of the simulation box,
193     \begin{displaymath}
194     \begin{array}{rclcl}
195    
196     & \underline{\mathrm{shearing}} & &
197     \underline{~~~~~~~~~~~~\mathrm{scaling}~~~~~~~~~~~~} \\
198     \mathbf{v}_i \leftarrow &
199     \mathbf{a}_c\ & + & c\cdot\left(\mathbf{v}_i - \langle\mathbf{v}_c
200     \rangle\right) + \langle\mathbf{v}_c\rangle \\
201     \mathbf{v}_j \leftarrow &
202     \mathbf{a}_h & + & h\cdot\left(\mathbf{v}_j - \langle\mathbf{v}_h
203     \rangle\right) + \langle\mathbf{v}_h\rangle .
204    
205     \end{array}
206     \end{displaymath}
207     Here $\langle\mathbf{v}_c\rangle$ and $\langle\mathbf{v}_h\rangle$ are
208     the center of mass velocities in the $C$ and $H$ slabs, respectively.
209     Within the two slabs, particles receive incremental changes or a
210     ``shear'' to their velocities. The amount of shear is governed by the
211     imposed momentum flux, $\mathbf{j}_z(\mathbf{p})$
212     \begin{eqnarray}
213     \mathbf{a}_c & = & - \mathbf{j}_z(\mathbf{p}) \Delta t / M_c \label{vss1}\\
214     \mathbf{a}_h & = & + \mathbf{j}_z(\mathbf{p}) \Delta t / M_h \label{vss2}
215     \end{eqnarray}
216     where $M_{\{c,h\}}$ is the total mass of particles within each of the
217     slabs and $\Delta t$ is the interval between two separate operations.
218    
219     To simultaneously impose a thermal flux ($J_z$) between the slabs we
220     use energy conservation constraints,
221     \begin{eqnarray}
222     K_c - J_z\Delta t & = & c^2 (K_c - \frac{1}{2}M_c \langle\mathbf{v}_c
223     \rangle^2) + \frac{1}{2}M_c (\langle \mathbf{v}_c \rangle + \mathbf{a}_c)^2 \label{vss3}\\
224     K_h + J_z\Delta t & = & h^2 (K_h - \frac{1}{2}M_h \langle\mathbf{v}_h
225     \rangle^2) + \frac{1}{2}M_h (\langle \mathbf{v}_h \rangle +
226     \mathbf{a}_h)^2 \label{vss4}.
227     \label{constraint}
228     \end{eqnarray}
229     Simultaneous solution of these quadratic formulae for the scaling
230     coefficients, $c$ and $h$, will ensure that the simulation samples from
231     the original microcanonical (NVE) ensemble. Here $K_{\{c,h\}}$ is the
232     instantaneous translational kinetic energy of each slab. At each time
233     interval, it is a simple matter to solve for $c$, $h$, $\mathbf{a}_c$,
234     and $\mathbf{a}_h$, subject to the imposed momentum flux,
235     $j_z(\mathbf{p})$, and thermal flux, $J_z$, values. Since the VSS
236     operations do not change the kinetic energy due to orientational
237     degrees of freedom or the potential energy of a system, configurations
238     after the VSS move have exactly the same energy (and linear
239     momentum) as before the move.
240    
241     As the simulation progresses, the VSS moves are performed on a regular
242     basis, and the system develops a thermal and/or velocity gradient in
243     response to the applied flux. In a bulk material it is quite simple
244     to use the slope of the temperature or velocity gradients to obtain
245     the thermal conductivity or shear viscosity.
246    
247     The VSS-RNEMD approach is versatile in that it may be used to
248     implement thermal and shear transport simultaneously. Perturbations
249     of velocities away from the ideal Maxwell-Boltzmann distributions are
250     minimal, as is thermal anisotropy. This ability to generate
251     simultaneous thermal and shear fluxes has been previously utilized to
252     map out the shear viscosity of SPC/E water over a wide range of
253     temperatures (90~K) with a single 1 ns simulation.\cite{Kuang12}
254    
255     Here we are using this method primarily to generate a shear between
256     the ice slab and the liquid phase, while using a weak thermal gradient
257     to maintaining the interface at the 225K target value. This will
258     insure minimal melting of the bulk ice phase and allows us to control
259     the exact temperature of the interface.
260    
261 gezelter 3897 \subsection{Computational Details}
262 gezelter 3918 All simulations were performed using OpenMD with a time step of 2 fs,
263     and periodic boundary conditions in all three dimensions. The systems
264     were divided into 100 artificial bins along the $z$-axis for the
265     VSS-RNEMD moves, which were attempted every 50 fs. The gradients were
266     allowed to develop for 1 ns before data collection was began. Once
267     established, snapshots of the system were taken every 1 ps, and the
268     average velocities and densities of each bin were accumulated every
269     attempted VSS-RNEMD move.
270 gezelter 3897
271     %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.
272     %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.
273     %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.
274     %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.
275    
276     \subsection{Measuring the Width of the Interface}
277 gezelter 3918 In order to characterize the ice/water interface, the local
278     tetrahedral order parameter as described by Kumar\cite{Kumar09} and
279     Errington\cite{Errington01} was used. The local tetrahedral order
280     parameter, $q$, is given by
281 gezelter 3897 \begin{equation}
282     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
283     \end{equation}
284 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.
285    
286     %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.
287    
288 gezelter 3897 The system was divided into 100 bins along the $z$-axis, and a $q$ value was determined for each snapshot of the system for each bin. The $q$ values for each bin were then averaged to give an average tetrahedrally profile of the system about the $z$- axis. The profile was then fit with a hyperbolic tangent function given by
289     \begin{equation}
290     q_{z}=q_{liq}+\frac{q_{ice}-q_{liq}}{2}(\tanh(\alpha(z-I_{L,m}))-\tanh(\alpha(z-I_{R,m})))
291     \end{equation}
292     where $q_{liq}$ and $q_{ice}$ are fitting parameters for the local tetrahedral order parameter for the liquid and ice, $\alpha$ is proportional to the inverse of the width of the interface, and $I_{L,m}$ and $I_{R,m}$ are the midpoints of the left and right interface. During the simulations where a kinetic energy flux was imposed, there was found to be a thermal influence in the liquid phase region of the tetrahedrally profile due to the thermal gradient developed in the system. To maximize the fit of the interface, another term was added to the hyperbolic tangent fitting function,
293     \begin{equation}
294     q_{z}=q_{liq}+\frac{q_{ice}-q_{liq}}{2}(\tanh(\alpha(z-I_{L,m}))-\tanh(\alpha(z-I_{R,m})))+\beta|(z-z_{mid})|
295     \end{equation}
296     where $\beta$ is a fitting parameter and $z_{mid}$ is the midpoint of the z dimension of the simulation box.
297    
298    
299     \section{Results and discussion}
300 plouden 3898
301     %Images to include: 3-long comic strip style of <Vx>, T, q_z as a function of z for the basal and prismatic faces. q_z by z with fit for basal and prismatic. interface width as a function of deltaVx (shear rate) with basal and prismatic on the same plot, error bars in the x and y. <Vx> by flux with basal and prismatic on same graph, back out slope from xmgr and error in slope to get lambda, friction coefficient of interface.
302    
303 plouden 3909 In Figures \ref{fig:bComic} and \ref{fig:pComic} we see the $z$-dimensional profiles for several parameters for the basal and prismatic systems respectively. 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. 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 can be found 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$5K, can be seen in panel (b). Lastly, the velocity gradient across the systems can be seen in panel (c). From panel (c), we can see liquid phase water molecules 5 to 12 \AA\ from the midpoint of the basal and prismatic interfaces are being dragged along with the ice block. This indicates that the shearing of ice water is in the stick boundary condition.
304 plouden 3904
305 gezelter 3914 \begin{figure}
306     \includegraphics[width=\linewidth]{bComicStrip}
307     \caption{\label{fig:bComic} The basal system: (a) The local tetrahedral order parameter,$q$, (black circles) fit by a hyperbolic tangent (red line), (b) the thermal gradient imposed on the system to maintain a stable interfacial temperature, and (c) the velocity gradient imposed on the system. The verticle dotted lines indicate the midpoint of the interfaces.}
308     \end{figure}
309 plouden 3904
310 gezelter 3914 %(a) The local tetrahedral order parameter across the z-dimension of the system (black circles) fit by a hyperbolic tangent (red line). (b) The thermal gradient imposed on the system to maintain a stable interfacial temperature. (c) The velocity gradient imposed on the system. The verticle dotted lines indicate the midpoint of the interfaces.
311    
312     \begin{figure}
313     \includegraphics[width=\linewidth]{pComicStrip}
314     \caption{\label{fig:pComic} The prismatic system: (a) The local tetrahedral order parameter,$q$, (black circles) fit by a hyperbolic tangent (red line), (b) the thermal gradient imposed on the system to maintain a stable interfacial temperature, and (c) the velocity gradient imposed on the system.The verticle dotted lines indicate the midpoint of the interfaces.}
315     \end{figure}
316    
317    
318 gezelter 3897 \subsection{Interfacial Width}
319 plouden 3909 %For the basal and prismatic systems, the ice blocks were sheared through the water at varying rates while an imposed thermal gradient kept the interface at the stable temperature range as described by Byrk and Haymet.
320     We found the interfacial width for the basal and prismatic systems to be 3.2$\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}, 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.
321 gezelter 3897
322 plouden 3919 %Need to reword the following paragraph
323     Beaglehole and Wilson have measured the ice/water interface to have a thickness approximately 10-20 \AA\ for both the basal and prismatic face of ice by ellipticity measurements \cite{Beaglehole93}. Haymet \emph{et al.} agrees with these measurements, our results do not. We are using a parameter from the hyperbolic tangent fit of the local tetrahedrality order parameter to determine the interfacial width, whereas Haymet and co-workers use the 10-90 widths of the translational, average density, diffusion, and orientational decay times \cite{Hayward01}.
324    
325 gezelter 3897 \subsection{Coefficient of Friction of the Interface}
326 gezelter 3915 As the ice is sheared through the liquid, there will be a friction between the ice and the interface. Balasubramanian has shown how to calculate the coefficient of friction for a solid-liquid interface. \cite{Balasubramanian99}
327 gezelter 3897 \begin{equation}
328     %<F_{x}^{w}>_{NE}(t)=-S\lambda_{wall}v_{x}(y_{wall})
329     \langle F_{x}^{w}\rangle(t)=-S\lambda_{wall}v_{x}(y_{wall})
330     \end{equation}
331     In this equation, $F_{x}^{w}$ is the total force of all the atoms acting on the fluid, $S$ is the surface area the force is being applied upon, and $\lambda_{wall}$ is the coefficient of friction of the interface. Since the imposed momentum flux, $J_{z}(p_{x})$, is known in the VSS-RNEMD simulations, and the $wall$ is the ice block in our simulations, the above equation can be rewritten as
332     \begin{equation}
333     J_{z}(p_{x})=-\lambda_{ice}v_{x}(y_{ice}).
334     \end{equation}
335    
336 plouden 3909 In Figure \ref{fig:CoeffFric}, the average velocity of the ice is plotted against the imposed momentum flux for the basal (black circles) and prismatic (red circles) systems. From the equation above, the slope of the linear fit of the data is $\lambda_{wall}$. The coefficient of friction of the interface for the basal face was calculated to be 11.0 $\pm$ 0.4 \AA\textsuperscript{-2}fs\textsuperscript{-1}, and the $\lambda_{wall}$ for the prismatic face was determined to be 19.9, $\pm$ 0.5 \AA\textsuperscript{-2}fs\textsuperscript{-1}.
337 gezelter 3897
338 plouden 3904 %Ask dan about truncating versus rounding the values for lambda.
339 plouden 3907 %The coefficient of friction of the interface for the basal face was calculated to be 11.02808 $\pm$ 0.4489844 \AA^{-2}fs^{-1}, and the $\lambda_{wall}$ for the prismatic face was determined to be 19.95948, $\pm$ 0.5370894 \AA^{-2}fs^{-1}.
340 gezelter 3914 \begin{figure}
341     \includegraphics[width=\linewidth]{CoeffFric}
342     \caption{\label{fig:CoeffFric} The average velocity of the ice for the basal (black circles) and prismatic (red circles) systems as a function off applied momentum flux. The slope of the fit line }
343     \end{figure}
344 plouden 3904
345 gezelter 3897 \section{Conclusion}
346 plouden 3909 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 appearant independence of the shear rate on the interfacial width. The coefficient of friction of the interface was also calculated for each of the facets. The $\lambda_{wall}$ for the basal face was calculated to be 11.0 $\pm$ 0.4 \AA\textsuperscript{-2}fs\textsuperscript{-1}, and 19.9, $\pm$ 0.5 \AA\textsuperscript{-2}fs\textsuperscript{-1} for the prismatic facet. For both facets, the shearing ice water was found to be in the no-slip boundary condition.
347 gezelter 3897
348 plouden 3909
349 gezelter 3914 \begin{acknowledgement}
350     Support for this project was provided by the National Science
351     Foundation under grant CHE-0848243. Computational time was provided
352     by the Center for Research Computing (CRC) at the University of
353     Notre Dame.
354     \end{acknowledgement}
355 gezelter 3897
356 gezelter 3914 \newpage
357     \bibstyle{achemso}
358 plouden 3909 \bibliography{iceWater}
359 gezelter 3897
360 gezelter 3914 \begin{tocentry}
361     \begin{wrapfigure}{l}{0.5\textwidth}
362     \begin{center}
363     \includegraphics[width=\linewidth]{SystemImage.png}
364     \end{center}
365     \end{wrapfigure}
366     An image of our system.
367     \end{tocentry}
368 gezelter 3897
369 gezelter 3914 \end{document}
370 gezelter 3897
371 plouden 3904 % basal: slope=11.02808, error in slope = 0.4489844
372     %prismatic: slope = 19.95948, error in slope = 0.5370894