ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/iceWater/iceWater.tex
Revision: 3946
Committed: Wed Sep 4 20:54:02 2013 UTC (10 years, 9 months ago) by gezelter
Content type: application/x-tex
File size: 35544 byte(s)
Log Message:
Added PDFs and edited up through page 13.

File Contents

# User Rev Content
1 gezelter 3914 \documentclass[journal = jpccck, manuscript = article]{achemso}
2     \setkeys{acs}{usetitle = true}
3     \usepackage{achemso}
4     \usepackage{natbib}
5     \usepackage{multirow}
6     \usepackage{wrapfig}
7     \usepackage{fixltx2e}
8     %\mciteErrorOnUnknownfalse
9 gezelter 3897
10 gezelter 3914 \usepackage[version=3]{mhchem} % this is a great package for formatting chemical reactions
11     \usepackage{url}
12 gezelter 3897
13 gezelter 3945 \title{Solid-liquid friction at ice-I$_\mathrm{h}$ / water interfaces}
14 gezelter 3897
15     \author{P. B. Louden}
16 gezelter 3914 \author{J. Daniel Gezelter}
17     \email{gezelter@nd.edu}
18     \affiliation[University of Notre Dame]{251 Nieuwland Science Hall\\
19     Department of Chemistry and Biochemistry\\ University of Notre
20     Dame\\ Notre Dame, Indiana 46556}
21 gezelter 3897
22 gezelter 3914 \keywords{}
23 gezelter 3897
24 gezelter 3914 \begin{document}
25 gezelter 3897
26 gezelter 3914 \begin{abstract}
27 gezelter 3945 We have investigated the structural and dynamic properties of the
28     basal and prismatic facets of an ice I$_\mathrm{h}$ / water
29     interface when the solid phase is being drawn through the liquid
30     (i.e. sheared relative to the fluid phase). To impose the shear, we
31     utilized a velocity-shearing and scaling (VSS) approach to reverse
32     non-equilibrium molecular dynamics (RNEMD). This method can create
33     simultaneous temperature and velocity gradients and allow the
34     measurement of transport properties at interfaces. The interfacial
35     width was found to be independent of relative velocity of the ice
36     and liquid layers over a wide range of shear rates. Decays of
37     molecular orientational time correlation functions for gave very
38     similar estimates for the width of the interfaces, although the
39     short- and longer-time decay components of the orientational
40     correlation functions behave differently closer to the interface.
41     Although both facets of ice are in ``stick'' boundary conditions in
42     liquid water, the solid-liquid friction coefficient was found to be
43     different for the basal and prismatic facets of ice.
44 gezelter 3914 \end{abstract}
45 gezelter 3897
46 gezelter 3914 \newpage
47 gezelter 3897
48     \section{Introduction}
49 plouden 3919 %-----Outline of Intro---------------
50     % in general, ice/water interface is important b/c ....
51     % here are some people who have worked on ice/water, trying to understand the processes above ....
52     % with the recent development of VSS-RNEMD, we can now look at the shearing problem
53     % talk about what we will present in this paper
54     % -------End Intro------------------
55 gezelter 3897
56 plouden 3919 %Gay02: cites many other ice/water papers, make sure to cite them.
57    
58 gezelter 3945 Understanding the ice/water interface is essential for explaining
59     complex processes such as nucleation and crystal
60     growth,\cite{Han92,Granasy95,Vanfleet95} crystal
61     melting,\cite{Weber83,Han92,Sakai96,Sakai96B} and some fascinating
62     biological processes, such as the behavior of the antifreeze proteins
63     found in winter flounder,\cite{Wierzbicki07, Chapsky97} and certain
64     terrestrial arthropods.\cite{Duman:2001qy,Meister29012013} There has
65     been significant progress on understanding the structure and dynamics
66     of quiescent ice/water interfaces utilizing both theory and
67     experiment. Haymet \emph{et al.} have done extensive work on ice I$_\mathrm{h}$,
68     including characterizing and determining the width of the ice/water
69     interface for the SPC, SPC/E, CF1, and TIP4P models for
70     water.\cite{Karim87,Karim88,Karim90,Hayward01,Bryk02,Hayward02,Gay02}
71     More recently, Haymet \emph{et al.} have investigated the effects
72     cations and anions have on crystal
73     nucleation\cite{Bryk04,Smith05,Wilson08,Wilson10}. Nada \emph{et al.}
74     have also studied ice/water
75     interfaces,\cite{Nada95,Nada00,Nada03,Nada12} and have found that the
76     differential growth rates of the facets of ice I$_\mathrm{h}$ are due to the the
77     reordering of the hydrogen bonding network\cite{Nada05}.
78 plouden 3919
79 gezelter 3945 The movement of liquid water over the facets of ice has been less
80     thoroughly studied than the quiescent surfaces. This process is
81     potentially important in understanding transport of large blocks of
82     ice in water (which has important implications in the earth sciences),
83     as well as the relative motion of crystal-crystal interfaces that have
84     been separated by nanometer-scale fluid domains. In addition to
85     understanding both the structure and thickness of the interfacial
86     regions, it is important to understand the molecular origin of
87     friction, drag, and other changes in dynamical properties of the
88     liquid in the regions close to the surface that are altered by the
89     presence of a shearing of the bulk fluid relative to the solid phase.
90 plouden 3919
91 gezelter 3945 In this work, we apply a recently-developed velocity shearing and
92     scaling approach to reverse non-equilibrium molecular dynamics
93     (VSS-RNEMD). This method makes it possible to calculate transport
94     properties like the interfacial thermal conductance across
95     heterogeneous interfaces,\cite{Kuang12} and can create simultaneous
96     temperature and velocity gradients and allow the measurement of
97     friction and thermal transport properties at interfaces. This has
98     allowed us to investigate the width of the ice/water interface as the
99     ice is sheared through the liquid, while simultaneously imposing a
100     weak thermal gradient to prevent frictional heating of the interface.
101     In the sections that follow, we discuss the methodology for creating
102     and simulating ice/water interfaces under shear and provide results
103     from both structural and dynamical correlation functions. We also
104     show that the solid-liquid interfacial friction coefficient depends
105     sensitively on the details of the surface morphology.
106    
107 gezelter 3897 \section{Methodology}
108    
109 gezelter 3945 \subsection{Stable ice I$_\mathrm{h}$ / water interfaces under shear}
110 gezelter 3914
111 gezelter 3946 The structure of ice I$_\mathrm{h}$ is well understood; it
112 gezelter 3914 crystallizes in a hexagonal space group P$6_3/mmc$, and the hexagonal
113     crystals of ice have two faces that are commonly exposed, the basal
114     face $\{0~0~0~1\}$, which forms the top and bottom of each hexagonal
115     plate, and the prismatic $\{1~0~\bar{1}~0\}$ face which forms the
116     sides of the plate. Other less-common, but still important, faces of
117 gezelter 3945 ice I$_\mathrm{h}$ are the secondary prism, $\{1~1~\bar{2}~0\}$, and
118 gezelter 3946 pyramidal, $\{2~0~\bar{2}~1\}$, faces. Ice I$_\mathrm{h}$ is normally
119     proton disordered in bulk crystals, although the surfaces probably
120     have a preference for proton ordering along strips of dangling H-atoms
121     and Oxygen lone pairs.\cite{Buch:2008fk}
122 gezelter 3914
123     \begin{wraptable}{r}{3.5in}
124 gezelter 3945 \caption{Mapping between the Miller indices of four facets of ice in
125     the $P6_3/mmc$ crystal system to the orthorhombic $P2_12_12_1$
126     system in reference \protect\cite{Hirsch04}}
127     \label{tab:equiv}
128 gezelter 3914 \begin{tabular}{|ccc|} \hline
129     & hexagonal & orthorhombic \\
130     & ($P6_3/mmc$) & ($P2_12_12_1$) \\
131     crystal face & Miller indices & equivalent \\ \hline
132     basal & $\{0~0~0~1\}$ & $\{0~0~1\}$ \\
133     prism & $\{1~0~\bar{1}~0\}$ & $\{1~0~0\}$ \\
134     secondary prism & $\{1~1~\bar{2}~0\}$ & $\{1~3~0\}$ \\
135 gezelter 3946 pyramidal & $\{2~0~\bar{2}~1\}$ & $\{2~0~1\}$ \\ \hline
136 gezelter 3914 \end{tabular}
137     \end{wraptable}
138    
139 gezelter 3946 For small simulated ice interfaces, it is useful to work with
140     proton-ordered, but zero-dipole crystal that exposes these strips of
141     dangling H-atoms and lone pairs. When placing another material in
142     contact with one of the ice crystalline planes, it is also quite
143     useful to have an orthorhombic (rectangular) box. Recent work by
144     Hirsch and Ojam\"{a}e describes a number of alternative crystal
145     systems for proton-ordered bulk ice I$_\mathrm{h}$ using orthorhombic
146     cells.\cite{Hirsch04}
147    
148     In this work, we are using Hirsch and Ojam\"{a}e's structure 6 which
149     is an orthorhombic cell ($P2_12_12_1$) that produces a proton-ordered
150     version of ice I$_\mathrm{h}$. Table \ref{tab:equiv} contains a
151     mapping between the Miller indices of common ice facets in the
152     P$6_3/mmc$ crystal system and those in the Hirsch and Ojam\"{a}e
153     $P2_12_12_1$ system.
154    
155 gezelter 3914 Structure 6 from the Hirsch and Ojam\"{a}e paper has lattice
156     parameters $a = 4.49225$, $b = 7.78080$, $c = 7.33581$ and two water
157 gezelter 3945 molecules whose atoms reside at fractional coordinates given in table
158     \ref{tab:p212121}. To construct the basal and prismatic interfaces,
159     these crystallographic coordinates were used to construct an
160     orthorhombic unit cell which was then replicated in all three
161     dimensions yielding a proton-ordered block of ice I$_{h}$. To expose
162     the desired face, the orthorhombic representation was then cut along
163     the ($001$) or ($100$) planes for the basal and prismatic faces
164     respectively. The resulting block was rotated so that the exposed
165 gezelter 3946 faces were aligned with the $z$-axis normal to the exposed face. The
166     block was then cut along two perpendicular directions in a way that
167     allowed for perfect periodic replication in the $x$ and $y$ axes,
168 gezelter 3945 creating a slab with either the basal or prismatic faces exposed along
169     the $z$ axis. The slab was then replicated in the $x$ and $y$
170     dimensions until a desired sample size was obtained.
171 gezelter 3914
172 gezelter 3946 \begin{wraptable}{r}{2.85in}
173 gezelter 3945 \caption{Fractional coordinates for water in the orthorhombic
174     $P2_12_12_1$ system for ice I$_\mathrm{h}$ in reference
175     \protect\cite{Hirsch04}}
176     \label{tab:p212121}
177     \begin{tabular}{|cccc|} \hline
178     atom type & x & y & z \\ \hline
179     O & 0.75 & 0.1667 & 0.4375 \\
180     H & 0.5735 & 0.2202 & 0.4836 \\
181     H & 0.7420 & 0.0517 & 0.4836 \\
182     O & 0.25 & 0.6667 & 0.4375 \\
183     H & 0.2580 & 0.6693 & 0.3071 \\
184     H & 0.4265 & 0.7255 & 0.4756 \\ \hline
185 gezelter 3914 \end{tabular}
186     \end{wraptable}
187    
188 gezelter 3946 Our ice / water interfaces were created using a box of liquid water
189     that had the same dimensions (in $x$ and $y$) as the ice block.
190     Although the experimental solid/liquid coexistence temperature under
191     atmospheric pressure is close to 273K, Haymet \emph{et al.} have done
192     extensive work on characterizing the ice/water interface, and find
193     that the coexistence temperature for simulated water is often quite a
194     bit different.\cite{Karim88,Karim90,Hayward01,Bryk02,Hayward02} They
195     have found that for the SPC/E water model,\cite{Berendsen87} which is
196     also used in this study, the ice/water interface is most stable at
197     225$\pm$5K.\cite{Bryk02} This liquid box was therefore equilibrated at
198     225 K and 1 atm of pressure in the NPAT ensemble (with the $z$ axis
199     allowed to fluctuate to equilibrate to the correct pressure). The
200     liquid and solid systems were combined by carving out any water
201     molecule from the liquid simulation cell that was within 3 \AA\ of any
202     atom in the ice slab.
203 gezelter 3914
204     Molecular translation and orientational restraints were applied in the
205     early stages of equilibration to prevent melting of the ice slab.
206     These restraints were removed during NVT equilibration, well before
207 gezelter 3945 data collection was carried out.
208 gezelter 3914
209 gezelter 3918 \subsection{Shearing ice / water interfaces without bulk melting}
210    
211 gezelter 3945 As a solid is dragged through a liquid, there is frictional heating
212     that will act to melt the interface. To study the behavior of the
213     interface under a shear stress without causing the interface to melt,
214 gezelter 3946 it is necessary to apply a weak thermal gradient in combination with
215     the momentum gradient. This can be accomplished using the velocity
216 gezelter 3945 shearing and scaling (VSS) variant of reverse non-equilibrium
217     molecular dynamics (RNEMD), which utilizes a series of simultaneous
218     velocity exchanges between two regions within the simulation
219     cell.\cite{Kuang12} One of these regions is centered within the ice
220 gezelter 3946 slab, while the other is centrally located in the liquid
221 gezelter 3918 region. VSS-RNEMD provides a set of conservation constraints for
222 gezelter 3946 creating either a momentum flux or a thermal flux (or both
223     simultaneously) between the two slabs. Satisfying the constraint
224     equations ensures that the new configurations are sampled from the
225     same NVE ensemble as before the VSS move.
226 gezelter 3918
227     The VSS moves are applied periodically to scale and shift the particle
228     velocities ($\mathbf{v}_i$ and $\mathbf{v}_j$) in two slabs ($H$ and
229     $C$) which are separated by half of the simulation box,
230     \begin{displaymath}
231     \begin{array}{rclcl}
232    
233     & \underline{\mathrm{shearing}} & &
234     \underline{~~~~~~~~~~~~\mathrm{scaling}~~~~~~~~~~~~} \\
235     \mathbf{v}_i \leftarrow &
236     \mathbf{a}_c\ & + & c\cdot\left(\mathbf{v}_i - \langle\mathbf{v}_c
237     \rangle\right) + \langle\mathbf{v}_c\rangle \\
238     \mathbf{v}_j \leftarrow &
239     \mathbf{a}_h & + & h\cdot\left(\mathbf{v}_j - \langle\mathbf{v}_h
240     \rangle\right) + \langle\mathbf{v}_h\rangle .
241    
242     \end{array}
243     \end{displaymath}
244     Here $\langle\mathbf{v}_c\rangle$ and $\langle\mathbf{v}_h\rangle$ are
245     the center of mass velocities in the $C$ and $H$ slabs, respectively.
246     Within the two slabs, particles receive incremental changes or a
247     ``shear'' to their velocities. The amount of shear is governed by the
248     imposed momentum flux, $\mathbf{j}_z(\mathbf{p})$
249     \begin{eqnarray}
250     \mathbf{a}_c & = & - \mathbf{j}_z(\mathbf{p}) \Delta t / M_c \label{vss1}\\
251     \mathbf{a}_h & = & + \mathbf{j}_z(\mathbf{p}) \Delta t / M_h \label{vss2}
252     \end{eqnarray}
253     where $M_{\{c,h\}}$ is the total mass of particles within each of the
254     slabs and $\Delta t$ is the interval between two separate operations.
255    
256     To simultaneously impose a thermal flux ($J_z$) between the slabs we
257     use energy conservation constraints,
258     \begin{eqnarray}
259     K_c - J_z\Delta t & = & c^2 (K_c - \frac{1}{2}M_c \langle\mathbf{v}_c
260     \rangle^2) + \frac{1}{2}M_c (\langle \mathbf{v}_c \rangle + \mathbf{a}_c)^2 \label{vss3}\\
261     K_h + J_z\Delta t & = & h^2 (K_h - \frac{1}{2}M_h \langle\mathbf{v}_h
262     \rangle^2) + \frac{1}{2}M_h (\langle \mathbf{v}_h \rangle +
263     \mathbf{a}_h)^2 \label{vss4}.
264     \label{constraint}
265     \end{eqnarray}
266     Simultaneous solution of these quadratic formulae for the scaling
267     coefficients, $c$ and $h$, will ensure that the simulation samples from
268     the original microcanonical (NVE) ensemble. Here $K_{\{c,h\}}$ is the
269     instantaneous translational kinetic energy of each slab. At each time
270     interval, it is a simple matter to solve for $c$, $h$, $\mathbf{a}_c$,
271     and $\mathbf{a}_h$, subject to the imposed momentum flux,
272     $j_z(\mathbf{p})$, and thermal flux, $J_z$, values. Since the VSS
273     operations do not change the kinetic energy due to orientational
274     degrees of freedom or the potential energy of a system, configurations
275     after the VSS move have exactly the same energy (and linear
276     momentum) as before the move.
277    
278     As the simulation progresses, the VSS moves are performed on a regular
279     basis, and the system develops a thermal and/or velocity gradient in
280 gezelter 3945 response to the applied flux. In a bulk material, it is quite simple
281 gezelter 3918 to use the slope of the temperature or velocity gradients to obtain
282 gezelter 3945 either the thermal conductivity or shear viscosity.
283 gezelter 3918
284     The VSS-RNEMD approach is versatile in that it may be used to
285     implement thermal and shear transport simultaneously. Perturbations
286     of velocities away from the ideal Maxwell-Boltzmann distributions are
287     minimal, as is thermal anisotropy. This ability to generate
288     simultaneous thermal and shear fluxes has been previously utilized to
289     map out the shear viscosity of SPC/E water over a wide range of
290     temperatures (90~K) with a single 1 ns simulation.\cite{Kuang12}
291    
292 gezelter 3945 For this work, we are using the VSS-RNEMD method primarily to generate
293     a shear between the ice slab and the liquid phase, while using a weak
294 gezelter 3946 thermal gradient to maintain the interface at the 225K target
295     value. This ensures minimal melting of the bulk ice phase and allows
296     us to control the exact temperature of the interface.
297 gezelter 3918
298 gezelter 3897 \subsection{Computational Details}
299 gezelter 3945 All simulations were performed using OpenMD,\cite{OOPSE,openmd} with a
300     time step of 2 fs and periodic boundary conditions in all three
301     dimensions. Electrostatics were handled using the damped-shifted
302     force real-space electrostatic kernel.\cite{Ewald} The systems were
303     divided into 100 bins along the $z$-axis for the VSS-RNEMD moves,
304     which were attempted every 50 fs.
305 gezelter 3897
306 gezelter 3945 The interfaces were equilibrated for a total of 10 ns at equilibrium
307     conditions before being exposed to either a shear or thermal gradient.
308     This consisted of 5 ns under a constant temperature (NVT) integrator
309     set to 225K followed by 5 ns under a microcanonical integrator. Weak
310     thermal gradients were allowed to develop using the VSS-RNEMD (NVE)
311     integrator using a a small thermal flux ($-2.0\times 10^{-6}$
312     kcal/mol/\AA$^2$/fs) for a duration of 5 ns to allow the gradient to
313 gezelter 3946 stabilize. The resulting temperature gradient was $\approx$ 10K over
314     the entire 100 \AA\ box length, which was sufficient to keep the
315 gezelter 3945 temperature at the interface within $\pm 1$ K of the 225K target.
316 gezelter 3897
317 gezelter 3945 Velocity gradients were then imposed using the VSS-RNEMD (NVE)
318     integrator with a range of momentum fluxes. These gradients were
319     allowed to stabilize for 1 ns before data collection began. Once
320     established, four successive 0.5 ns runs were performed for each shear
321     rate. During these simulations, snapshots of the system were taken
322     every 1 ps, and statistics on the structure and dynamics in each bin
323     were accumulated throughout the simulations.
324    
325 plouden 3941 \section{Results and discussion}
326    
327 gezelter 3946 \subsection{Interfacial width}
328     Any order parameter or time correlation function that changes as one
329     crosses an interface from a bulk liquid to a solid can be used to
330     measure the width of the interface. In previous work on the ice/water
331     interface, Haymet {\it et al.}\cite{} have utilized structural
332     features (including the density) as well as dynamic properties
333     (including the diffusion constant) to estimate the width of the
334     interfaces for a number of facets of the ice crystals. Because
335     VSS-RNEMD imposes a lateral flow, parameters that depend on
336     translational motion of the molecules (e.g. diffusion) may be
337     artifically skewed by the RNEMD moves. A structural parameter is not
338     influenced by the RNEMD perturbations to the same degree. Here, we
339     have used the local tetraherdal order parameter as described by
340     Kumar\cite{Kumar09} and Errington\cite{Errington01} as our principal
341 gezelter 3945 measure of the interfacial width.
342 plouden 3936
343 gezelter 3945 The local tetrahedral order parameter, $q(z)$, is given by
344 gezelter 3897 \begin{equation}
345 gezelter 3946 q(z) = \int_0^L \sum_{k=1}^{N} \Bigg(1 -\frac{3}{8}\sum_{i=1}^{3}
346     \sum_{j=i+1}^{4} \bigg(\cos\psi_{ikj}+\frac{1}{3}\bigg)^2\Bigg)
347     \delta(z_{k}-z)\mathrm{d}z \Bigg/ N_z
348 gezelter 3945 \label{eq:qz}
349 gezelter 3897 \end{equation}
350 gezelter 3945 where $\psi_{ikj}$ is the angle formed between the oxygen site on
351 gezelter 3946 central molecule $k$, and the oxygen sites on two of the four closest
352     molecules, $i$ and $j$. Molecules $i$ and $j$ are further restricted
353     to lie within the first solvation shell of molecule $k$. $N_z = \int
354     \delta(z_k - z) \mathrm{d}z$ is a normalization factor to account for
355     the varying population of molecules within each finite-width bin. The
356     local tetrahedral order parameter has a range of $(0,1)$, where the
357     larger values of $q$ indicate a larger degree of tetrahedral ordering
358     of the local environment. In perfect ice I$_\mathrm{h}$ structures,
359     the parameter can approach 1 at low temperatures, while in liquid
360     water, the ordering is significantly less tetrahedral, and values of
361     $q(z) \approx 0.75$ are more common.
362 plouden 3909
363 gezelter 3946 To estimate the interfacial width, the system was divided into 100
364     bins along the $z$-dimension, and a cutoff radius for the first
365     solvation shell was set to 3.41 \AA\ . The $q_{z}$ function was
366     time-averaged to give yield a tetrahedrality profile of the
367     system. The profile was then fit to a hyperbolic tangent that smoothly
368     links the liquid and solid states,
369 plouden 3941 \begin{equation}\label{tet_fit}
370 gezelter 3946 q(z) \approx
371     q_{liq}+\frac{q_{ice}-q_{liq}}{2}\left[\tanh\left(\frac{z-l}{w}\right)-\tanh\left(\frac{z-r}{w}\right)\right]+\beta\left|z-
372     \frac{r+l}{2}\right|.
373 gezelter 3897 \end{equation}
374 gezelter 3946 Here $q_{liq}$ and $q_{ice}$ are the local tetrahedral order parameter
375     for the bulk liquid and ice domains, respectively, $w$ is the width of
376     the interface. $l$ and $r$ are the midpoints of the left and right
377     interfaces, respectively. The last term in equation \eqref{tet_fit}
378     accounts for the influence that the weak thermal gradient has on the
379     tetrahedrality profile in the liquid region. To estimate the
380     10\%-90\% widths commonly used in previous studies,\cite{} it is a
381     simple matter to scale the widths obtained from the hyberbolic tangent
382     fits to obtain $w_{10-90} = 2.9 w$.\cite{}
383 gezelter 3897
384 gezelter 3946 In Figures \ref{fig:bComic} and \ref{fig:pComic} we see the
385     $z$-coordinate profiles for tetrahedrality, temperature, and the
386     $x$-component of the velocity for the basal and prismatic interfaces.
387     The lower panels show the $q(z)$ (black circles) along with the
388     hyperbolic tangent fits (red lines). In the liquid region, the local
389     tetrahedral order parameter, $q(z) \approx 0.75$ while in the
390     crystalline region, $q(z) \approx 0.94$, indicating a more tetrahedral
391     environment. The vertical dotted lines denote the midpoint of the
392     interfaces ($r$ and $l$ in equation \eqref{tet_fit}). The weak thermal
393     gradient applied to the systems in order to keep the interface at
394     225$\pm$5K, can be seen in middle panels. The tranverse velocity
395     profile is shown in the upper panels. It is clear from the upper
396     panels that water molecules in close proximity to the surface (i.e.
397     within 10 \AA\ to 15 \AA\ of the interfaces) have transverse
398     velocities quite close to the velocities within the ice block. There
399     is no velocity discontinuity at the interface, which indicates that
400     the shearing of ice/water interfaces occurs in the ``stick'' or
401     no-slip boundary conditions.
402 gezelter 3897
403 gezelter 3914 \begin{figure}
404 gezelter 3946 \includegraphics[width=\linewidth]{bComicStrip.pdf}
405     \caption{\label{fig:bComic} The basal interfaces. Lower panel: the
406     local tetrahedral order parameter, $q(z)$, (black circles) and the
407     hyperbolic tangent fit (red line). Middle panel: the imposed
408     thermal gradient required to maintain a fixed interfacial
409     temperature. Upper panel: the transverse velocity gradient that
410     develops in response to an imposed momentum flux. The vertical
411     dotted lines indicate the locations of the midpoints of the two
412     interfaces.}
413 gezelter 3914 \end{figure}
414 plouden 3904
415 gezelter 3914 \begin{figure}
416 gezelter 3946 \includegraphics[width=\linewidth]{pComicStrip.pdf}
417     \caption{\label{fig:pComic} The prismatic interfaces. Panel
418     descriptions match those in figure \ref{fig:bComic}}
419 gezelter 3914 \end{figure}
420    
421 gezelter 3946 From the fits using equation \eqref{tet_fit}, we find the interfacial
422     width for the basal and prismatic systems to be 3.2$\pm$0.4 \AA\ and
423     3.6$\pm$0.2 \AA\ , respectively, with no applied momentum flux. Over
424     the range of shear rates investigated, $0.6 \pm 0.3 \mathrm{ms}^{-1}
425     \rightarrow 5.3 \pm 0.5 \mathrm{ms}^{-1}$ for the basal system and
426     $0.9 \pm 0.2 \mathrm{ms}^{-1} \rightarrow 4.5 \pm 0.1
427     \mathrm{ms}^{-1}$ for the prismatic, we found no appreciable change in
428     the interface width. The fit values for the interfacial width ($w$)
429     over all shear rates contained the values reported above within their
430     error bars.
431 gezelter 3914
432 plouden 3941 \subsubsection{Orientational Time Correlation Function}
433 gezelter 3946 The orientational time correlation function,
434 plouden 3941 \begin{equation}\label{C(t)1}
435 gezelter 3946 C_{2}(t)=\langle P_{2}(\mathbf{u}(0) \cdot \mathbf{u}(t)) \rangle,
436 plouden 3941 \end{equation}
437 gezelter 3946 gives insight into the local dynamic environment around the water
438     molecules. The rate at which the function decays provides information
439     about hindered motions and the timescales for relaxation. In
440     eq. \eqref{C(t)1}, $P_{2}$ is the second-order Legendre polynomial,
441     the vector $\mathbf{u}$ is usually taken as HOH bisector, although
442     slightly different behavior is observed when $\mathbf{u}$ is the
443     vector along one of the OH bonds. The angle brackets denote an
444     ensemble average over all water molecules in a given spatial region.
445 gezelter 3897
446 gezelter 3946 To investigate the dynamic behavior of water at the ice interfaces, we
447     have computed $C_{2}(z,t)$ for molecules that are present within a
448     particular slab along the $z$- axis at the initial time. The change
449     in the decay behavior as a function of the $z$ coordinate is another
450     measure of the change of how the environment changes as we traverse
451     the ice/water interface. To compute these correlation functions, each
452     of the 0.5 ns simulations was followed by a shorter 200 ps simulation
453     where the positions of every molecule in the system were recorded
454     every 0.1 ps. The systems were then divided into 30 bins and $C_2(t)$
455     was evaluated for each bin.
456 plouden 3919
457 gezelter 3946 In simulations at biological interfaces, it has been shown that
458     $C_2(t)$ for water can be fit by a triexponential decay\cite{Furse08},
459     where the three components of the decay correspond to a fast (<200 fs)
460     reorientational piece driven by the restoring forces of existing
461     hydrogen bonds, a middle (on the order of several ps) piece describing
462     the large angle jumps that occur during the breaking and formation of
463     new hydrogen bonds,\cite{Laage08,Laage11} and a slow (on the order of
464     tens of ps) contribution describing the translational motion of the
465     molecules. We have similarly fit our correlation functions
466 gezelter 3897 \begin{equation}
467 gezelter 3946 C_{2}(t) \approx a e^{-t/\tau_\mathrm{short}} + b e^{-t/\tau_\mathrm{middle}} + c
468     e^{-t/\tau_\mathrm{long}} + (1-a-b-c)
469 gezelter 3897 \end{equation}
470 gezelter 3946 An average value and standard deviation for each $\tau$ was obtained
471     for each bin from the four runs. To improve statistics, the data is
472     shown as a function of distance from the center of the ice slab in
473     figures \ref{fig:basal_Tau_comic_strip} and
474     \ref{fig:prismatic_Tau_comic_strip}.
475 gezelter 3897
476 plouden 3941 \begin{figure}
477 gezelter 3946 \includegraphics[width=\linewidth]{basal_Tau_comic_strip.pdf}
478 plouden 3941 \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. }
479     \end{figure}
480 gezelter 3897
481 gezelter 3914 \begin{figure}
482 gezelter 3946 \includegraphics[width=\linewidth]{prismatic_Tau_comic_strip.pdf}
483 plouden 3941 \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.}
484 gezelter 3914 \end{figure}
485 plouden 3904
486 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.
487 plouden 3937
488 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.
489 plouden 3937
490 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.
491 plouden 3937
492 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.
493 plouden 3937
494 plouden 3941 \subsection{Coefficient of Friction of the Interface}
495 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}
496     \begin{equation}\label{Pit}
497     \lambda=\eta/\delta
498     \end{equation}
499     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}
500     \begin{equation}\label{Kuang}
501     j_{z}(p_{x})=-\eta\frac{\partial v_{x}}{\partial z}
502     \end{equation}
503     Solving eq. \eqref{Kuang} for $\eta$ and substituting the result into eq. \eqref{Pit}, we obtain an alternate expression for the coefficient of friction.
504 plouden 3941 \begin{equation}
505 plouden 3942 \lambda=-\frac{j_{z}(p_{x})}{\delta \frac{\partial v_{x}}{\partial z}}
506 plouden 3941 \end{equation}
507 plouden 3937
508 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$0.007) amu fs\textsuperscript{-1}. It is known that the basal and prismatic faces have different surface structures. The basal face is smoother than the prismatic with small alternating valleys and crests, while the prismatic surface has deep corrugating channels. We believe the reason that the prismatic face's coefficient of friction was found to be smaller than the basal's is due to the direction of the shear. The shear of the ice/water was in the same direction of the corrugating channels, allowing water molecules to pass through the channels during the shear.
509 plouden 3939
510 plouden 3942 \begin{figure}
511 gezelter 3946 \includegraphics[width=\linewidth]{delta_example.pdf}
512 plouden 3942 \caption{\label{fig:delta_example} A schematic of determining the slip length ($\delta$). The slip length is the difference of the structural starting point of the ice and the point of intersection of the linear regressions of the liquid phase velocity profile (red) and of the solid ice velocity profile (black). The dotted line indicates the location of the ice as determined by the tetrahedrality profile.}
513     \end{figure}
514    
515    
516 gezelter 3897 \section{Conclusion}
517 gezelter 3945 Here we have simulated the basal and prismatic facets of an SPC/E model of the ice I$_\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.
518 gezelter 3897
519 gezelter 3914 \begin{acknowledgement}
520     Support for this project was provided by the National Science
521     Foundation under grant CHE-0848243. Computational time was provided
522     by the Center for Research Computing (CRC) at the University of
523     Notre Dame.
524     \end{acknowledgement}
525 gezelter 3897
526 gezelter 3914 \newpage
527     \bibstyle{achemso}
528 plouden 3909 \bibliography{iceWater}
529 gezelter 3897
530 gezelter 3914 \begin{tocentry}
531     \begin{wrapfigure}{l}{0.5\textwidth}
532     \begin{center}
533     \includegraphics[width=\linewidth]{SystemImage.png}
534     \end{center}
535     \end{wrapfigure}
536     An image of our system.
537     \end{tocentry}
538 gezelter 3897
539 gezelter 3914 \end{document}
540 gezelter 3897
541 plouden 3924 %**************************************************************
542     %Non-mass weighted slopes (amu*Angstroms^-2 * fs^-1)
543     % basal: slope=0.090677616, error in slope = 0.003691743
544     %prismatic: slope = 0.050101506, error in slope = 0.001348181
545     %Mass weighted slopes (Angstroms^-2 * fs^-1)
546     %basal slope = 4.76598E-06, error in slope = 1.94037E-07
547     %prismatic slope = 3.23131E-06, error in slope = 8.69514E-08
548     %**************************************************************