ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/iceWater/iceWater.tex
Revision: 3957
Committed: Tue Sep 10 14:31:32 2013 UTC (11 years ago) by gezelter
Content type: application/x-tex
File size: 36824 byte(s)
Log Message:
Nearly final version.

File Contents

# User Rev Content
1 gezelter 3957 \documentclass[11pt]{article}
2     \usepackage{amsmath}
3     \usepackage{amssymb}
4     \usepackage{setspace}
5     %\usepackage{endfloat}
6     \usepackage{caption}
7     %\usepackage{epsf}
8     %\usepackage{tabularx}
9     \usepackage{graphicx}
10 gezelter 3914 \usepackage{multirow}
11     \usepackage{wrapfig}
12 gezelter 3957 %\usepackage{booktabs}
13     %\usepackage{bibentry}
14     %\usepackage{mathrsfs}
15     %\usepackage[ref]{overcite}
16     \usepackage[square, comma, sort&compress]{natbib}
17     \usepackage{url}
18     \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
19     \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
20     9.0in \textwidth 6.5in \brokenpenalty=10000
21 gezelter 3897
22 gezelter 3957 % double space list of tables and figures
23     %\AtBeginDelayedFloats{\renewcommand{\baselinestretch}{1.66}}
24     \setlength{\abovecaptionskip}{20 pt}
25     \setlength{\belowcaptionskip}{30 pt}
26    
27     %\renewcommand\citemid{\ } % no comma in optional referenc note
28     \bibpunct{}{}{,}{s}{}{;}
29     \bibliographystyle{aip}
30    
31    
32     % \documentclass[journal = jpccck, manuscript = article]{achemso}
33     % \setkeys{acs}{usetitle = true}
34     % \usepackage{achemso}
35     % \usepackage{natbib}
36     % \usepackage{multirow}
37     % \usepackage{wrapfig}
38     % \usepackage{fixltx2e}
39    
40 gezelter 3914 \usepackage[version=3]{mhchem} % this is a great package for formatting chemical reactions
41     \usepackage{url}
42 gezelter 3897
43 gezelter 3957
44     \begin{document}
45    
46 gezelter 3949 \title{Simulations of solid-liquid friction at ice-I$_\mathrm{h}$ /
47     water interfaces}
48 gezelter 3897
49 gezelter 3957 \author{Patrick B. Louden and J. Daniel
50     Gezelter\footnote{Corresponding author. \ Electronic mail:
51     gezelter@nd.edu} \\
52     Department of Chemistry and Biochemistry,\\
53     University of Notre Dame\\
54     Notre Dame, Indiana 46556}
55 gezelter 3897
56 gezelter 3957 \date{\today}
57     \maketitle
58     \begin{doublespace}
59 gezelter 3897
60    
61 gezelter 3914 \begin{abstract}
62 gezelter 3945 We have investigated the structural and dynamic properties of the
63 gezelter 3954 basal and prismatic facets of the ice I$_\mathrm{h}$ / water
64     interface when the solid phase is drawn through the liquid
65 gezelter 3945 (i.e. sheared relative to the fluid phase). To impose the shear, we
66     utilized a velocity-shearing and scaling (VSS) approach to reverse
67     non-equilibrium molecular dynamics (RNEMD). This method can create
68     simultaneous temperature and velocity gradients and allow the
69     measurement of transport properties at interfaces. The interfacial
70 gezelter 3954 width was found to be independent of the relative velocity of the
71     ice and liquid layers over a wide range of shear rates. Decays of
72     molecular orientational time correlation functions gave similar
73     estimates for the width of the interfaces, although the short- and
74     longer-time decay components behave differently closer to the
75     interface. Although both facets of ice are in ``stick'' boundary
76     conditions in liquid water, the solid-liquid friction coefficients
77     were found to be significantly different for the basal and prismatic
78     facets of ice.
79 gezelter 3914 \end{abstract}
80 gezelter 3897
81 gezelter 3914 \newpage
82 gezelter 3897
83     \section{Introduction}
84 plouden 3919 %-----Outline of Intro---------------
85     % in general, ice/water interface is important b/c ....
86     % here are some people who have worked on ice/water, trying to understand the processes above ....
87     % with the recent development of VSS-RNEMD, we can now look at the shearing problem
88     % talk about what we will present in this paper
89     % -------End Intro------------------
90 gezelter 3897
91 plouden 3919 %Gay02: cites many other ice/water papers, make sure to cite them.
92    
93 gezelter 3945 Understanding the ice/water interface is essential for explaining
94     complex processes such as nucleation and crystal
95     growth,\cite{Han92,Granasy95,Vanfleet95} crystal
96     melting,\cite{Weber83,Han92,Sakai96,Sakai96B} and some fascinating
97     biological processes, such as the behavior of the antifreeze proteins
98     found in winter flounder,\cite{Wierzbicki07, Chapsky97} and certain
99     terrestrial arthropods.\cite{Duman:2001qy,Meister29012013} There has
100     been significant progress on understanding the structure and dynamics
101     of quiescent ice/water interfaces utilizing both theory and
102     experiment. Haymet \emph{et al.} have done extensive work on ice I$_\mathrm{h}$,
103     including characterizing and determining the width of the ice/water
104 gezelter 3957 interface for the SPC,\cite{Karim90} SPC/E,\cite{Gay02,Bryk02} CF1,\cite{Hayward01,Hayward02} and TIP4P~\cite{Karim88} models for
105 plouden 3950 water.
106 gezelter 3945 More recently, Haymet \emph{et al.} have investigated the effects
107     cations and anions have on crystal
108 gezelter 3957 nucleation.\cite{Bryk04,Smith05,Wilson08,Wilson10} Nada \emph{et al.}
109 gezelter 3945 have also studied ice/water
110     interfaces,\cite{Nada95,Nada00,Nada03,Nada12} and have found that the
111     differential growth rates of the facets of ice I$_\mathrm{h}$ are due to the the
112 gezelter 3957 reordering of the hydrogen bonding network.\cite{Nada05}
113 plouden 3919
114 gezelter 3945 The movement of liquid water over the facets of ice has been less
115     thoroughly studied than the quiescent surfaces. This process is
116     potentially important in understanding transport of large blocks of
117     ice in water (which has important implications in the earth sciences),
118     as well as the relative motion of crystal-crystal interfaces that have
119     been separated by nanometer-scale fluid domains. In addition to
120     understanding both the structure and thickness of the interfacial
121     regions, it is important to understand the molecular origin of
122     friction, drag, and other changes in dynamical properties of the
123     liquid in the regions close to the surface that are altered by the
124     presence of a shearing of the bulk fluid relative to the solid phase.
125 plouden 3919
126 gezelter 3945 In this work, we apply a recently-developed velocity shearing and
127     scaling approach to reverse non-equilibrium molecular dynamics
128     (VSS-RNEMD). This method makes it possible to calculate transport
129     properties like the interfacial thermal conductance across
130     heterogeneous interfaces,\cite{Kuang12} and can create simultaneous
131     temperature and velocity gradients and allow the measurement of
132     friction and thermal transport properties at interfaces. This has
133     allowed us to investigate the width of the ice/water interface as the
134     ice is sheared through the liquid, while simultaneously imposing a
135     weak thermal gradient to prevent frictional heating of the interface.
136     In the sections that follow, we discuss the methodology for creating
137     and simulating ice/water interfaces under shear and provide results
138     from both structural and dynamical correlation functions. We also
139     show that the solid-liquid interfacial friction coefficient depends
140     sensitively on the details of the surface morphology.
141    
142 gezelter 3897 \section{Methodology}
143    
144 gezelter 3945 \subsection{Stable ice I$_\mathrm{h}$ / water interfaces under shear}
145 gezelter 3914
146 gezelter 3946 The structure of ice I$_\mathrm{h}$ is well understood; it
147 gezelter 3914 crystallizes in a hexagonal space group P$6_3/mmc$, and the hexagonal
148     crystals of ice have two faces that are commonly exposed, the basal
149     face $\{0~0~0~1\}$, which forms the top and bottom of each hexagonal
150     plate, and the prismatic $\{1~0~\bar{1}~0\}$ face which forms the
151     sides of the plate. Other less-common, but still important, faces of
152 gezelter 3945 ice I$_\mathrm{h}$ are the secondary prism, $\{1~1~\bar{2}~0\}$, and
153 gezelter 3946 pyramidal, $\{2~0~\bar{2}~1\}$, faces. Ice I$_\mathrm{h}$ is normally
154     proton disordered in bulk crystals, although the surfaces probably
155     have a preference for proton ordering along strips of dangling H-atoms
156     and Oxygen lone pairs.\cite{Buch:2008fk}
157 gezelter 3914
158 gezelter 3957 \begin{table}[h]
159     \centering
160 gezelter 3945 \caption{Mapping between the Miller indices of four facets of ice in
161     the $P6_3/mmc$ crystal system to the orthorhombic $P2_12_12_1$
162 gezelter 3957 system in reference \bibpunct{}{}{,}{n}{}{,} \protect\citep{Hirsch04}.}
163 gezelter 3945 \label{tab:equiv}
164 gezelter 3914 \begin{tabular}{|ccc|} \hline
165     & hexagonal & orthorhombic \\
166     & ($P6_3/mmc$) & ($P2_12_12_1$) \\
167     crystal face & Miller indices & equivalent \\ \hline
168     basal & $\{0~0~0~1\}$ & $\{0~0~1\}$ \\
169     prism & $\{1~0~\bar{1}~0\}$ & $\{1~0~0\}$ \\
170     secondary prism & $\{1~1~\bar{2}~0\}$ & $\{1~3~0\}$ \\
171 gezelter 3946 pyramidal & $\{2~0~\bar{2}~1\}$ & $\{2~0~1\}$ \\ \hline
172 gezelter 3914 \end{tabular}
173 gezelter 3957 \end{table}
174 gezelter 3946 For small simulated ice interfaces, it is useful to work with
175     proton-ordered, but zero-dipole crystal that exposes these strips of
176     dangling H-atoms and lone pairs. When placing another material in
177     contact with one of the ice crystalline planes, it is also quite
178     useful to have an orthorhombic (rectangular) box. Recent work by
179     Hirsch and Ojam\"{a}e describes a number of alternative crystal
180     systems for proton-ordered bulk ice I$_\mathrm{h}$ using orthorhombic
181     cells.\cite{Hirsch04}
182    
183     In this work, we are using Hirsch and Ojam\"{a}e's structure 6 which
184     is an orthorhombic cell ($P2_12_12_1$) that produces a proton-ordered
185     version of ice I$_\mathrm{h}$. Table \ref{tab:equiv} contains a
186     mapping between the Miller indices of common ice facets in the
187     P$6_3/mmc$ crystal system and those in the Hirsch and Ojam\"{a}e
188     $P2_12_12_1$ system.
189    
190 gezelter 3914 Structure 6 from the Hirsch and Ojam\"{a}e paper has lattice
191 gezelter 3957 parameters $a = 4.49225$ \AA\ , $b = 7.78080$ \AA\ , $c = 7.33581$ \AA\
192     and two water molecules whose atoms reside at fractional coordinates
193     given in table \ref{tab:p212121}. To construct the basal and prismatic
194     interfaces, these crystallographic coordinates were used to construct
195     an orthorhombic unit cell which was then replicated in all three
196 gezelter 3945 dimensions yielding a proton-ordered block of ice I$_{h}$. To expose
197     the desired face, the orthorhombic representation was then cut along
198     the ($001$) or ($100$) planes for the basal and prismatic faces
199 gezelter 3957 respectively. The resulting block was rotated so that the exposed
200     faces were aligned with the $z$-axis normal to the exposed face. The
201 gezelter 3946 block was then cut along two perpendicular directions in a way that
202     allowed for perfect periodic replication in the $x$ and $y$ axes,
203 gezelter 3945 creating a slab with either the basal or prismatic faces exposed along
204 gezelter 3957 the $z$ axis. The slab was then replicated in the $x$ and $y$
205 gezelter 3945 dimensions until a desired sample size was obtained.
206 gezelter 3914
207 gezelter 3957 \begin{table}[h]
208     \centering
209 gezelter 3945 \caption{Fractional coordinates for water in the orthorhombic
210 gezelter 3957 $P2_12_12_1$ system for ice I$_\mathrm{h}$ in reference \bibpunct{}{}{,}{n}{}{,} \protect\citep{Hirsch04}.}
211 gezelter 3945 \label{tab:p212121}
212     \begin{tabular}{|cccc|} \hline
213     atom type & x & y & z \\ \hline
214 gezelter 3957 O & 0.7500 & 0.1667 & 0.4375 \\
215 gezelter 3945 H & 0.5735 & 0.2202 & 0.4836 \\
216     H & 0.7420 & 0.0517 & 0.4836 \\
217 gezelter 3957 O & 0.2500 & 0.6667 & 0.4375 \\
218 gezelter 3945 H & 0.2580 & 0.6693 & 0.3071 \\
219     H & 0.4265 & 0.7255 & 0.4756 \\ \hline
220 gezelter 3914 \end{tabular}
221 gezelter 3957 \end{table}
222 gezelter 3914
223 gezelter 3946 Our ice / water interfaces were created using a box of liquid water
224     that had the same dimensions (in $x$ and $y$) as the ice block.
225     Although the experimental solid/liquid coexistence temperature under
226 gezelter 3957 atmospheric pressure is close to 273~K, Haymet \emph{et al.} have done
227 gezelter 3946 extensive work on characterizing the ice/water interface, and find
228     that the coexistence temperature for simulated water is often quite a
229     bit different.\cite{Karim88,Karim90,Hayward01,Bryk02,Hayward02} They
230     have found that for the SPC/E water model,\cite{Berendsen87} which is
231     also used in this study, the ice/water interface is most stable at
232 gezelter 3957 225~$\pm$5~K.\cite{Bryk02} This liquid box was therefore equilibrated at
233     225~K and 1~atm of pressure in the NPAT ensemble (with the $z$ axis
234 gezelter 3946 allowed to fluctuate to equilibrate to the correct pressure). The
235     liquid and solid systems were combined by carving out any water
236 gezelter 3957 molecule from the liquid simulation cell that was within 3~\AA\ of any
237 gezelter 3946 atom in the ice slab.
238 gezelter 3914
239     Molecular translation and orientational restraints were applied in the
240     early stages of equilibration to prevent melting of the ice slab.
241     These restraints were removed during NVT equilibration, well before
242 gezelter 3945 data collection was carried out.
243 gezelter 3914
244 gezelter 3918 \subsection{Shearing ice / water interfaces without bulk melting}
245    
246 gezelter 3945 As a solid is dragged through a liquid, there is frictional heating
247     that will act to melt the interface. To study the behavior of the
248     interface under a shear stress without causing the interface to melt,
249 gezelter 3946 it is necessary to apply a weak thermal gradient in combination with
250     the momentum gradient. This can be accomplished using the velocity
251 gezelter 3945 shearing and scaling (VSS) variant of reverse non-equilibrium
252     molecular dynamics (RNEMD), which utilizes a series of simultaneous
253     velocity exchanges between two regions within the simulation
254     cell.\cite{Kuang12} One of these regions is centered within the ice
255 gezelter 3946 slab, while the other is centrally located in the liquid
256 gezelter 3918 region. VSS-RNEMD provides a set of conservation constraints for
257 gezelter 3946 creating either a momentum flux or a thermal flux (or both
258     simultaneously) between the two slabs. Satisfying the constraint
259     equations ensures that the new configurations are sampled from the
260     same NVE ensemble as before the VSS move.
261 gezelter 3918
262     The VSS moves are applied periodically to scale and shift the particle
263     velocities ($\mathbf{v}_i$ and $\mathbf{v}_j$) in two slabs ($H$ and
264     $C$) which are separated by half of the simulation box,
265     \begin{displaymath}
266     \begin{array}{rclcl}
267    
268     & \underline{\mathrm{shearing}} & &
269     \underline{~~~~~~~~~~~~\mathrm{scaling}~~~~~~~~~~~~} \\
270     \mathbf{v}_i \leftarrow &
271     \mathbf{a}_c\ & + & c\cdot\left(\mathbf{v}_i - \langle\mathbf{v}_c
272     \rangle\right) + \langle\mathbf{v}_c\rangle \\
273     \mathbf{v}_j \leftarrow &
274     \mathbf{a}_h & + & h\cdot\left(\mathbf{v}_j - \langle\mathbf{v}_h
275     \rangle\right) + \langle\mathbf{v}_h\rangle .
276    
277     \end{array}
278     \end{displaymath}
279     Here $\langle\mathbf{v}_c\rangle$ and $\langle\mathbf{v}_h\rangle$ are
280     the center of mass velocities in the $C$ and $H$ slabs, respectively.
281     Within the two slabs, particles receive incremental changes or a
282     ``shear'' to their velocities. The amount of shear is governed by the
283     imposed momentum flux, $\mathbf{j}_z(\mathbf{p})$
284     \begin{eqnarray}
285     \mathbf{a}_c & = & - \mathbf{j}_z(\mathbf{p}) \Delta t / M_c \label{vss1}\\
286     \mathbf{a}_h & = & + \mathbf{j}_z(\mathbf{p}) \Delta t / M_h \label{vss2}
287     \end{eqnarray}
288     where $M_{\{c,h\}}$ is the total mass of particles within each of the
289     slabs and $\Delta t$ is the interval between two separate operations.
290    
291     To simultaneously impose a thermal flux ($J_z$) between the slabs we
292     use energy conservation constraints,
293     \begin{eqnarray}
294     K_c - J_z\Delta t & = & c^2 (K_c - \frac{1}{2}M_c \langle\mathbf{v}_c
295     \rangle^2) + \frac{1}{2}M_c (\langle \mathbf{v}_c \rangle + \mathbf{a}_c)^2 \label{vss3}\\
296     K_h + J_z\Delta t & = & h^2 (K_h - \frac{1}{2}M_h \langle\mathbf{v}_h
297     \rangle^2) + \frac{1}{2}M_h (\langle \mathbf{v}_h \rangle +
298     \mathbf{a}_h)^2 \label{vss4}.
299     \label{constraint}
300     \end{eqnarray}
301     Simultaneous solution of these quadratic formulae for the scaling
302     coefficients, $c$ and $h$, will ensure that the simulation samples from
303     the original microcanonical (NVE) ensemble. Here $K_{\{c,h\}}$ is the
304     instantaneous translational kinetic energy of each slab. At each time
305     interval, it is a simple matter to solve for $c$, $h$, $\mathbf{a}_c$,
306     and $\mathbf{a}_h$, subject to the imposed momentum flux,
307     $j_z(\mathbf{p})$, and thermal flux, $J_z$, values. Since the VSS
308     operations do not change the kinetic energy due to orientational
309     degrees of freedom or the potential energy of a system, configurations
310     after the VSS move have exactly the same energy (and linear
311     momentum) as before the move.
312    
313     As the simulation progresses, the VSS moves are performed on a regular
314     basis, and the system develops a thermal and/or velocity gradient in
315 gezelter 3945 response to the applied flux. In a bulk material, it is quite simple
316 gezelter 3918 to use the slope of the temperature or velocity gradients to obtain
317 gezelter 3945 either the thermal conductivity or shear viscosity.
318 gezelter 3918
319     The VSS-RNEMD approach is versatile in that it may be used to
320     implement thermal and shear transport simultaneously. Perturbations
321     of velocities away from the ideal Maxwell-Boltzmann distributions are
322     minimal, as is thermal anisotropy. This ability to generate
323     simultaneous thermal and shear fluxes has been previously utilized to
324     map out the shear viscosity of SPC/E water over a wide range of
325 gezelter 3957 temperatures (90~K) with a single 1~ns simulation.\cite{Kuang12}
326 gezelter 3918
327 gezelter 3945 For this work, we are using the VSS-RNEMD method primarily to generate
328     a shear between the ice slab and the liquid phase, while using a weak
329 gezelter 3957 thermal gradient to maintain the interface at the 225~K target
330 gezelter 3946 value. This ensures minimal melting of the bulk ice phase and allows
331     us to control the exact temperature of the interface.
332 gezelter 3918
333 gezelter 3897 \subsection{Computational Details}
334 gezelter 3945 All simulations were performed using OpenMD,\cite{OOPSE,openmd} with a
335     time step of 2 fs and periodic boundary conditions in all three
336     dimensions. Electrostatics were handled using the damped-shifted
337     force real-space electrostatic kernel.\cite{Ewald} The systems were
338     divided into 100 bins along the $z$-axis for the VSS-RNEMD moves,
339 gezelter 3957 which were attempted every 50~fs.
340 gezelter 3897
341 gezelter 3945 The interfaces were equilibrated for a total of 10 ns at equilibrium
342     conditions before being exposed to either a shear or thermal gradient.
343     This consisted of 5 ns under a constant temperature (NVT) integrator
344     set to 225K followed by 5 ns under a microcanonical integrator. Weak
345     thermal gradients were allowed to develop using the VSS-RNEMD (NVE)
346     integrator using a a small thermal flux ($-2.0\times 10^{-6}$
347     kcal/mol/\AA$^2$/fs) for a duration of 5 ns to allow the gradient to
348 gezelter 3946 stabilize. The resulting temperature gradient was $\approx$ 10K over
349     the entire 100 \AA\ box length, which was sufficient to keep the
350 gezelter 3945 temperature at the interface within $\pm 1$ K of the 225K target.
351 gezelter 3897
352 gezelter 3945 Velocity gradients were then imposed using the VSS-RNEMD (NVE)
353     integrator with a range of momentum fluxes. These gradients were
354 gezelter 3957 allowed to stabilize for 1~ns before data collection began. Once
355     established, four successive 0.5~ns runs were performed for each shear
356 gezelter 3945 rate. During these simulations, snapshots of the system were taken
357 gezelter 3957 every 1~ps, and statistics on the structure and dynamics in each bin
358 gezelter 3945 were accumulated throughout the simulations.
359    
360 plouden 3941 \section{Results and discussion}
361    
362 gezelter 3946 \subsection{Interfacial width}
363     Any order parameter or time correlation function that changes as one
364     crosses an interface from a bulk liquid to a solid can be used to
365     measure the width of the interface. In previous work on the ice/water
366 gezelter 3954 interface, Haymet {\it et al.}\cite{Bryk02} have utilized structural
367 gezelter 3946 features (including the density) as well as dynamic properties
368     (including the diffusion constant) to estimate the width of the
369     interfaces for a number of facets of the ice crystals. Because
370     VSS-RNEMD imposes a lateral flow, parameters that depend on
371     translational motion of the molecules (e.g. diffusion) may be
372 gezelter 3954 artificially skewed by the RNEMD moves. A structural parameter is not
373 gezelter 3946 influenced by the RNEMD perturbations to the same degree. Here, we
374 gezelter 3954 have used the local tetrahedral order parameter as described by
375 gezelter 3946 Kumar\cite{Kumar09} and Errington\cite{Errington01} as our principal
376 gezelter 3945 measure of the interfacial width.
377 plouden 3936
378 gezelter 3945 The local tetrahedral order parameter, $q(z)$, is given by
379 gezelter 3897 \begin{equation}
380 gezelter 3946 q(z) = \int_0^L \sum_{k=1}^{N} \Bigg(1 -\frac{3}{8}\sum_{i=1}^{3}
381     \sum_{j=i+1}^{4} \bigg(\cos\psi_{ikj}+\frac{1}{3}\bigg)^2\Bigg)
382     \delta(z_{k}-z)\mathrm{d}z \Bigg/ N_z
383 gezelter 3945 \label{eq:qz}
384 gezelter 3897 \end{equation}
385 gezelter 3945 where $\psi_{ikj}$ is the angle formed between the oxygen site on
386 gezelter 3946 central molecule $k$, and the oxygen sites on two of the four closest
387     molecules, $i$ and $j$. Molecules $i$ and $j$ are further restricted
388     to lie within the first solvation shell of molecule $k$. $N_z = \int
389     \delta(z_k - z) \mathrm{d}z$ is a normalization factor to account for
390     the varying population of molecules within each finite-width bin. The
391     local tetrahedral order parameter has a range of $(0,1)$, where the
392     larger values of $q$ indicate a larger degree of tetrahedral ordering
393     of the local environment. In perfect ice I$_\mathrm{h}$ structures,
394     the parameter can approach 1 at low temperatures, while in liquid
395     water, the ordering is significantly less tetrahedral, and values of
396     $q(z) \approx 0.75$ are more common.
397 plouden 3909
398 gezelter 3946 To estimate the interfacial width, the system was divided into 100
399     bins along the $z$-dimension, and a cutoff radius for the first
400 gezelter 3957 solvation shell was set to 3.41~\AA\ . The $q_{z}$ function was
401 gezelter 3946 time-averaged to give yield a tetrahedrality profile of the
402     system. The profile was then fit to a hyperbolic tangent that smoothly
403     links the liquid and solid states,
404 plouden 3941 \begin{equation}\label{tet_fit}
405 gezelter 3946 q(z) \approx
406     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-
407     \frac{r+l}{2}\right|.
408 gezelter 3897 \end{equation}
409 gezelter 3946 Here $q_{liq}$ and $q_{ice}$ are the local tetrahedral order parameter
410     for the bulk liquid and ice domains, respectively, $w$ is the width of
411     the interface. $l$ and $r$ are the midpoints of the left and right
412 gezelter 3957 interfaces, respectively. The last term in eq. \eqref{tet_fit}
413 gezelter 3946 accounts for the influence that the weak thermal gradient has on the
414     tetrahedrality profile in the liquid region. To estimate the
415 gezelter 3954 10\%-90\% widths commonly used in previous studies,\cite{Bryk02} it is
416     a simple matter to scale the widths obtained from the hyperbolic
417     tangent fits to obtain $w_{10-90} = 2.1971 \times w$.\cite{Bryk02}
418 gezelter 3897
419 gezelter 3946 In Figures \ref{fig:bComic} and \ref{fig:pComic} we see the
420     $z$-coordinate profiles for tetrahedrality, temperature, and the
421     $x$-component of the velocity for the basal and prismatic interfaces.
422     The lower panels show the $q(z)$ (black circles) along with the
423     hyperbolic tangent fits (red lines). In the liquid region, the local
424     tetrahedral order parameter, $q(z) \approx 0.75$ while in the
425     crystalline region, $q(z) \approx 0.94$, indicating a more tetrahedral
426     environment. The vertical dotted lines denote the midpoint of the
427 gezelter 3957 interfaces ($r$ and $l$ in eq. \eqref{tet_fit}). The weak thermal
428 gezelter 3946 gradient applied to the systems in order to keep the interface at
429 gezelter 3957 225~$\pm$~5K, can be seen in middle panels. The transverse velocity
430 gezelter 3946 profile is shown in the upper panels. It is clear from the upper
431     panels that water molecules in close proximity to the surface (i.e.
432 gezelter 3957 within 10~\AA\ to 15~\AA~of the interfaces) have transverse
433 gezelter 3946 velocities quite close to the velocities within the ice block. There
434     is no velocity discontinuity at the interface, which indicates that
435     the shearing of ice/water interfaces occurs in the ``stick'' or
436     no-slip boundary conditions.
437 gezelter 3897
438 gezelter 3914 \begin{figure}
439 gezelter 3957 \includegraphics[width=\linewidth]{bComicStrip}
440 gezelter 3946 \caption{\label{fig:bComic} The basal interfaces. Lower panel: the
441     local tetrahedral order parameter, $q(z)$, (black circles) and the
442     hyperbolic tangent fit (red line). Middle panel: the imposed
443     thermal gradient required to maintain a fixed interfacial
444     temperature. Upper panel: the transverse velocity gradient that
445     develops in response to an imposed momentum flux. The vertical
446     dotted lines indicate the locations of the midpoints of the two
447     interfaces.}
448 gezelter 3914 \end{figure}
449 plouden 3904
450 gezelter 3914 \begin{figure}
451 gezelter 3957 \includegraphics[width=\linewidth]{pComicStrip}
452 gezelter 3946 \caption{\label{fig:pComic} The prismatic interfaces. Panel
453     descriptions match those in figure \ref{fig:bComic}}
454 gezelter 3914 \end{figure}
455    
456 gezelter 3957 From the fits using eq. \eqref{tet_fit}, we find the interfacial
457     width for the basal and prismatic systems to be 3.2~$\pm$~0.4~\AA\ and
458     3.6~$\pm$~0.2~\AA\ , respectively, with no applied momentum flux. Over
459 gezelter 3946 the range of shear rates investigated, $0.6 \pm 0.3 \mathrm{ms}^{-1}
460     \rightarrow 5.3 \pm 0.5 \mathrm{ms}^{-1}$ for the basal system and
461     $0.9 \pm 0.2 \mathrm{ms}^{-1} \rightarrow 4.5 \pm 0.1
462     \mathrm{ms}^{-1}$ for the prismatic, we found no appreciable change in
463     the interface width. The fit values for the interfacial width ($w$)
464     over all shear rates contained the values reported above within their
465     error bars.
466 gezelter 3914
467 gezelter 3954 \subsubsection{Orientational Dynamics}
468 gezelter 3946 The orientational time correlation function,
469 plouden 3941 \begin{equation}\label{C(t)1}
470 gezelter 3946 C_{2}(t)=\langle P_{2}(\mathbf{u}(0) \cdot \mathbf{u}(t)) \rangle,
471 plouden 3941 \end{equation}
472 gezelter 3946 gives insight into the local dynamic environment around the water
473     molecules. The rate at which the function decays provides information
474     about hindered motions and the timescales for relaxation. In
475     eq. \eqref{C(t)1}, $P_{2}$ is the second-order Legendre polynomial,
476 gezelter 3948 the vector $\mathbf{u}$ is often taken as HOH bisector, although
477     slightly different behavior can be observed when $\mathbf{u}$ is the
478 gezelter 3946 vector along one of the OH bonds. The angle brackets denote an
479     ensemble average over all water molecules in a given spatial region.
480 gezelter 3897
481 gezelter 3946 To investigate the dynamic behavior of water at the ice interfaces, we
482     have computed $C_{2}(z,t)$ for molecules that are present within a
483     particular slab along the $z$- axis at the initial time. The change
484     in the decay behavior as a function of the $z$ coordinate is another
485 gezelter 3948 measure of the change of how the local environment changes across the
486     ice/water interface. To compute these correlation functions, each of
487     the 0.5 ns simulations was followed by a shorter 200 ps microcanonical
488     (NVE) simulation in which the positions and orientations of every
489     molecule in the system were recorded every 0.1 ps. The systems were
490 gezelter 3954 then divided into 30 bins along the $z$-axis and $C_2(t)$ was
491     evaluated for each bin.
492 plouden 3919
493 gezelter 3948 In simulations of water at biological interfaces, Furse {\em et al.}
494     fit $C_2(t)$ functions for water with triexponential
495     functions,\cite{Furse08} where the three components of the decay
496 gezelter 3957 correspond to a fast ($<$200 fs) reorientational piece driven by the
497 gezelter 3948 restoring forces of existing hydrogen bonds, a middle (on the order of
498     several ps) piece describing the large angle jumps that occur during
499     the breaking and formation of new hydrogen bonds,and a slow (on the
500     order of tens of ps) contribution describing the translational motion
501     of the molecules. The model for orientational decay presented
502     recently by Laage and Hynes {\em et al.}\cite{Laage08,Laage11} also
503     includes three similar decay constants, although two of the time
504     constants are linked, and the resulting decay curve has two parameters
505     governing the dynamics of decay.
506    
507     In our ice/water interfaces, we are at substantially lower
508     temperatures, and the water molecules are further perturbed by the
509     presence of the ice phase nearby. We have obtained the most
510     reasonable fits using triexponential functions with three distinct
511 gezelter 3954 time domains, as well as a constant piece to account for the water
512 gezelter 3948 stuck in the ice phase that does not experience any long-time
513     orientational decay,
514 gezelter 3897 \begin{equation}
515 gezelter 3946 C_{2}(t) \approx a e^{-t/\tau_\mathrm{short}} + b e^{-t/\tau_\mathrm{middle}} + c
516     e^{-t/\tau_\mathrm{long}} + (1-a-b-c)
517 gezelter 3897 \end{equation}
518 gezelter 3948 Average values for the three decay constants (and error estimates)
519     were obtained for each bin. In figures \ref{fig:basal_Tau_comic_strip}
520     and \ref{fig:prismatic_Tau_comic_strip}, the three orientational decay
521     times are shown as a function of distance from the center of the ice
522     slab.
523 gezelter 3897
524 plouden 3941 \begin{figure}
525 gezelter 3957 \includegraphics[width=\linewidth]{basal_Tau_comic_strip}
526 gezelter 3949 \caption{\label{fig:basal_Tau_comic_strip} The three decay constants
527     of the orientational time correlation function, $C_2(t)$, for water
528     as a function of distance from the center of the ice slab. The
529     dashed line indicates the location of the basal face (as determined
530     from the tetrahedrality order parameter). The moderate and long
531     time contributions slow down close to the interface which would be
532 gezelter 3954 expected under reorganizations that involve large motions of the
533 gezelter 3949 molecules (e.g. frame-reorientations and jumps). The observed
534     speed-up in the short time contribution is surprising, but appears
535     to reflect the restricted motion of librations closer to the
536     interface.}
537 plouden 3941 \end{figure}
538 gezelter 3897
539 gezelter 3914 \begin{figure}
540 gezelter 3957 \includegraphics[width=\linewidth]{prismatic_Tau_comic_strip}
541 gezelter 3949 \caption{\label{fig:prismatic_Tau_comic_strip}
542     Decay constants for $C_2(t)$ at the prismatic interface. Panel
543     descriptions match those in figure \ref{fig:basal_Tau_comic_strip}.}
544 gezelter 3914 \end{figure}
545 plouden 3904
546 gezelter 3949 Figures \ref{fig:basal_Tau_comic_strip} and
547     \ref{fig:prismatic_Tau_comic_strip} show the three decay constants for
548     the orientational time correlation function for water at varying
549     displacements from the center of the ice slab for both the basal and
550     prismatic interfaces. The vertical dotted lines indicate the
551     locations of the midpoints of the interfaces as determined by the
552     tetrahedrality fits. In the liquid regions, $\tau_{middle}$ and
553     $\tau_{long}$ have consistent values around 3-4 ps and 20-40 ps,
554     respectively, and increase in value approaching the interface.
555     According to the jump model of Laage and Hynes {\em et
556     al.},\cite{Laage08,Laage11} $\tau_{middle}$ corresponds to the
557     breaking and making of hydrogen bonds and $\tau_{long}$ is explained
558     with translational motion of the molecules (i.e. frame reorientation).
559     The shortest of the three decay constants, the librational time
560     $\tau_\mathrm{short}$ has a value of about 70 fs in the liquid region,
561     and decreases in value approaching the interface. The observed
562     speed-up in the short time contribution is surprising, but appears to
563     reflect the restricted motion of librations closer to the interface.
564 plouden 3937
565 gezelter 3949 The control systems (with no applied momentum flux) are shown with
566     black symbols in figs. \ref{fig:basal_Tau_comic_strip} and
567     \ref{fig:prismatic_Tau_comic_strip}, while those obtained while a
568     shear was active are shown in red.
569 plouden 3937
570 gezelter 3952 Two notable features deserve clarification. First, there are
571     nearly-constant liquid-state values for $\tau_{short}$,
572 gezelter 3949 $\tau_{middle}$, and $\tau_{long}$ at large displacements from the
573 gezelter 3952 interface. Second, there appears to be a single distance, $d_{basal}$
574     or $d_{prismatic}$, from the interface at which all three decay times
575     begin to deviate from their bulk liquid values. We find these
576 gezelter 3957 distances to be approximately 15~\AA\ and 8~\AA\, respectively,
577 gezelter 3952 although significantly finer binning of the $C_2(t)$ data would be
578     necessary to provide better estimates of a ``dynamic'' interfacial
579     thickness.
580 plouden 3937
581 gezelter 3952 Beaglehole and Wilson have measured the ice/water interface using
582 gezelter 3957 ellipsometry and find a thickness of approximately 10~\AA\ for both
583 gezelter 3952 the basal and prismatic faces.\cite{Beaglehole93} Structurally, we
584 gezelter 3957 have found the basal and prismatic interfacial width to be
585     3.2~$\pm$~0.4~\AA\ and 3.6~$\pm$~0.2~\AA. However, decomposition of
586     the spatial dependence of the decay times of $C_2(t)$ indicates that a
587     somewhat thicker interfacial region exists in which the orientational
588     dynamics of the water molecules begin to resemble the trapped
589     interfacial water more than the surrounding liquid.
590 plouden 3937
591 gezelter 3952 Our results indicate that the dynamics of the water molecules within
592     $d_{basal}$ and $d_{prismatic}$ are being significantly perturbed by
593     the interface, even though the structural width of the interface via
594     analysis of the tetrahedrality profile indicates that bulk liquid
595     structure of water is recovered after about 4 \AA\ from the edge of
596     the ice.
597    
598 plouden 3941 \subsection{Coefficient of Friction of the Interface}
599 gezelter 3954 As liquid water flows over an ice interface, there is a distance from
600     the structural interface where bulk-like hydrodynamics are recovered.
601     Bocquet and Barrat constructed a theory for the hydrodynamic boundary
602     parameters, which include the slipping length
603 gezelter 3952 $\left(\delta_\mathrm{wall}\right)$ of this boundary layer and the
604     ``hydrodynamic position'' of the boundary
605 gezelter 3954 $\left(z_\mathrm{wall}\right)$.\cite{PhysRevLett.70.2726,PhysRevE.49.3079}
606     This last parameter is the location (relative to a solid surface)
607 gezelter 3952 where the bulk-like behavior is recovered. Work by Mundy {\it et al.}
608 gezelter 3954 has helped to combine these parameters into a liquid-solid friction
609 gezelter 3952 coefficient, which quantifies the resistance to pulling the solid
610 gezelter 3954 interface through a liquid,\cite{Mundy1997305}
611 gezelter 3952 \begin{equation}
612     \lambda_\mathrm{wall} = \frac{\eta}{\delta_\mathrm{wall}}.
613     \end{equation}
614 gezelter 3954 This expression is nearly identical to one provided by Pit {\it et
615     al.} for the solid-liquid friction of an interface,\cite{Pit99}
616 plouden 3942 \begin{equation}\label{Pit}
617 gezelter 3952 \lambda=\frac{\eta}{\delta}
618 plouden 3942 \end{equation}
619 gezelter 3952 where $\delta$ is the slip length for the liquid measured at the
620     location of the interface itself.
621    
622     In both of these expressions, $\eta$ is the shear viscosity of the
623     bulk-like region of the liquid, a quantity which is easily obtained in
624     VSS-RNEMD simulations by fitting the velocity profile in the region
625 gezelter 3954 far from the surface.\cite{Kuang12} Assuming linear response in the
626 gezelter 3952 bulk-like region,
627 plouden 3942 \begin{equation}\label{Kuang}
628 gezelter 3952 j_{z}(p_{x})=-\eta \left(\frac{\partial v_{x}}{\partial z}\right)
629 plouden 3942 \end{equation}
630 gezelter 3952 Substituting this result into eq. \eqref{Pit}, we can estimate the
631     solid-liquid coefficient using the slip length,
632 plouden 3941 \begin{equation}
633 gezelter 3954 \lambda=-\frac{j_{z}(p_{x})} {\left(\frac{\partial v_{x}}{\partial
634     z}\right) \delta}
635 plouden 3941 \end{equation}
636 plouden 3937
637 gezelter 3954 For ice / water interfaces, the boundary conditions are markedly
638     no-slip, so projecting the bulk liquid state velocity profile yields a
639 gezelter 3957 negative slip length. This length is the difference between the
640 gezelter 3954 structural edge of the ice (determined by the tetrahedrality profile)
641     and the location where the projected velocity of the bulk liquid
642     intersects the solid phase velocity (see Figure
643     \ref{fig:delta_example}). The coefficients of friction for the basal
644 gezelter 3957 and the prismatic facets are found to be
645     0.07~$\pm$~0.01~amu~fs\textsuperscript{-1} and
646     0.032~$\pm$~0.007~amu~fs\textsuperscript{-1}, respectively. These
647     results may seem surprising as the basal face is smoother than the
648     prismatic with only small undulations of the oxygen positions, while
649     the prismatic surface has deep corrugated channels. The applied
650     momentum flux used in our simulations is parallel to these channels,
651     however, and this results in a flow of water in the same direction as
652     the corrugations, allowing water molecules to pass through the
653     channels during the shear.
654 plouden 3939
655 plouden 3942 \begin{figure}
656 gezelter 3957 \includegraphics[width=\linewidth]{delta_example}
657 gezelter 3952 \caption{\label{fig:delta_example} Determining the (negative) slip
658     length ($\delta$) for the ice-water interfaces (which have decidedly
659     non-slip behavior). This length is the difference between the
660     structural edge of the ice (determined by the tetrahedrality
661     profile) and the location where the projected velocity of the bulk
662     liquid (dashed red line) intersects the solid phase velocity (solid
663     black line). The dotted line indicates the location of the ice as
664     determined by the tetrahedrality profile.}
665 plouden 3942 \end{figure}
666    
667    
668 gezelter 3897 \section{Conclusion}
669 gezelter 3952 We have simulated the basal and prismatic facets of an SPC/E model of
670     the ice I$_\mathrm{h}$ / water interface. Using VSS-RNEMD, the ice
671     was sheared relative to the liquid while simultaneously being exposed
672     to a weak thermal gradient which kept the interface at a stable
673     temperature. Calculation of the local tetrahedrality order parameter
674     has shown an apparent independence of the interfacial width on the
675 gezelter 3957 shear rate. This width was found to be 3.2~$\pm$0.4~\AA\ and
676     3.6~$\pm$0.2~\AA\ for the basal and prismatic systems, respectively.
677 gezelter 3897
678 gezelter 3952 Orientational time correlation functions were calculated at varying
679     displacements from the interface, and were found to be similarly
680     independent of the applied momentum flux. The short decay due to the
681     restoring forces of existing hydrogen bonds decreased close to the
682     interface, while the longer-time decay constants increased in close
683     proximity to the interface. There is also an apparent dynamic
684     interface width, $d_{basal}$ and $d_{prismatic}$, at which these
685 gezelter 3954 deviations from bulk liquid values begin. We found $d_{basal}$ and
686 gezelter 3957 $d_{prismatic}$ to be approximately 15~\AA\ and 8~\AA\ . This implies
687 gezelter 3954 that the dynamics of water molecules which have similar structural
688     environments to liquid phase molecules are dynamically perturbed by
689     the presence of the ice interface.
690 gezelter 3952
691 gezelter 3954 The coefficient of liquid-solid friction for each of the facets was
692 gezelter 3957 also determined. They were found to be
693     0.07~$\pm$~0.01~amu~fs\textsuperscript{-1} and
694     0.032~$\pm$~0.007~amu~fs\textsuperscript{-1} for the basal and
695     prismatic facets respectively. We attribute the large difference
696     between the two friction coefficients to the direction of the shear
697     and to the surface structure of the crystal facets.
698 gezelter 3952
699 gezelter 3957 \section{Acknowledgements}
700 gezelter 3914 Support for this project was provided by the National Science
701     Foundation under grant CHE-0848243. Computational time was provided
702     by the Center for Research Computing (CRC) at the University of
703     Notre Dame.
704 gezelter 3897
705 gezelter 3914 \newpage
706 plouden 3909 \bibliography{iceWater}
707 gezelter 3897
708 gezelter 3957 \end{doublespace}
709 gezelter 3897
710 gezelter 3957 % \begin{tocentry}
711     % \begin{wrapfigure}{l}{0.5\textwidth}
712     % \begin{center}
713     % \includegraphics[width=\linewidth]{SystemImage.png}
714     % \end{center}
715     % \end{wrapfigure}
716     % The cell used to simulate liquid-solid shear in ice I$_\mathrm{h}$ /
717     % water interfaces. Velocity gradients were applied using the velocity
718     % shearing and scaling variant of reverse non-equilibrium molecular
719     % dynamics (VSS-RNEMD) with a weak thermal gradient to prevent melting.
720     % The interface width is relatively robust in both structual and dynamic
721     % measures as a function of the applied shear.
722     % \end{tocentry}
723    
724 gezelter 3914 \end{document}
725 gezelter 3897
726 plouden 3924 %**************************************************************
727     %Non-mass weighted slopes (amu*Angstroms^-2 * fs^-1)
728     % basal: slope=0.090677616, error in slope = 0.003691743
729     %prismatic: slope = 0.050101506, error in slope = 0.001348181
730     %Mass weighted slopes (Angstroms^-2 * fs^-1)
731     %basal slope = 4.76598E-06, error in slope = 1.94037E-07
732     %prismatic slope = 3.23131E-06, error in slope = 8.69514E-08
733     %**************************************************************