ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/iceWater/iceWater.tex
Revision: 3968
Committed: Wed Oct 23 23:25:02 2013 UTC (10 years, 10 months ago) by plouden
Content type: application/x-tex
File size: 40269 byte(s)
Log Message:
revised manuscript, added fit equation for d(basal) and d(prismatic). Edited text changing values for these displacements, and changed comparison text in conclusion to match the new arguement.

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 3967 atom in the ice slab. The resulting basal system was $23.87 \times 35.83
238     \times 98.64$ \AA\ with 900 SPC/E molecules in the ice slab, and 1846 in
239     the liquid phase. Similarly, the prismatic system was $36.12 \times 36.43
240     \times 86.10$ \AA\ with 1000 SPC/E molecules in the ice slab and 2684 in
241     the liquid.
242 gezelter 3914
243     Molecular translation and orientational restraints were applied in the
244     early stages of equilibration to prevent melting of the ice slab.
245     These restraints were removed during NVT equilibration, well before
246 gezelter 3945 data collection was carried out.
247 gezelter 3914
248 gezelter 3918 \subsection{Shearing ice / water interfaces without bulk melting}
249    
250 gezelter 3945 As a solid is dragged through a liquid, there is frictional heating
251     that will act to melt the interface. To study the behavior of the
252     interface under a shear stress without causing the interface to melt,
253 gezelter 3946 it is necessary to apply a weak thermal gradient in combination with
254     the momentum gradient. This can be accomplished using the velocity
255 gezelter 3945 shearing and scaling (VSS) variant of reverse non-equilibrium
256     molecular dynamics (RNEMD), which utilizes a series of simultaneous
257     velocity exchanges between two regions within the simulation
258     cell.\cite{Kuang12} One of these regions is centered within the ice
259 gezelter 3946 slab, while the other is centrally located in the liquid
260 gezelter 3918 region. VSS-RNEMD provides a set of conservation constraints for
261 gezelter 3946 creating either a momentum flux or a thermal flux (or both
262     simultaneously) between the two slabs. Satisfying the constraint
263     equations ensures that the new configurations are sampled from the
264     same NVE ensemble as before the VSS move.
265 gezelter 3918
266     The VSS moves are applied periodically to scale and shift the particle
267     velocities ($\mathbf{v}_i$ and $\mathbf{v}_j$) in two slabs ($H$ and
268     $C$) which are separated by half of the simulation box,
269     \begin{displaymath}
270     \begin{array}{rclcl}
271    
272     & \underline{\mathrm{shearing}} & &
273     \underline{~~~~~~~~~~~~\mathrm{scaling}~~~~~~~~~~~~} \\
274     \mathbf{v}_i \leftarrow &
275     \mathbf{a}_c\ & + & c\cdot\left(\mathbf{v}_i - \langle\mathbf{v}_c
276     \rangle\right) + \langle\mathbf{v}_c\rangle \\
277     \mathbf{v}_j \leftarrow &
278     \mathbf{a}_h & + & h\cdot\left(\mathbf{v}_j - \langle\mathbf{v}_h
279     \rangle\right) + \langle\mathbf{v}_h\rangle .
280    
281     \end{array}
282     \end{displaymath}
283     Here $\langle\mathbf{v}_c\rangle$ and $\langle\mathbf{v}_h\rangle$ are
284     the center of mass velocities in the $C$ and $H$ slabs, respectively.
285     Within the two slabs, particles receive incremental changes or a
286     ``shear'' to their velocities. The amount of shear is governed by the
287     imposed momentum flux, $\mathbf{j}_z(\mathbf{p})$
288     \begin{eqnarray}
289     \mathbf{a}_c & = & - \mathbf{j}_z(\mathbf{p}) \Delta t / M_c \label{vss1}\\
290     \mathbf{a}_h & = & + \mathbf{j}_z(\mathbf{p}) \Delta t / M_h \label{vss2}
291     \end{eqnarray}
292     where $M_{\{c,h\}}$ is the total mass of particles within each of the
293     slabs and $\Delta t$ is the interval between two separate operations.
294    
295     To simultaneously impose a thermal flux ($J_z$) between the slabs we
296     use energy conservation constraints,
297     \begin{eqnarray}
298     K_c - J_z\Delta t & = & c^2 (K_c - \frac{1}{2}M_c \langle\mathbf{v}_c
299     \rangle^2) + \frac{1}{2}M_c (\langle \mathbf{v}_c \rangle + \mathbf{a}_c)^2 \label{vss3}\\
300     K_h + J_z\Delta t & = & h^2 (K_h - \frac{1}{2}M_h \langle\mathbf{v}_h
301     \rangle^2) + \frac{1}{2}M_h (\langle \mathbf{v}_h \rangle +
302     \mathbf{a}_h)^2 \label{vss4}.
303     \label{constraint}
304     \end{eqnarray}
305     Simultaneous solution of these quadratic formulae for the scaling
306     coefficients, $c$ and $h$, will ensure that the simulation samples from
307     the original microcanonical (NVE) ensemble. Here $K_{\{c,h\}}$ is the
308     instantaneous translational kinetic energy of each slab. At each time
309     interval, it is a simple matter to solve for $c$, $h$, $\mathbf{a}_c$,
310     and $\mathbf{a}_h$, subject to the imposed momentum flux,
311     $j_z(\mathbf{p})$, and thermal flux, $J_z$, values. Since the VSS
312     operations do not change the kinetic energy due to orientational
313     degrees of freedom or the potential energy of a system, configurations
314     after the VSS move have exactly the same energy (and linear
315     momentum) as before the move.
316    
317     As the simulation progresses, the VSS moves are performed on a regular
318     basis, and the system develops a thermal and/or velocity gradient in
319 gezelter 3945 response to the applied flux. In a bulk material, it is quite simple
320 gezelter 3918 to use the slope of the temperature or velocity gradients to obtain
321 gezelter 3945 either the thermal conductivity or shear viscosity.
322 gezelter 3918
323     The VSS-RNEMD approach is versatile in that it may be used to
324     implement thermal and shear transport simultaneously. Perturbations
325     of velocities away from the ideal Maxwell-Boltzmann distributions are
326     minimal, as is thermal anisotropy. This ability to generate
327     simultaneous thermal and shear fluxes has been previously utilized to
328     map out the shear viscosity of SPC/E water over a wide range of
329 gezelter 3957 temperatures (90~K) with a single 1~ns simulation.\cite{Kuang12}
330 gezelter 3918
331 gezelter 3945 For this work, we are using the VSS-RNEMD method primarily to generate
332     a shear between the ice slab and the liquid phase, while using a weak
333 gezelter 3957 thermal gradient to maintain the interface at the 225~K target
334 gezelter 3946 value. This ensures minimal melting of the bulk ice phase and allows
335     us to control the exact temperature of the interface.
336 gezelter 3918
337 gezelter 3897 \subsection{Computational Details}
338 gezelter 3945 All simulations were performed using OpenMD,\cite{OOPSE,openmd} with a
339     time step of 2 fs and periodic boundary conditions in all three
340     dimensions. Electrostatics were handled using the damped-shifted
341     force real-space electrostatic kernel.\cite{Ewald} The systems were
342     divided into 100 bins along the $z$-axis for the VSS-RNEMD moves,
343 gezelter 3957 which were attempted every 50~fs.
344 gezelter 3897
345 gezelter 3945 The interfaces were equilibrated for a total of 10 ns at equilibrium
346     conditions before being exposed to either a shear or thermal gradient.
347     This consisted of 5 ns under a constant temperature (NVT) integrator
348     set to 225K followed by 5 ns under a microcanonical integrator. Weak
349     thermal gradients were allowed to develop using the VSS-RNEMD (NVE)
350 plouden 3966 integrator using a small thermal flux ($-2.0\times 10^{-6}$
351 gezelter 3945 kcal/mol/\AA$^2$/fs) for a duration of 5 ns to allow the gradient to
352 gezelter 3946 stabilize. The resulting temperature gradient was $\approx$ 10K over
353     the entire 100 \AA\ box length, which was sufficient to keep the
354 gezelter 3945 temperature at the interface within $\pm 1$ K of the 225K target.
355 gezelter 3897
356 gezelter 3945 Velocity gradients were then imposed using the VSS-RNEMD (NVE)
357     integrator with a range of momentum fluxes. These gradients were
358 gezelter 3957 allowed to stabilize for 1~ns before data collection began. Once
359     established, four successive 0.5~ns runs were performed for each shear
360 gezelter 3945 rate. During these simulations, snapshots of the system were taken
361 gezelter 3957 every 1~ps, and statistics on the structure and dynamics in each bin
362 gezelter 3967 were accumulated throughout the simulations. Although there was some
363     small variation in the measured interfacial width between succcessive
364     runs, no indication of bulk melting (or crystallization) was observed.
365 gezelter 3945
366 plouden 3941 \section{Results and discussion}
367    
368 gezelter 3946 \subsection{Interfacial width}
369     Any order parameter or time correlation function that changes as one
370     crosses an interface from a bulk liquid to a solid can be used to
371     measure the width of the interface. In previous work on the ice/water
372 gezelter 3954 interface, Haymet {\it et al.}\cite{Bryk02} have utilized structural
373 gezelter 3946 features (including the density) as well as dynamic properties
374     (including the diffusion constant) to estimate the width of the
375     interfaces for a number of facets of the ice crystals. Because
376     VSS-RNEMD imposes a lateral flow, parameters that depend on
377     translational motion of the molecules (e.g. diffusion) may be
378 gezelter 3954 artificially skewed by the RNEMD moves. A structural parameter is not
379 gezelter 3946 influenced by the RNEMD perturbations to the same degree. Here, we
380 gezelter 3954 have used the local tetrahedral order parameter as described by
381 gezelter 3946 Kumar\cite{Kumar09} and Errington\cite{Errington01} as our principal
382 gezelter 3967 measure of the interfacial width. A previous study by Bryk and Haymet
383     also used local tetrahedrality as an order parameter for ice/water
384     interfaces.\cite{Bryk2004b}
385 plouden 3936
386 gezelter 3945 The local tetrahedral order parameter, $q(z)$, is given by
387 gezelter 3897 \begin{equation}
388 gezelter 3946 q(z) = \int_0^L \sum_{k=1}^{N} \Bigg(1 -\frac{3}{8}\sum_{i=1}^{3}
389     \sum_{j=i+1}^{4} \bigg(\cos\psi_{ikj}+\frac{1}{3}\bigg)^2\Bigg)
390     \delta(z_{k}-z)\mathrm{d}z \Bigg/ N_z
391 gezelter 3945 \label{eq:qz}
392 gezelter 3897 \end{equation}
393 gezelter 3945 where $\psi_{ikj}$ is the angle formed between the oxygen site on
394 gezelter 3946 central molecule $k$, and the oxygen sites on two of the four closest
395     molecules, $i$ and $j$. Molecules $i$ and $j$ are further restricted
396 gezelter 3967 to lie withing the first peak in the pair distribution function for
397     molecule $k$ (typically $<$ 3.41 \AA\ for water). $N_z = \int
398 gezelter 3946 \delta(z_k - z) \mathrm{d}z$ is a normalization factor to account for
399     the varying population of molecules within each finite-width bin. The
400     local tetrahedral order parameter has a range of $(0,1)$, where the
401     larger values of $q$ indicate a larger degree of tetrahedral ordering
402     of the local environment. In perfect ice I$_\mathrm{h}$ structures,
403     the parameter can approach 1 at low temperatures, while in liquid
404     water, the ordering is significantly less tetrahedral, and values of
405     $q(z) \approx 0.75$ are more common.
406 plouden 3909
407 gezelter 3946 To estimate the interfacial width, the system was divided into 100
408 gezelter 3967 bins along the $z$-dimension. The $q_{z}$ function was time-averaged
409     to give yield a tetrahedrality profile of the system. The profile was
410     then fit to a hyperbolic tangent that smoothly links the liquid and
411     solid states,
412 plouden 3941 \begin{equation}\label{tet_fit}
413 gezelter 3946 q(z) \approx
414     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-
415     \frac{r+l}{2}\right|.
416 gezelter 3897 \end{equation}
417 gezelter 3946 Here $q_{liq}$ and $q_{ice}$ are the local tetrahedral order parameter
418     for the bulk liquid and ice domains, respectively, $w$ is the width of
419     the interface. $l$ and $r$ are the midpoints of the left and right
420 gezelter 3957 interfaces, respectively. The last term in eq. \eqref{tet_fit}
421 gezelter 3946 accounts for the influence that the weak thermal gradient has on the
422 gezelter 3967 tetrahedrality profile in the liquid region.
423 gezelter 3897
424 gezelter 3946 In Figures \ref{fig:bComic} and \ref{fig:pComic} we see the
425     $z$-coordinate profiles for tetrahedrality, temperature, and the
426     $x$-component of the velocity for the basal and prismatic interfaces.
427     The lower panels show the $q(z)$ (black circles) along with the
428     hyperbolic tangent fits (red lines). In the liquid region, the local
429     tetrahedral order parameter, $q(z) \approx 0.75$ while in the
430     crystalline region, $q(z) \approx 0.94$, indicating a more tetrahedral
431     environment. The vertical dotted lines denote the midpoint of the
432 gezelter 3957 interfaces ($r$ and $l$ in eq. \eqref{tet_fit}). The weak thermal
433 gezelter 3946 gradient applied to the systems in order to keep the interface at
434 gezelter 3957 225~$\pm$~5K, can be seen in middle panels. The transverse velocity
435 gezelter 3946 profile is shown in the upper panels. It is clear from the upper
436     panels that water molecules in close proximity to the surface (i.e.
437 gezelter 3967 within 10~\AA\ to 15~\AA~of the interfaces) have transverse velocities
438     quite close to the velocities within the ice block. There is no
439     velocity discontinuity at the interface, which indicates that the
440     shearing of ice/water interfaces occurs in the ``stick'' or no-slip
441     boundary conditions.
442 gezelter 3897
443 gezelter 3914 \begin{figure}
444 gezelter 3957 \includegraphics[width=\linewidth]{bComicStrip}
445 gezelter 3967 \caption{\label{fig:bComic} The basal interface with a shear rate of
446 plouden 3968 1.3 ms\textsuperscript{-1}. Lower panel: the local tetrahedral order parameter, $q(z)$,
447 gezelter 3967 (black circles) and the hyperbolic tangent fit (red line). Middle
448     panel: the imposed thermal gradient required to maintain a fixed
449     interfacial temperature. Upper panel: the transverse velocity
450     gradient that develops in response to an imposed momentum flux. The
451     vertical dotted lines indicate the locations of the midpoints of the
452     two interfaces.}
453 gezelter 3914 \end{figure}
454 plouden 3904
455 gezelter 3914 \begin{figure}
456 gezelter 3957 \includegraphics[width=\linewidth]{pComicStrip}
457 gezelter 3967 \caption{\label{fig:pComic} The prismatic interface with a shear rate
458 plouden 3968 of 2.0 ms\textsuperscript{-1}. Panel
459 gezelter 3946 descriptions match those in figure \ref{fig:bComic}}
460 gezelter 3914 \end{figure}
461    
462 gezelter 3967 From the fits using eq. \eqref{tet_fit}, we find the interfacial width
463     for the basal and prismatic systems to be 3.2~$\pm$~0.4~\AA\ and
464 gezelter 3957 3.6~$\pm$~0.2~\AA\ , respectively, with no applied momentum flux. Over
465 gezelter 3946 the range of shear rates investigated, $0.6 \pm 0.3 \mathrm{ms}^{-1}
466     \rightarrow 5.3 \pm 0.5 \mathrm{ms}^{-1}$ for the basal system and
467     $0.9 \pm 0.2 \mathrm{ms}^{-1} \rightarrow 4.5 \pm 0.1
468     \mathrm{ms}^{-1}$ for the prismatic, we found no appreciable change in
469     the interface width. The fit values for the interfacial width ($w$)
470     over all shear rates contained the values reported above within their
471 gezelter 3967 error bars. Note that the interfacial widths reported here are based
472     on the hyperbolic tangent parameter $w$ in Eq. \ref{tet_fit}. This is
473     related to, but not identical with, the 10\%-90\% intefacial widths
474     commonly used in previous studies.\cite{Bryk02,Bryk2004b} To estimate
475     the 10\%-90\% widths, it is a simple matter to scale the widths
476     obtained from the hyperbolic tangent fits to obtain $w_{10-90} =
477     2.1971 \times w$.\cite{Bryk02,Bryk2004b} This results in $w_{10-90}$
478     values of 7.0~$\pm$~0.9~\AA\ for the basal face, and 7.9~$\pm$~0.4
479     \AA\ for the prismatic face. These are somewhat smaller than
480     previously reported values.
481 gezelter 3914
482 gezelter 3967 \begin{figure}
483     \includegraphics[width=\linewidth]{interface_width_by_shear_rate}
484     \caption{\label{fig:widthByShear} The width of the ice water
485     interfaces (as measured by Eq. \ref{tet_fit}) exhibits no dependence
486     on the applied shear rate between the ice and water regions.}
487     \end{figure}
488    
489    
490    
491 gezelter 3954 \subsubsection{Orientational Dynamics}
492 gezelter 3946 The orientational time correlation function,
493 plouden 3941 \begin{equation}\label{C(t)1}
494 gezelter 3946 C_{2}(t)=\langle P_{2}(\mathbf{u}(0) \cdot \mathbf{u}(t)) \rangle,
495 plouden 3941 \end{equation}
496 gezelter 3946 gives insight into the local dynamic environment around the water
497     molecules. The rate at which the function decays provides information
498     about hindered motions and the timescales for relaxation. In
499     eq. \eqref{C(t)1}, $P_{2}$ is the second-order Legendre polynomial,
500 gezelter 3948 the vector $\mathbf{u}$ is often taken as HOH bisector, although
501     slightly different behavior can be observed when $\mathbf{u}$ is the
502 gezelter 3946 vector along one of the OH bonds. The angle brackets denote an
503     ensemble average over all water molecules in a given spatial region.
504 gezelter 3897
505 gezelter 3946 To investigate the dynamic behavior of water at the ice interfaces, we
506     have computed $C_{2}(z,t)$ for molecules that are present within a
507     particular slab along the $z$- axis at the initial time. The change
508     in the decay behavior as a function of the $z$ coordinate is another
509 gezelter 3948 measure of the change of how the local environment changes across the
510     ice/water interface. To compute these correlation functions, each of
511     the 0.5 ns simulations was followed by a shorter 200 ps microcanonical
512     (NVE) simulation in which the positions and orientations of every
513     molecule in the system were recorded every 0.1 ps. The systems were
514 gezelter 3954 then divided into 30 bins along the $z$-axis and $C_2(t)$ was
515     evaluated for each bin.
516 plouden 3919
517 gezelter 3948 In simulations of water at biological interfaces, Furse {\em et al.}
518     fit $C_2(t)$ functions for water with triexponential
519     functions,\cite{Furse08} where the three components of the decay
520 gezelter 3957 correspond to a fast ($<$200 fs) reorientational piece driven by the
521 gezelter 3948 restoring forces of existing hydrogen bonds, a middle (on the order of
522     several ps) piece describing the large angle jumps that occur during
523     the breaking and formation of new hydrogen bonds,and a slow (on the
524     order of tens of ps) contribution describing the translational motion
525     of the molecules. The model for orientational decay presented
526     recently by Laage and Hynes {\em et al.}\cite{Laage08,Laage11} also
527     includes three similar decay constants, although two of the time
528     constants are linked, and the resulting decay curve has two parameters
529     governing the dynamics of decay.
530    
531     In our ice/water interfaces, we are at substantially lower
532     temperatures, and the water molecules are further perturbed by the
533     presence of the ice phase nearby. We have obtained the most
534     reasonable fits using triexponential functions with three distinct
535 gezelter 3954 time domains, as well as a constant piece to account for the water
536 gezelter 3948 stuck in the ice phase that does not experience any long-time
537     orientational decay,
538 gezelter 3897 \begin{equation}
539 gezelter 3946 C_{2}(t) \approx a e^{-t/\tau_\mathrm{short}} + b e^{-t/\tau_\mathrm{middle}} + c
540     e^{-t/\tau_\mathrm{long}} + (1-a-b-c)
541 gezelter 3897 \end{equation}
542 gezelter 3948 Average values for the three decay constants (and error estimates)
543     were obtained for each bin. In figures \ref{fig:basal_Tau_comic_strip}
544     and \ref{fig:prismatic_Tau_comic_strip}, the three orientational decay
545     times are shown as a function of distance from the center of the ice
546     slab.
547 gezelter 3897
548 plouden 3941 \begin{figure}
549 gezelter 3957 \includegraphics[width=\linewidth]{basal_Tau_comic_strip}
550 gezelter 3949 \caption{\label{fig:basal_Tau_comic_strip} The three decay constants
551     of the orientational time correlation function, $C_2(t)$, for water
552     as a function of distance from the center of the ice slab. The
553     dashed line indicates the location of the basal face (as determined
554     from the tetrahedrality order parameter). The moderate and long
555     time contributions slow down close to the interface which would be
556 gezelter 3954 expected under reorganizations that involve large motions of the
557 gezelter 3949 molecules (e.g. frame-reorientations and jumps). The observed
558     speed-up in the short time contribution is surprising, but appears
559     to reflect the restricted motion of librations closer to the
560     interface.}
561 plouden 3941 \end{figure}
562 gezelter 3897
563 gezelter 3914 \begin{figure}
564 gezelter 3957 \includegraphics[width=\linewidth]{prismatic_Tau_comic_strip}
565 gezelter 3949 \caption{\label{fig:prismatic_Tau_comic_strip}
566     Decay constants for $C_2(t)$ at the prismatic interface. Panel
567     descriptions match those in figure \ref{fig:basal_Tau_comic_strip}.}
568 gezelter 3914 \end{figure}
569 plouden 3904
570 gezelter 3949 Figures \ref{fig:basal_Tau_comic_strip} and
571     \ref{fig:prismatic_Tau_comic_strip} show the three decay constants for
572     the orientational time correlation function for water at varying
573     displacements from the center of the ice slab for both the basal and
574     prismatic interfaces. The vertical dotted lines indicate the
575     locations of the midpoints of the interfaces as determined by the
576     tetrahedrality fits. In the liquid regions, $\tau_{middle}$ and
577     $\tau_{long}$ have consistent values around 3-4 ps and 20-40 ps,
578     respectively, and increase in value approaching the interface.
579     According to the jump model of Laage and Hynes {\em et
580     al.},\cite{Laage08,Laage11} $\tau_{middle}$ corresponds to the
581     breaking and making of hydrogen bonds and $\tau_{long}$ is explained
582     with translational motion of the molecules (i.e. frame reorientation).
583     The shortest of the three decay constants, the librational time
584     $\tau_\mathrm{short}$ has a value of about 70 fs in the liquid region,
585     and decreases in value approaching the interface. The observed
586     speed-up in the short time contribution is surprising, but appears to
587     reflect the restricted motion of librations closer to the interface.
588 plouden 3937
589 gezelter 3949 The control systems (with no applied momentum flux) are shown with
590     black symbols in figs. \ref{fig:basal_Tau_comic_strip} and
591     \ref{fig:prismatic_Tau_comic_strip}, while those obtained while a
592     shear was active are shown in red.
593 plouden 3937
594 gezelter 3952 Two notable features deserve clarification. First, there are
595     nearly-constant liquid-state values for $\tau_{short}$,
596 gezelter 3949 $\tau_{middle}$, and $\tau_{long}$ at large displacements from the
597 gezelter 3952 interface. Second, there appears to be a single distance, $d_{basal}$
598     or $d_{prismatic}$, from the interface at which all three decay times
599 plouden 3968 begin to deviate from their bulk liquid values. To quantify this
600     distance, each of the decay constant $z$-profiles were fit to
601     \begin{equation}\label{tauFit}
602     \tau(z)\approx\tau_{liquid}+(\tau_{solid}-\tau_{liquid})e^{-(z-z_{wall})/d}
603     \end{equation}
604     where $\tau_{liquid}$ and $\tau_{solid}$ are the liquid and projected
605     solid values of the decay constants, $z_{wall}$ is the location of the
606     interface, and $d$ is the displacement the deviations occur at (see
607     Figures \ref{fig:basal_Tau_comic_strip} and
608     \ref{fig:prismatic_Tau_comic_strip}). The displacements $d_{basal}$
609     and $d_{prismatic}$ were determined for each of the three decay
610     constants, and then averaged for better statistics.
611     For the basal system, we found $d_{basal}$ for the control set to be
612     2.9 \AA\, and 2.8 \AA\ for a simulation with a shear rate of 1.3
613     ms\textsuperscript{-1}. We found $d_{prismatic}$ to be slightly
614     larger than $d_{basal}$ for both the control and an applied shear,
615     with displacements of 3.6 \AA\ for the control system and 3.5 \AA\ for
616     a simulation with a 2 ms\textsuperscript{-1} shear rate. From this we
617     can conclude there is no apparent dependence on the shear rate for the dynamic interface
618     width.
619 plouden 3937
620 plouden 3968 %%%%%%%%Should we keep this paragraph???%%%%%%%%%%%%%%%
621 gezelter 3952 Beaglehole and Wilson have measured the ice/water interface using
622 gezelter 3957 ellipsometry and find a thickness of approximately 10~\AA\ for both
623 gezelter 3952 the basal and prismatic faces.\cite{Beaglehole93} Structurally, we
624 gezelter 3957 have found the basal and prismatic interfacial width to be
625 plouden 3968 3.2~$\pm$~0.4~\AA\ and 3.6~$\pm$~0.2~\AA. Decomposition of
626     the spatial dependence of the decay times of $C_2(t)$ shows good
627     agreement with the structural interfacial width determined by the
628     local tetrahedrality.
629     %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
630 plouden 3937
631 gezelter 3952
632 plouden 3941 \subsection{Coefficient of Friction of the Interface}
633 gezelter 3954 As liquid water flows over an ice interface, there is a distance from
634     the structural interface where bulk-like hydrodynamics are recovered.
635     Bocquet and Barrat constructed a theory for the hydrodynamic boundary
636     parameters, which include the slipping length
637 gezelter 3952 $\left(\delta_\mathrm{wall}\right)$ of this boundary layer and the
638     ``hydrodynamic position'' of the boundary
639 gezelter 3954 $\left(z_\mathrm{wall}\right)$.\cite{PhysRevLett.70.2726,PhysRevE.49.3079}
640     This last parameter is the location (relative to a solid surface)
641 gezelter 3952 where the bulk-like behavior is recovered. Work by Mundy {\it et al.}
642 gezelter 3954 has helped to combine these parameters into a liquid-solid friction
643 gezelter 3952 coefficient, which quantifies the resistance to pulling the solid
644 gezelter 3954 interface through a liquid,\cite{Mundy1997305}
645 gezelter 3952 \begin{equation}
646     \lambda_\mathrm{wall} = \frac{\eta}{\delta_\mathrm{wall}}.
647     \end{equation}
648 gezelter 3954 This expression is nearly identical to one provided by Pit {\it et
649     al.} for the solid-liquid friction of an interface,\cite{Pit99}
650 plouden 3942 \begin{equation}\label{Pit}
651 gezelter 3952 \lambda=\frac{\eta}{\delta}
652 plouden 3942 \end{equation}
653 gezelter 3952 where $\delta$ is the slip length for the liquid measured at the
654 gezelter 3967 location of the interface itself. In our simulations, the shoulder on
655     the velocity profile indicating the location of the hydrodynamic
656     boundary in the liquid is not always apparent. In some cases, the
657     linear behavior persists nearly up to the interfacial region. For
658     this reason, the hydrodynamic position of the boundary is not always
659     computable, while the Pit approach (Eq. \ref{Pit}) can be used to find
660     the solid-liquid friction coefficient more reliably.
661 gezelter 3952
662 gezelter 3967 In both the Pit and hydrodynamic boundary expressions, $\eta$ is the
663     shear viscosity of the bulk-like region of the liquid, a quantity
664     which is easily obtained in VSS-RNEMD simulations by fitting the
665     velocity profile in the region far from the surface.\cite{Kuang12}
666     Assuming linear response in the bulk-like region,
667 plouden 3942 \begin{equation}\label{Kuang}
668 gezelter 3952 j_{z}(p_{x})=-\eta \left(\frac{\partial v_{x}}{\partial z}\right)
669 plouden 3942 \end{equation}
670 gezelter 3952 Substituting this result into eq. \eqref{Pit}, we can estimate the
671     solid-liquid coefficient using the slip length,
672 plouden 3941 \begin{equation}
673 gezelter 3954 \lambda=-\frac{j_{z}(p_{x})} {\left(\frac{\partial v_{x}}{\partial
674     z}\right) \delta}
675 plouden 3941 \end{equation}
676 plouden 3937
677 gezelter 3967 For ice / water interfaces, the boundary conditions are no-slip, so
678     projecting the bulk liquid state velocity profile yields a negative
679     slip length. This length is the difference between the structural edge
680     of the ice (determined by the tetrahedrality profile) and the location
681     where the projected velocity of the bulk liquid intersects the solid
682     phase velocity (see Figure \ref{fig:delta_example}). The coefficients
683     of friction for the basal and the prismatic facets were determined for
684     shearing along both the $x$ and $y$ axes. The values are given in
685     table \ref{tab:lambda}.
686 plouden 3939
687 gezelter 3967 Note that the measured friction coefficient for the basal face is
688     twice that of the prismatic face (regardless of drag direction).
689     These results may seem surprising as the basalface appears smoother
690     than the prismatic with only small undulations of the oxygen
691     positions, while the prismatic surface has deep corrugated channels
692     along the $x$ direction in the crystal system used in this work.
693     However, the corrugations are relatively thin, and the liquid phase
694     water does not appear to populate the channels. The prismatic face
695     therefore effectively presents stripes of solid-phase molecules
696     (making up approximately half of the exposed surface area) with nearly
697     empty space between them. The interfacial friction appears to be
698     independent of the drag direction, so flow parallel to these channels
699     does not explain the lower friction of the prismatic face. A more
700     likely explanation is that the effective contact between the liquid
701     phase and the prismatic face is reduced by the empty corrugations.
702    
703     \begin{figure}
704 gezelter 3957 \includegraphics[width=\linewidth]{delta_example}
705 gezelter 3952 \caption{\label{fig:delta_example} Determining the (negative) slip
706     length ($\delta$) for the ice-water interfaces (which have decidedly
707     non-slip behavior). This length is the difference between the
708     structural edge of the ice (determined by the tetrahedrality
709     profile) and the location where the projected velocity of the bulk
710     liquid (dashed red line) intersects the solid phase velocity (solid
711     black line). The dotted line indicates the location of the ice as
712 gezelter 3967 determined by the tetrahedrality profile. This example is taken
713 plouden 3968 from the basal-face simulation with an applied shear rate of 3.0 ms\textsuperscript{-1}.}
714 plouden 3942 \end{figure}
715    
716    
717 gezelter 3967 \begin{table}[h]
718     \centering
719     \caption{Solid-liquid friction coefficients (measured in amu~fs\textsuperscript{-1}) }
720     \label{tab:lambda}
721     \begin{tabular}{|ccc|} \hline
722     & \multicolumn{2}{c|}{Drag direction} \\
723     Interface & $x$ & $y$ \\ \hline
724     basal & $0.08 \pm 0.02$ & $0.09 \pm 0.03$ \\
725     prismatic & $0.037 \pm 0.008$ & $0.04 \pm 0.01$ \\ \hline
726     \end{tabular}
727     \end{table}
728    
729    
730 gezelter 3897 \section{Conclusion}
731 gezelter 3952 We have simulated the basal and prismatic facets of an SPC/E model of
732     the ice I$_\mathrm{h}$ / water interface. Using VSS-RNEMD, the ice
733     was sheared relative to the liquid while simultaneously being exposed
734     to a weak thermal gradient which kept the interface at a stable
735     temperature. Calculation of the local tetrahedrality order parameter
736     has shown an apparent independence of the interfacial width on the
737 gezelter 3957 shear rate. This width was found to be 3.2~$\pm$0.4~\AA\ and
738     3.6~$\pm$0.2~\AA\ for the basal and prismatic systems, respectively.
739 gezelter 3897
740 gezelter 3952 Orientational time correlation functions were calculated at varying
741     displacements from the interface, and were found to be similarly
742     independent of the applied momentum flux. The short decay due to the
743     restoring forces of existing hydrogen bonds decreased close to the
744     interface, while the longer-time decay constants increased in close
745     proximity to the interface. There is also an apparent dynamic
746     interface width, $d_{basal}$ and $d_{prismatic}$, at which these
747 gezelter 3954 deviations from bulk liquid values begin. We found $d_{basal}$ and
748 plouden 3968 $d_{prismatic}$ to be approximately 2.8~\AA\ and 3.5~\AA\ . This
749     interfacial width is in good agreement with values determined by the
750     structural analysis of the interface, by the hyperbolic tangent fit of
751     the local tetrahedral order parameter.
752 gezelter 3952
753 gezelter 3954 The coefficient of liquid-solid friction for each of the facets was
754 gezelter 3957 also determined. They were found to be
755     0.07~$\pm$~0.01~amu~fs\textsuperscript{-1} and
756     0.032~$\pm$~0.007~amu~fs\textsuperscript{-1} for the basal and
757     prismatic facets respectively. We attribute the large difference
758     between the two friction coefficients to the direction of the shear
759     and to the surface structure of the crystal facets.
760 gezelter 3952
761 gezelter 3957 \section{Acknowledgements}
762 gezelter 3914 Support for this project was provided by the National Science
763     Foundation under grant CHE-0848243. Computational time was provided
764     by the Center for Research Computing (CRC) at the University of
765     Notre Dame.
766 gezelter 3897
767 gezelter 3914 \newpage
768 plouden 3909 \bibliography{iceWater}
769 gezelter 3897
770 gezelter 3957 \end{doublespace}
771 gezelter 3897
772 gezelter 3957 % \begin{tocentry}
773     % \begin{wrapfigure}{l}{0.5\textwidth}
774     % \begin{center}
775     % \includegraphics[width=\linewidth]{SystemImage.png}
776     % \end{center}
777     % \end{wrapfigure}
778     % The cell used to simulate liquid-solid shear in ice I$_\mathrm{h}$ /
779     % water interfaces. Velocity gradients were applied using the velocity
780     % shearing and scaling variant of reverse non-equilibrium molecular
781     % dynamics (VSS-RNEMD) with a weak thermal gradient to prevent melting.
782     % The interface width is relatively robust in both structual and dynamic
783     % measures as a function of the applied shear.
784     % \end{tocentry}
785    
786 gezelter 3914 \end{document}
787 gezelter 3897
788 plouden 3924 %**************************************************************
789     %Non-mass weighted slopes (amu*Angstroms^-2 * fs^-1)
790     % basal: slope=0.090677616, error in slope = 0.003691743
791     %prismatic: slope = 0.050101506, error in slope = 0.001348181
792     %Mass weighted slopes (Angstroms^-2 * fs^-1)
793     %basal slope = 4.76598E-06, error in slope = 1.94037E-07
794     %prismatic slope = 3.23131E-06, error in slope = 8.69514E-08
795     %**************************************************************