ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/iceWater/iceWater.tex
(Generate patch)

Comparing trunk/iceWater/iceWater.tex (file contents):
Revision 3949 by gezelter, Fri Sep 6 22:02:54 2013 UTC vs.
Revision 3966 by plouden, Wed Oct 23 17:29:18 2013 UTC

# Line 1 | Line 1
1 < \documentclass[journal = jpccck, manuscript = article]{achemso}
2 < \setkeys{acs}{usetitle = true}
3 < \usepackage{achemso}
4 < \usepackage{natbib}
1 > \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   \usepackage{multirow}
11   \usepackage{wrapfig}
12 < \usepackage{fixltx2e}
13 < %\mciteErrorOnUnknownfalse
12 > %\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  
22 + % 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   \usepackage[version=3]{mhchem}  % this is a great package for formatting chemical reactions
41   \usepackage{url}
42  
43 +
44 + \begin{document}
45 +
46   \title{Simulations of solid-liquid friction at ice-I$_\mathrm{h}$ /
47    water interfaces}
48  
49 < \author{P. B. Louden}
50 < \author{J. Daniel Gezelter}
51 < \email{gezelter@nd.edu}
52 < \affiliation[University of Notre Dame]{251 Nieuwland Science Hall\\
53 <  Department of Chemistry and Biochemistry\\ University of Notre
54 <  Dame\\ Notre Dame, Indiana 46556}
49 > \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  
56 < \keywords{}
56 > \date{\today}
57 > \maketitle
58 > \begin{doublespace}
59  
25 \begin{document}
60  
61   \begin{abstract}
62    We have investigated the structural and dynamic properties of the
63 <  basal and prismatic facets of an ice I$_\mathrm{h}$ / water
64 <  interface when the solid phase is being drawn through the liquid
63 >  basal and prismatic facets of the ice I$_\mathrm{h}$ / water
64 >  interface when the solid phase is drawn through the liquid
65    (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 <  width was found to be independent of relative velocity of the ice
71 <  and liquid layers over a wide range of shear rates.  Decays of
72 <  molecular orientational time correlation functions for gave very
73 <  similar estimates for the width of the interfaces, although the
74 <  short- and longer-time decay components of the orientational
75 <  correlation functions behave differently closer to the interface.
76 <  Although both facets of ice are in ``stick'' boundary conditions in
77 <  liquid water, the solid-liquid friction coefficient was found to be
78 <  different for the basal and prismatic facets of ice.
70 >  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   \end{abstract}
80  
81   \newpage
# Line 67 | Line 101 | interface for the SPC, SPC/E, CF1, and TIP4P models fo
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 < interface for the SPC, SPC/E, CF1, and TIP4P models for
105 < water.\cite{Karim87,Karim88,Karim90,Hayward01,Bryk02,Hayward02,Gay02}
104 > interface for the SPC,\cite{Karim90} SPC/E,\cite{Gay02,Bryk02} CF1,\cite{Hayward01,Hayward02} and TIP4P~\cite{Karim88} models for
105 > water.
106   More recently, Haymet \emph{et al.} have investigated the effects
107   cations and anions have on crystal
108 < nucleation\cite{Bryk04,Smith05,Wilson08,Wilson10}. Nada \emph{et al.}
108 > nucleation.\cite{Bryk04,Smith05,Wilson08,Wilson10} Nada \emph{et al.}
109   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 < reordering of the hydrogen bonding network\cite{Nada05}.
112 > reordering of the hydrogen bonding network.\cite{Nada05}
113  
114   The movement of liquid water over the facets of ice has been less
115   thoroughly studied than the quiescent surfaces. This process is
# Line 121 | Line 155 | and Oxygen lone pairs.\cite{Buch:2008fk}
155   have a preference for proton ordering along strips of dangling H-atoms
156   and Oxygen lone pairs.\cite{Buch:2008fk}
157  
158 < \begin{wraptable}{r}{3.5in}
158 > \begin{table}[h]
159 > \centering
160    \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 <    system in reference \protect\cite{Hirsch04}}
162 >    system in reference \bibpunct{}{}{,}{n}{}{,} \protect\citep{Hirsch04}.}
163   \label{tab:equiv}
164   \begin{tabular}{|ccc|} \hline
165   & hexagonal & orthorhombic \\
# Line 135 | Line 170 | pyramidal & $\{2~0~\bar{2}~1\}$ & $\{2~0~1\}$ \\ \hlin
170   secondary prism & $\{1~1~\bar{2}~0\}$ & $\{1~3~0\}$ \\
171   pyramidal & $\{2~0~\bar{2}~1\}$ & $\{2~0~1\}$ \\ \hline
172   \end{tabular}
173 < \end{wraptable}
139 <
173 > \end{table}
174   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
# Line 154 | Line 188 | parameters $a = 4.49225$, $b = 7.78080$, $c = 7.33581$
188   $P2_12_12_1$ system.
189  
190   Structure 6 from the Hirsch and Ojam\"{a}e paper has lattice
191 < parameters $a = 4.49225$, $b = 7.78080$, $c = 7.33581$ and two water
192 < molecules whose atoms reside at fractional coordinates given in table
193 < \ref{tab:p212121}.  To construct the basal and prismatic interfaces,
194 < these crystallographic coordinates were used to construct an
195 < orthorhombic unit cell which was then replicated in all three
191 > 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   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 < respectively.  The resulting block was rotated so that the exposed
200 < faces were aligned with the $z$-axis normal to the exposed face.  The
199 > 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   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   creating a slab with either the basal or prismatic faces exposed along
204 < the $z$ axis.  The slab was then replicated in the $x$ and $y$
204 > the $z$ axis. The slab was then replicated in the $x$ and $y$
205   dimensions until a desired sample size was obtained.
206  
207 < \begin{wraptable}{r}{2.85in}
207 > \begin{table}[h]
208 > \centering
209    \caption{Fractional coordinates for water in the orthorhombic
210 <    $P2_12_12_1$ system for ice I$_\mathrm{h}$ in reference
176 <    \protect\cite{Hirsch04}}
210 >    $P2_12_12_1$ system for ice I$_\mathrm{h}$ in reference \bibpunct{}{}{,}{n}{}{,} \protect\citep{Hirsch04}.}
211   \label{tab:p212121}
212   \begin{tabular}{|cccc|}  \hline
213   atom type & x & y & z \\ \hline
214 < O & 0.75 & 0.1667 & 0.4375 \\
214 > O & 0.7500 & 0.1667 & 0.4375 \\
215   H & 0.5735 & 0.2202 & 0.4836 \\
216   H & 0.7420 & 0.0517 & 0.4836 \\
217 < O & 0.25 & 0.6667 & 0.4375 \\
217 > O & 0.2500 & 0.6667 & 0.4375 \\
218   H & 0.2580 & 0.6693 & 0.3071 \\
219   H & 0.4265 & 0.7255 & 0.4756 \\ \hline
220   \end{tabular}
221 < \end{wraptable}
221 > \end{table}
222  
223   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 < atmospheric pressure is close to 273K, Haymet \emph{et al.} have done
226 > atmospheric pressure is close to 273~K, Haymet \emph{et al.} have done
227   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 < 225$\pm$5K.\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
232 > 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   allowed to fluctuate to equilibrate to the correct pressure).  The
235   liquid and solid systems were combined by carving out any water
236 < molecule from the liquid simulation cell that was within 3 \AA\ of any
237 < atom in the ice slab.
236 > molecule from the liquid simulation cell that was within 3~\AA\ of any
237 > atom in the ice slab. The resulting basal system was 24 \AA\ x 36 \AA\ x 99 \AA\ with 900 SPC/E molecules in the ice slab and 1846 SPC/E molecules in the liquid phase. Similarly, the prismatic system was constructed as 36 \AA\ x 36 \AA\ x 86 \AA\ with 1000 SPC/E molecules in the ice slab and 2684 SPC/E molecules in the liquid phase.
238  
239   Molecular translation and orientational restraints were applied in the
240   early stages of equilibration to prevent melting of the ice slab.
# Line 288 | Line 322 | temperatures (90~K) with a single 1 ns simulation.\cit
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 < temperatures (90~K) with a single 1 ns simulation.\cite{Kuang12}
325 > temperatures (90~K) with a single 1~ns simulation.\cite{Kuang12}
326  
327   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 < thermal gradient to maintain the interface at the 225K target
329 > thermal gradient to maintain the interface at the 225~K target
330   value. This ensures minimal melting of the bulk ice phase and allows
331   us to control the exact temperature of the interface.
332  
# Line 302 | Line 336 | which were attempted every 50 fs.
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 < which were attempted every 50 fs.
339 > which were attempted every 50~fs.
340  
341   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 ($-2.0\times 10^{-6}$
346 > integrator using 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   stabilize.  The resulting temperature gradient was $\approx$ 10K over
349   the entire 100 \AA\  box length, which was sufficient to keep the
# Line 317 | Line 351 | allowed to stabilize for 1 ns before data collection b
351  
352   Velocity gradients were then imposed using the VSS-RNEMD (NVE)
353   integrator with a range of momentum fluxes.  These gradients were
354 < allowed to stabilize for 1 ns before data collection began. Once
355 < established, four successive 0.5 ns runs were performed for each shear
354 > 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   rate.  During these simulations, snapshots of the system were taken
357 < every 1 ps, and statistics on the structure and dynamics in each bin
357 > every 1~ps, and statistics on the structure and dynamics in each bin
358   were accumulated throughout the simulations.
359  
360   \section{Results and discussion}
# Line 329 | Line 363 | interface, Haymet {\it et al.}\cite{} have utilized st
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 < interface, Haymet {\it et al.}\cite{} have utilized structural
366 > interface, Haymet {\it et al.}\cite{Bryk02} have utilized structural
367   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 < artifically skewed by the RNEMD moves.  A structural parameter is not
372 > artificially skewed by the RNEMD moves.  A structural parameter is not
373   influenced by the RNEMD perturbations to the same degree. Here, we
374 < have used the local tetraherdal order parameter as described by
374 > have used the local tetrahedral order parameter as described by
375   Kumar\cite{Kumar09} and Errington\cite{Errington01} as our principal
376   measure of the interfacial width.
377  
# Line 363 | Line 397 | solvation shell was set to 3.41 \AA\ .  The $q_{z}$ fu
397  
398   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 < solvation shell was set to 3.41 \AA\ .  The $q_{z}$ function was
400 > solvation shell was set to 3.41~\AA\ .  The $q_{z}$ function was
401   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,
# Line 375 | Line 409 | interfaces, respectively.  The last term in equation \
409   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 < interfaces, respectively.  The last term in equation \eqref{tet_fit}
412 > interfaces, respectively.  The last term in eq. \eqref{tet_fit}
413   accounts for the influence that the weak thermal gradient has on the
414   tetrahedrality profile in the liquid region.  To estimate the
415 < 10\%-90\% widths commonly used in previous studies,\cite{} it is a
416 < simple matter to scale the widths obtained from the hyberbolic tangent
417 < fits to obtain $w_{10-90} = 2.9 w$.\cite{}
415 > 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  
419   In Figures \ref{fig:bComic} and \ref{fig:pComic} we see the
420   $z$-coordinate profiles for tetrahedrality, temperature, and the
# Line 390 | Line 424 | interfaces ($r$ and $l$ in equation \eqref{tet_fit}).
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 < interfaces ($r$ and $l$ in equation \eqref{tet_fit}). The weak thermal
427 > interfaces ($r$ and $l$ in eq. \eqref{tet_fit}). The weak thermal
428   gradient applied to the systems in order to keep the interface at
429 < 225$\pm$5K, can be seen in middle panels.  The tranverse velocity
429 > 225~$\pm$~5K, can be seen in middle panels.  The transverse velocity
430   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 < within 10 \AA\ to 15 \AA\ of the interfaces) have transverse
432 > within 10~\AA\ to 15~\AA~of the interfaces) have transverse
433   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  
438   \begin{figure}
439 < \includegraphics[width=\linewidth]{bComicStrip.pdf}
439 > \includegraphics[width=\linewidth]{bComicStrip}
440   \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
# Line 414 | Line 448 | no-slip boundary conditions.
448   \end{figure}
449  
450   \begin{figure}
451 < \includegraphics[width=\linewidth]{pComicStrip.pdf}
451 > \includegraphics[width=\linewidth]{pComicStrip}
452   \caption{\label{fig:pComic} The prismatic interfaces.  Panel
453    descriptions match those in figure \ref{fig:bComic}}
454   \end{figure}
455  
456 < From the fits using equation \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$0.2 \AA\ , respectively, with no applied momentum flux. Over
456 > From the fits using eq. \eqref{tet_fit}, we find the interfacial
457 > width for the basal and prismatic systems to be 3.2~$\pm$~0.4~\AA\ and
458 > 3.6~$\pm$~0.2~\AA\ , respectively, with no applied momentum flux. Over
459   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
# Line 430 | Line 464 | error bars.
464   over all shear rates contained the values reported above within their
465   error bars.
466  
467 < \subsubsection{Orientational Time Correlation Function}
467 > \subsubsection{Orientational Dynamics}
468   The orientational time correlation function,
469   \begin{equation}\label{C(t)1}
470    C_{2}(t)=\langle P_{2}(\mathbf{u}(0) \cdot \mathbf{u}(t)) \rangle,
# Line 453 | Line 487 | then divided into 30 bins and $C_2(t)$ was evaluated f
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 < then divided into 30 bins and $C_2(t)$ was evaluated for each bin.
490 > then divided into 30 bins along the $z$-axis and $C_2(t)$ was
491 > evaluated for each bin.
492  
493   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 < correspond to a fast (<200 fs) reorientational piece driven by the
496 > correspond to a fast ($<$200 fs) reorientational piece driven by the
497   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
# Line 473 | Line 508 | time domains, as well as a constant piece that account
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 < time domains, as well as a constant piece that accounts for the water
511 > time domains, as well as a constant piece to account for the water
512   stuck in the ice phase that does not experience any long-time
513   orientational decay,
514   \begin{equation}
# Line 487 | Line 522 | slab.
522   slab.
523  
524   \begin{figure}
525 < \includegraphics[width=\linewidth]{basal_Tau_comic_strip.pdf}
525 > \includegraphics[width=\linewidth]{basal_Tau_comic_strip}
526   \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 <  expected under reorganizations that inolve large motions of the
532 >  expected under reorganizations that involve large motions of the
533    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
# Line 502 | Line 537 | slab.
537   \end{figure}
538  
539   \begin{figure}
540 < \includegraphics[width=\linewidth]{prismatic_Tau_comic_strip.pdf}
540 > \includegraphics[width=\linewidth]{prismatic_Tau_comic_strip}
541   \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}.}
# Line 532 | Line 567 | One remarkable feature, is that for each of the interf
567   \ref{fig:prismatic_Tau_comic_strip}, while those obtained while a
568   shear was active are shown in red.
569  
570 < One remarkable feature, is that for each of the interfaces, there is
571 < an apparent fixed liquid-state value for $\tau_{short}$,
570 > Two notable features deserve clarification.  First, there are
571 > nearly-constant liquid-state values for $\tau_{short}$,
572   $\tau_{middle}$, and $\tau_{long}$ at large displacements from the
573 < interface. There also appears to be a single distance, $d_{basal}$ or
574 < $d_{prismatic}$, from the interface at which all three decay times
575 < begin to deviate from their bulk liquid values. We find $d_{basal}$
576 < and $d_{prismatic}$ to be roughly 15 \AA\ and 8 \AA\ respectively.
577 < These two results indicate that the dynamics of the water molecules
578 < within $d_{basal}$ and $d_{prismatic}$ are being significantly
579 < perturbed by the interface, even though the structural width of the
545 < interface via analysis of the tetrahedrality profile indicates that
546 < bulk liquid structure of water is recovered after about 4 \AA\ from
547 < the edge of the ice.
573 > 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 > distances to be approximately 15~\AA\ and 8~\AA\, respectively,
577 > 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  
581 < Beaglehole and Wilson have measured the ice/water interface to have a
582 < thickness approximately 10 \AA\ for both the basal and prismatic face
583 < of ice by ellipticity measurements \cite{Beaglehole93}. Structurally,
584 < we 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 the
586 < spatial dependence of the decay times of $C_2(t)$ appears to indicate
587 < that a somewhat thicker interfacial region exists in which the
588 < orientational dynamics of the water molecules begin to resemble the
589 < trapped interfacial water more than the surrounding bulk.
581 > Beaglehole and Wilson have measured the ice/water interface using
582 > ellipsometry and find a thickness of approximately 10~\AA\ for both
583 > the basal and prismatic faces.\cite{Beaglehole93} Structurally, we
584 > 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  
591 + 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   \subsection{Coefficient of Friction of the Interface}
599 < As the ice is sheared through the liquid, there will be a friction
600 < between the solid and the liquid. Pit has shown how to calculate the
601 < coefficient of friction $\lambda$ for a solid-liquid interface for a
602 < Newtonian fluid of viscosity $\eta$ and has a slip length of
603 < $\delta$. \cite{Pit99}
599 > 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 > $\left(\delta_\mathrm{wall}\right)$ of this boundary layer and the
604 > ``hydrodynamic position'' of the boundary
605 > $\left(z_\mathrm{wall}\right)$.\cite{PhysRevLett.70.2726,PhysRevE.49.3079}
606 > This last parameter is the location (relative to a solid surface)
607 > where the bulk-like behavior is recovered.  Work by Mundy {\it et al.}
608 > has helped to combine these parameters into a liquid-solid friction
609 > coefficient, which quantifies the resistance to pulling the solid
610 > interface through a liquid,\cite{Mundy1997305}
611 > \begin{equation}
612 > \lambda_\mathrm{wall} = \frac{\eta}{\delta_\mathrm{wall}}.
613 > \end{equation}
614 > 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   \begin{equation}\label{Pit}
617 < \lambda=\eta/\delta
617 >  \lambda=\frac{\eta}{\delta}
618   \end{equation}
619 < From linear response theory, $\eta$ can be obtained from the imposed
620 < momentum flux and the slope of the velocity about the dimension of the
621 < imposed flux.\cite{Kuang12}
619 > 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 > far from the surface.\cite{Kuang12} Assuming linear response in the
626 > bulk-like region,
627   \begin{equation}\label{Kuang}
628 < j_{z}(p_{x})=-\eta\frac{\partial v_{x}}{\partial z}
628 > j_{z}(p_{x})=-\eta \left(\frac{\partial v_{x}}{\partial z}\right)
629   \end{equation}
630 < Solving eq. \eqref{Kuang} for $\eta$ and substituting the result into
631 < eq. \eqref{Pit}, we obtain an alternate expression for the coefficient
576 < of friction.
630 > Substituting this result into eq. \eqref{Pit}, we can estimate the
631 > solid-liquid coefficient using the slip length,
632   \begin{equation}
633 < \lambda=-\frac{j_{z}(p_{x})}{\delta \frac{\partial v_{x}}{\partial z}}
633 > \lambda=-\frac{j_{z}(p_{x})} {\left(\frac{\partial v_{x}}{\partial
634 >      z}\right) \delta}
635   \end{equation}
636  
637 < For our simulations, we obtain $\delta$ from the difference between the structural edge of the ice block determined by the tetrahedrality profile fit, and the intersection of the linear regression of the $v_{x}$ profiles about the $z$-dimension for the ice and liquid. (See Figure \ref{fig:delta_example}) The coefficient of friction for the basal and the prismatic facets were found to be (0.07$\pm$0.01) amu fs\textsuperscript{-1} and (0.032$\pm$0.007) amu fs\textsuperscript{-1}. It is known that the basal and prismatic faces have different surface structures. The basal face is smoother than the prismatic with small alternating valleys and crests, while the prismatic surface has deep corrugating channels. We believe the reason that the prismatic face's coefficient of friction was found to be smaller than the basal's is due to the direction of the shear. The shear of the ice/water was in the same direction of the corrugating channels, allowing water molecules to pass through the channels during the shear.
637 > For ice / water interfaces, the boundary conditions are markedly
638 > no-slip, so projecting the bulk liquid state velocity profile yields a
639 > negative slip length. This length is the difference between the
640 > structural edge of the ice (determined by the tetrahedrality profile)
641 > and the location where the projected velocity of the bulk liquid
642 > intersects the solid phase velocity (see Figure
643 > \ref{fig:delta_example}). The coefficients of friction for the basal
644 > and the prismatic facets are found to be
645 > 0.07~$\pm$~0.01~amu~fs\textsuperscript{-1} and
646 > 0.032~$\pm$~0.007~amu~fs\textsuperscript{-1}, respectively. These
647 > results may seem surprising as the basal face is smoother than the
648 > prismatic with only small undulations of the oxygen positions, while
649 > the prismatic surface has deep corrugated channels. The applied
650 > momentum flux used in our simulations is parallel to these channels,
651 > however, and this results in a flow of water in the same direction as
652 > the corrugations, allowing water molecules to pass through the
653 > channels during the shear.
654  
655   \begin{figure}
656 < \includegraphics[width=\linewidth]{delta_example.pdf}
657 < \caption{\label{fig:delta_example} A schematic of determining the slip length ($\delta$). The slip length is the difference of the structural starting point of the ice and the point of intersection of the linear regressions of the liquid phase velocity profile (red) and of the solid ice velocity profile (black). The dotted line indicates the location of the ice as determined by the tetrahedrality profile.}
656 > \includegraphics[width=\linewidth]{delta_example}
657 > \caption{\label{fig:delta_example} Determining the (negative) slip
658 >  length ($\delta$) for the ice-water interfaces (which have decidedly
659 >  non-slip behavior).  This length is the difference between the
660 >  structural edge of the ice (determined by the tetrahedrality
661 >  profile) and the location where the projected velocity of the bulk
662 >  liquid (dashed red line) intersects the solid phase velocity (solid
663 >  black line).  The dotted line indicates the location of the ice as
664 >  determined by the tetrahedrality profile.}
665   \end{figure}
666  
667  
668   \section{Conclusion}
669 < Here we have simulated the basal and prismatic facets of an SPC/E model of the ice I$_\mathrm{h}$ / water interface. Using VSS-RNEMD, the ice was sheared relative to the liquid while imposed thermal gradients kept the interface at a stable temperature. Caculation of the local tetrahedrality order parameter has shown an apparent independence of the shear rate on the interfacial width, which was found to be 3.2$\pm$0.4 \AA\ and 3.6$\pm$0.2 \AA\ for the basal and prismatic systems. The orientational time correlation function was calculated from varying displacements from the interface. Decomposition by a triexponential decay also showed an apparent independence of the shear rate. The short time decay due to the restoring forces of existing hydrogen bonds decreased at close displacements from the interface, while the middle and long time decays were found to increase. There is also an apparent displacement, $d_{basal}$ and $d_{prismatic}$, from the interface at which these deviations from bulk liquid values occurs. We found $d_{basal}$ and $d_{prismatic}$ to be approximately 15 \AA\ and 8 \AA\ . This implies that the dynamics of water molecules which are structurally equivalent to bulk phase molecules are being perturbed by the presence of the ice and/or the interface. The coefficient of friction of each of the facets was also determined. They were found to be (0.07$\pm$0.01) amu fs\textsuperscript{-1} and (0.032$\pm$0.007) amu fs\textsuperscript{-1} for the basal and prismatic facets respectively. We believe the large difference between the two friction coefficients is due to the direction of the shear and the surface structure of the crystal facets.
669 > We have simulated the basal and prismatic facets of an SPC/E model of
670 > the ice I$_\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 > 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  
678 < \begin{acknowledgement}
678 > 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 > deviations from bulk liquid values begin.  We found $d_{basal}$ and
686 > $d_{prismatic}$ to be approximately 15~\AA\ and 8~\AA\ . This implies
687 > 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 >
691 > The coefficient of liquid-solid friction for each of the facets was
692 > 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 >
699 > \section{Acknowledgements}
700    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.
597 \end{acknowledgement}
704  
705   \newpage
600 \bibstyle{achemso}
706   \bibliography{iceWater}
707  
708 < \begin{tocentry}
604 < \begin{wrapfigure}{l}{0.5\textwidth}
605 < \begin{center}
606 < \includegraphics[width=\linewidth]{SystemImage.png}
607 < \end{center}
608 < \end{wrapfigure}
609 < An image of our system.
610 < \end{tocentry}
708 > \end{doublespace}
709  
710 + % \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   \end{document}
725  
726   %**************************************************************

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines