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 3918 by gezelter, Mon Jul 15 15:07:36 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 < \title{Do the facets of ice $I_\mathrm{h}$ crystals have different
15 <  friction coefficients?  Simulating shear in ice/water interfaces}
44 > \begin{document}
45  
46 < \author{P. B. Louden}
47 < \author{J. Daniel Gezelter}
19 < \email{gezelter@nd.edu}
20 < \affiliation[University of Notre Dame]{251 Nieuwland Science Hall\\
21 <  Department of Chemistry and Biochemistry\\ University of Notre
22 <  Dame\\ Notre Dame, Indiana 46556}
46 > \title{Simulations of solid-liquid friction at ice-I$_\mathrm{h}$ /
47 >  water interfaces}
48  
49 < \keywords{}
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 < \begin{document}
56 > \date{\today}
57 > \maketitle
58 > \begin{doublespace}
59  
60 +
61   \begin{abstract}
62 < We have investigated the structural properties of the basal and
63 < prismatic facets of an SPC/E model of the ice Ih / water interface
64 < when the solid phase is being drawn through liquid water (i.e. sheared
65 < relative to the fluid phase). To impose the shear, we utilized a
66 < reverse non-equilibrium molecular dynamics (RNEMD) method that creates
67 < non-equilibrium conditions using velocity shearing and scaling (VSS)
68 < moves of the molecules in two physically separated slabs in the
69 < simulation cell. This method can create simultaneous temperature and
70 < velocity gradients and allow the measurement of friction transport
71 < properties at interfaces. We present calculations of the interfacial
72 < friction coefficients and the apparent independence of shear rate on
73 < interfacial width and show that water moving over a flat ice/water
74 < interface is close to the no-slip boundary condition.
62 >  We have investigated the structural and dynamic properties of the
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 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
82  
83   \section{Introduction}
84 + %-----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  
91 < %Other people looking at the ice/water interface
49 < %Geologists are concerned with the flow of water over ice
50 < %Antifreeze protein in fish--Haymet's group has cited this before
91 > %Gay02: cites many other ice/water papers, make sure to cite them.
92  
93 < %Paragraph explaining why the ice/water interface is important
94 < %Paragraph on what other people have done / lead into what hasn't been done
95 < %Paragraph on what I'm going to do
93 > 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 > 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.}
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}
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
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  
126 + 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  
58
59 With the recent development of velocity shearing and scaling reverse non-equilibrium molecular dynamics (VSS-RNEMD), it is now possible to calculate transport properties from heterogeneous systems.\cite{Kuang12} This method can create simultaneous temperature and velocity gradients and allow the measurement of friction and thermal transport properties at interfaces. This allows for the study of the width of the ice/water interface as the ice is sheared through the liquid, while imposing a thermal gradient to prevent frictional heating of the interface.
60
61 as well as determining the friction coefficient of the interface.
62
63 In this paper, we investigate the width and the friction coefficient of the ice/water interface as the ice is sheared through the liquid.
64
65
66
142   \section{Methodology}
143  
144 < \subsection{Stable ice I$_\mathrm{h}$ / water interfaces}
144 > \subsection{Stable ice I$_\mathrm{h}$ / water interfaces under shear}
145  
146   The structure of ice I$_\mathrm{h}$ is well understood; it
147   crystallizes in a hexagonal space group P$6_3/mmc$, and the hexagonal
# Line 74 | Line 149 | ice I$_\mathrm{h}$ are the secondary prism face, $\{1~
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 < ice I$_\mathrm{h}$ are the secondary prism face, $\{1~1~\bar{2}~0\}$,
153 < and the prismatic face, $\{2~0~\bar{2}~1\}$.
152 > ice I$_\mathrm{h}$ are the secondary prism, $\{1~1~\bar{2}~0\}$, and
153 > 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  
158 < Ice I$_\mathrm{h}$ is normally proton disordered in bulk crystals,
159 < although the surfaces probably have a preference for proton ordering
160 < along strips of dangling H-atoms and Oxygen lone
161 < pairs.\cite{Buch:2008fk}
162 <
163 < For small simulated ice interfaces, it is useful to have a
86 < proton-ordered, but zero-dipole crystal that exposes these strips of
87 < dangling H-atoms and lone pairs.  Also, if we're going to place
88 < another material in contact with one of the ice crystalline planes, it
89 < is useful to have an orthorhombic (rectangular) box to work with.  A
90 < recent paper by Hirsch and Ojam\"{a}e describes how to create
91 < proton-ordered bulk ice I$_\mathrm{h}$ in alternative orthorhombic
92 < cells.\cite{Hirsch04}
93 <
94 < We have using Hirsch and Ojam\"{a}e's structure 6 which is an
95 < orthorhombic cell ($P2_12_12_1$) that produces a proton-ordered
96 < version of ice Ih.  Table \ref{tab:equiv} contains a mapping between
97 < the Miller indices in the P$6_3/mmc$ crystal system and those in the
98 < Hirsch and Ojam\"{a}e $P2_12_12_1$ system.
99 <
100 < \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 \bibpunct{}{}{,}{n}{}{,} \protect\citep{Hirsch04}.}
163 > \label{tab:equiv}
164   \begin{tabular}{|ccc|} \hline
165   & hexagonal & orthorhombic \\
166   & ($P6_3/mmc$) & ($P2_12_12_1$) \\
# Line 105 | Line 168 | pryamidal & $\{2~0~\bar{2}~1\}$ & $\{2~0~1\}$ \\ \hlin
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 < pryamidal & $\{2~0~\bar{2}~1\}$ & $\{2~0~1\}$ \\ \hline
171 > pyramidal & $\{2~0~\bar{2}~1\}$ & $\{2~0~1\}$ \\ \hline
172   \end{tabular}
173 < \end{wraptable}
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
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   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 the following fractional coordinates:
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
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 $y$
205 > dimensions until a desired sample size was obtained.
206  
207 < \begin{wraptable}{r}{3.25in}
208 < \begin{tabular}{|ccccc|}  \hline
209 < atom label & type & x & y & z \\ \hline
210 < O$_{a}$    & O & 0.75 & 0.1667 & 0.4375 \\
211 < H$_{1a}$ & H & 0.5735 & 0.2202 & 0.4836 \\
212 < H$_{2a}$ & H & 0.7420 & 0.0517 & 0.4836 \\
213 < O$_{b}$    & O & 0.25 & 0.6667 & 0.4375 \\
214 < H$_{1b}$ & H & 0.2580 & 0.6693 & 0.3071 \\
215 < H$_{2b}$ & H & 0.4265 & 0.7255 & 0.4756 \\ \hline
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 \bibpunct{}{}{,}{n}{}{,} \protect\citep{Hirsch04}.}
211 > \label{tab:p212121}
212 > \begin{tabular}{|cccc|}  \hline
213 > atom type & x & y & z \\ \hline
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 & 0.2500 & 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 < To construct the basal and prismatic interfaces, the crystallographic
224 < coordinates above were used to construct an orthorhombic unit cell
225 < which was then replicated in all three dimensions yielding a
226 < proton-ordered block of ice I$_{h}$. To expose the desired face, the
227 < orthorhombic representation was then cut along the ($001$) or ($100$)
228 < planes for the basal and prismatic faces respectively.  The resulting
229 < block was rotated so that the exposed faces were aligned with the $z$
230 < axis normal to the exposed face.  The block was then cut along two
231 < perpendicular directions in a way that allowed for perfect periodic
232 < replication in the $x$ and $y$ axes, creating a slab with either the
233 < basal or prismatic faces exposed along the $z$ axis.  The slab was
234 < then replicated in the $x$ and $y$ dimensions until a desired sample
235 < size was obtained.  
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 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$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. 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  
142 Although experimental solid/liquid coexistant temperature under normal
143 pressure are close to 273K, Haymet \emph{et al.} have done extensive
144 work on characterizing the ice/water
145 interface.\cite{Karim88,Karim90,Hayward01,Bryk02,Hayward02} They have
146 found for the SPC/E water model,\cite{Berendsen87} which is also used
147 in this study, the ice/water interface is most stable at
148 225$\pm$5K.\cite{Bryk02} To create a ice / water interface, a box of
149 liquid water that had the same dimensions in $x$ and $y$ was
150 equilibrated at 225 K and 1 atm of pressure in the NPAT ensemble (with
151 the $z$ axis allowed to fluctuate to equilibrate to the correct
152 pressure).  The liquid and solid systems were combined by carving out
153 any water molecule from the liquid simulation cell that was within 3
154 \AA\ of any atom in the ice slab.
155
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 < data collection was carried out.
242 > data collection was carried out.  
243  
244   \subsection{Shearing ice / water interfaces without bulk melting}
245  
246 < As one drags a solid through a liquid, there will be frictional
247 < heating that will act to melt the interface.  To study the frictional
248 < behavior of the interface without causing the interface to melt, it is
249 < necessary to apply a weak thermal gradient along with the momentum
250 < gradient.  This can be accomplished with of the newly-developed
251 < approaches to reverse non-equilibrium molecular dynamics (RNEMD).  The
252 < velocity shearing and scaling (VSS) variant of RNEMD utilizes a series
253 < of simultaneous velocity exchanges between two regions within the
254 < simulation cell.\cite{Kuang12} One of these regions is centered within
255 < the ice slab, while the other is centrally located in the liquid phase
246 > 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 > it is necessary to apply a weak thermal gradient in combination with
250 > the momentum gradient.  This can be accomplished using the velocity
251 > 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 > slab, while the other is centrally located in the liquid
256   region. VSS-RNEMD provides a set of conservation constraints for
257 < simultaneously creating either a momentum flux or a thermal flux (or
258 < both) between the two slabs.  Satisfying the constraint equations
259 < ensures that the new configurations are sampled from the same NVE
260 < ensemble as previously.
257 > 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  
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
# Line 229 | Line 312 | response to the applied flux.  In a bulk material it i
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 < response to the applied flux.  In a bulk material it is quite simple
315 > response to the applied flux.  In a bulk material, it is quite simple
316   to use the slope of the temperature or velocity gradients to obtain
317 < the thermal conductivity or shear viscosity.
317 > either the thermal conductivity or shear viscosity.
318  
319   The VSS-RNEMD approach is versatile in that it may be used to
320   implement thermal and shear transport simultaneously.  Perturbations
# Line 239 | 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 < Here we are using this method primarily to generate a shear between
328 < the ice slab and the liquid phase, while using a weak thermal gradient
329 < to maintaining the interface at the 225K target value.  This will
330 < insure minimal melting of the bulk ice phase and allows us to control
331 < the exact temperature of the interface.
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 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  
333   \subsection{Computational Details}
334 < All simulations were performed using OpenMD with a time step of 2 fs,
335 < and periodic boundary conditions in all three dimensions. The systems
336 < were divided into 100 artificial bins along the $z$-axis for the
337 < VSS-RNEMD moves, which were attempted every 50 fs. The gradients were
338 < allowed to develop for 1 ns before data collection was began. Once
339 < established, snapshots of the system were taken every 1 ps, and the
257 < average velocities and densities of each bin were accumulated every
258 < attempted VSS-RNEMD move.
334 > 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 > which were attempted every 50~fs.
340  
341 < %A paragraph on the equilibration procedure of the system? Shenyu did some amount of equilibration to the files and then I was handed them. I performed 5 ns of NVT at 225K for both systems, then 5 ns of NVE at 225K for both systems, with no gradients imposed.
342 < %For the basal, once the thermal gradient was found which gave me the interfacial temperature I wanted (-2.0E-6 kcal/mol/A^2/fs), I equilibrated the file for 5 ns letting this gradient stabilize. Then I continued to use this thermal gradient as I imposed momentum gradients and watched the response of the interface.
343 < %For the prismatic, a gradient was not found that would give me the interfacial temperature I desired, so while imposing a thermal gradient that had the interface at 220K, I raised the temperature of the system to 230K. This resulted in a thermal gradient which gave my interface at 225K, equilibrated for ins NVT, then ins NVE while this gradient was still imposed, then I began dragging.
344 < %I have run each system for 1 ns under PTgrads to allow them to develop, then ran each system for an additional 2 ns in segments of 0.5 ns in order to calculate statistics of the calculated values.
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 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
350 > temperature at the interface within $\pm 1$ K of the 225K target.
351  
352 < \subsection{Measuring the Width of the Interface}
353 < In order to characterize the ice/water interface, the local
354 < tetrahedral order parameter as described by Kumar\cite{Kumar09} and
355 < Errington\cite{Errington01} was used. The local tetrahedral order
356 < parameter, $q$, is given by
357 < \begin{equation}
358 < q_{k} \equiv 1 -\frac{3}{8}\sum_{i=1}^{3} \sum_{j=i+1}^{4} \Bigg[\cos\psi_{ikj}+\frac{1}{3}\Bigg]^2
272 < \end{equation}
273 < where $\psi_{ikj}$ is the angle formed by the oxygen sites on molecule $k$, and the oxygen site on its two closest neighbors, molecules $i$ and $j$. The local tetrahedral order parameter function has a range of (0,1), where the larger the value $q$ has the more tetrahedral the ordering of the local environment is. A $q$ value of one describes a perfectly tetrahedral environment relative to it and its four nearest neighbors, and the parameter's value decreases as the local ordering becomes less tetrahedral.
274 <
275 < %If the central water molecule has a perfect tetrahedral geometry with its four nearest neighbors, the parameter goes to one, and decreases to zero as the geometry deviates from the ideal configuration.
276 <
277 < The system was divided into 100 bins along the $z$-axis, and a $q$ value was determined for each snapshot of the system for each bin. The $q$ values for each bin were then averaged to give an average tetrahedrally profile of the system about the $z$- axis. The profile was then fit with a hyperbolic tangent function given by
278 < \begin{equation}
279 < q_{z}=q_{liq}+\frac{q_{ice}-q_{liq}}{2}(\tanh(\alpha(z-I_{L,m}))-\tanh(\alpha(z-I_{R,m})))
280 < \end{equation}
281 < where $q_{liq}$ and $q_{ice}$ are fitting parameters for the local tetrahedral order parameter for the liquid and ice, $\alpha$ is proportional to the inverse of the width of the interface, and $I_{L,m}$ and $I_{R,m}$ are the midpoints of the left and right interface. During the simulations where a kinetic energy flux was imposed, there was found to be a thermal influence in the liquid phase region of the tetrahedrally profile due to the thermal gradient developed in the system. To maximize the fit of the interface, another term was added to the hyperbolic tangent fitting function,
282 < \begin{equation}
283 < q_{z}=q_{liq}+\frac{q_{ice}-q_{liq}}{2}(\tanh(\alpha(z-I_{L,m}))-\tanh(\alpha(z-I_{R,m})))+\beta|(z-z_{mid})|
284 < \end{equation}
285 < where $\beta$ is a fitting parameter and $z_{mid}$ is the midpoint of the z dimension of the simulation box.
286 <
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
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
358 > were accumulated throughout the simulations.
359  
360   \section{Results and discussion}
361  
362 < %Images to include: 3-long comic strip style of <Vx>, T, q_z as a function of z for the basal and prismatic faces. q_z by z with fit for basal and prismatic. interface width as a function of deltaVx (shear rate) with basal and prismatic on the same plot, error bars in the x and y. <Vx> by flux with basal and prismatic on same graph, back out slope from xmgr and error in slope to get lambda, friction coefficient of interface.
362 > \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 > 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 > 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 tetrahedral order parameter as described by
375 > Kumar\cite{Kumar09} and Errington\cite{Errington01} as our principal
376 > measure of the interfacial width.
377  
378 < In Figures \ref{fig:bComic} and \ref{fig:pComic} we see the $z$-dimensional profiles for several parameters for the basal and prismatic systems respectively. In panel (a) of the figures we see the tetrahedrality profile of the systems (black circles). In the liquid region of the system, the local tetrahedral order parameter is approximately 0.75. In the solid region, the parameter is approximately 0.94, indicating a more tetrahedral structure of the water molecules. The hyperbolic tangent function used to fit the tetrahedrality profiles can be found in red and the verticle dotted lines denote the midpoint of the interfaces. The weak thermal gradient applied to the systems in order to keep the interface at a stable temperature, 225$\pm$5K, can be seen in panel (b).  Lastly, the velocity gradient across the systems can be seen in panel (c). From panel (c), we can see liquid phase water molecules 5 to 12 \AA\ from the midpoint of the basal and prismatic interfaces are being dragged along with the ice block. This indicates that the shearing of ice water is in the stick boundary condition.  
378 > The local tetrahedral order parameter, $q(z)$, is given by
379 > \begin{equation}
380 > 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 > \label{eq:qz}
384 > \end{equation}
385 > where $\psi_{ikj}$ is the angle formed between the oxygen site on
386 > 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  
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
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,
404 + \begin{equation}\label{tet_fit}
405 + 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 + \end{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 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{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
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 + 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 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
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}
440 < \caption{\label{fig:bComic} The basal system: (a) The local tetrahedral order parameter,$q$, (black circles) fit by a hyperbolic tangent (red line), (b) the thermal gradient imposed on the system to maintain a stable interfacial temperature, and (c) the velocity gradient imposed on the system. The verticle dotted lines indicate the midpoint of the interfaces.}
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
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   \end{figure}
449  
299 %(a) The local tetrahedral order parameter across the z-dimension of the system (black circles) fit by a hyperbolic tangent (red line). (b) The thermal gradient imposed on the system to maintain a stable interfacial temperature. (c) The velocity gradient imposed on the system. The verticle dotted lines indicate the midpoint of the interfaces.
300
450   \begin{figure}
451   \includegraphics[width=\linewidth]{pComicStrip}
452 < \caption{\label{fig:pComic} The prismatic system: (a) The local tetrahedral order parameter,$q$, (black circles) fit by a hyperbolic tangent (red line), (b) the thermal gradient imposed on the system to maintain a stable interfacial temperature, and (c) the velocity gradient imposed on the system.The verticle dotted lines indicate the midpoint of the interfaces.}
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 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 + $0.9 \pm 0.2 \mathrm{ms}^{-1} \rightarrow 4.5 \pm 0.1
462 + \mathrm{ms}^{-1}$ 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  
467 < \subsection{Interfacial Width}
468 < %For the basal and prismatic systems, the ice blocks were sheared through the water at varying rates while an imposed thermal gradient kept the interface at the stable temperature range as described by Byrk and Haymet.
469 < We found the interfacial width for the basal and prismatic systems to be 3.2$\pm$0.4 \AA\ and 3.6$\pm$0.2 \AA\ with no applied momentum flux. Over the range of shear rates investigated, 0.6$\pm$0.3 ms\textsuperscript{-1} to 5.3$\pm$0.5 ms\textsuperscript{-1} for the basal system and 0.9$\pm$0.2 ms\textsuperscript{-1} to 4.5$\pm$0.1 ms\textsuperscript{-1}, there was no appreciable change in the interface width found. The calculated values for the interfacial width over all shear rates investigated contained the control values within their error bars.
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,
471 > \end{equation}
472 > 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 > the vector $\mathbf{u}$ is often taken as HOH bisector, although
477 > slightly different behavior can be observed when $\mathbf{u}$ is the
478 > 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  
481 + 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 + 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 + 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
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
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 + 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}
515 + 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 + \end{equation}
518 + 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 +
524 + \begin{figure}
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 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
536 +  interface.}
537 + \end{figure}
538 +
539 + \begin{figure}
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}.}
544 + \end{figure}
545 +
546 + 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 +
565 + 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 +
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. 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 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 between the ice and the interface. Balasubramanian has shown how to calculate the coefficient of friction for a solid-liquid interface. \cite{Balasubramanian99}
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 < %<F_{x}^{w}>_{NE}(t)=-S\lambda_{wall}v_{x}(y_{wall})
315 < \langle F_{x}^{w}\rangle(t)=-S\lambda_{wall}v_{x}(y_{wall})
612 > \lambda_\mathrm{wall} = \frac{\eta}{\delta_\mathrm{wall}}.
613   \end{equation}
614 < In this equation, $F_{x}^{w}$ is the total force of all the atoms acting on the fluid, $S$ is the surface area the force is being applied upon, and $\lambda_{wall}$ is the coefficient of friction of the interface. Since the imposed momentum flux, $J_{z}(p_{x})$, is known in the VSS-RNEMD simulations, and the $wall$ is the ice block in our simulations, the above equation can be rewritten as
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=\frac{\eta}{\delta}
618 > \end{equation}
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 \left(\frac{\partial v_{x}}{\partial z}\right)
629 > \end{equation}
630 > Substituting this result into eq. \eqref{Pit}, we can estimate the
631 > solid-liquid coefficient using the slip length,
632   \begin{equation}
633 < J_{z}(p_{x})=-\lambda_{ice}v_{x}(y_{ice}).
633 > \lambda=-\frac{j_{z}(p_{x})} {\left(\frac{\partial v_{x}}{\partial
634 >      z}\right) \delta}
635   \end{equation}
636  
637 < In Figure \ref{fig:CoeffFric}, the average velocity of the ice is plotted against the imposed momentum flux for the basal (black circles) and prismatic (red circles) systems. From the equation above, the slope of the linear fit of the data is $\lambda_{wall}$. The coefficient of friction of the interface for the basal face was calculated to be 11.0 $\pm$ 0.4 \AA\textsuperscript{-2}fs\textsuperscript{-1}, and the $\lambda_{wall}$ for the prismatic face was determined to be 19.9, $\pm$ 0.5 \AA\textsuperscript{-2}fs\textsuperscript{-1}.  
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 < %Ask dan about truncating versus rounding the values for lambda.
656 < %The coefficient of friction of the interface for the basal face was calculated to be 11.02808 $\pm$  0.4489844 \AA^{-2}fs^{-1}, and the $\lambda_{wall}$ for the prismatic face was determined to be 19.95948, $\pm$ 0.5370894 \AA^{-2}fs^{-1}.
657 < \begin{figure}
658 < \includegraphics[width=\linewidth]{CoeffFric}
659 < \caption{\label{fig:CoeffFric} The average velocity of the ice for the basal (black circles) and prismatic (red circles) systems as a function off applied momentum flux. The slope of the fit line   }
655 > \begin{figure}
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 Ih / water interface. Using VSS-RNEMD, the ice was sheared relative to the liquid while imposed thermal gradients kept the interface at a stable temperature. Caculation of the local tetrahedrality order parameter has shown an appearant independence of the shear rate on the interfacial width. The coefficient of friction of the interface was also calculated for each of the facets. The $\lambda_{wall}$ for the basal face was calculated to be 11.0 $\pm$ 0.4 \AA\textsuperscript{-2}fs\textsuperscript{-1}, and 19.9, $\pm$ 0.5 \AA\textsuperscript{-2}fs\textsuperscript{-1} for the prismatic facet. For both facets, the shearing ice water was found to be in the no-slip boundary condition.  
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 + 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 < \begin{acknowledgement}
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.
340 \end{acknowledgement}
704  
705   \newpage
343 \bibstyle{achemso}
706   \bibliography{iceWater}
707  
708 < \begin{tocentry}
347 < \begin{wrapfigure}{l}{0.5\textwidth}
348 < \begin{center}
349 < \includegraphics[width=\linewidth]{SystemImage.png}
350 < \end{center}
351 < \end{wrapfigure}
352 < An image of our system.
353 < \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 < % basal: slope=11.02808, error in slope = 0.4489844
727 < %prismatic: slope = 19.95948, error in slope  = 0.5370894
726 > %**************************************************************
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 > %**************************************************************

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines