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 3944 by plouden, Fri Aug 30 18:13:50 2013 UTC vs.
Revision 3945 by gezelter, Tue Sep 3 21:16:58 2013 UTC

# Line 11 | Line 11
11   \usepackage{url}
12  
13  
14 < \title{Do the facets of ice $I_\mathrm{h}$ crystals have different
15 <  friction coefficients?  Simulating shear in ice/water interfaces}
14 > \title{Solid-liquid friction at ice-I$_\mathrm{h}$ / water interfaces}
15  
16   \author{P. B. Louden}
17   \author{J. Daniel Gezelter}
# Line 26 | Line 25 | We have investigated the structural properties of the
25   \begin{document}
26  
27   \begin{abstract}
28 < We have investigated the structural properties of the basal and
29 < prismatic facets of an SPC/E model of the ice Ih / water interface
30 < when the solid phase is being drawn through liquid water (i.e. sheared
31 < relative to the fluid phase). To impose the shear, we utilized a
32 < reverse non-equilibrium molecular dynamics (RNEMD) method that creates
33 < non-equilibrium conditions using velocity shearing and scaling (VSS)
34 < moves of the molecules in two physically separated slabs in the
35 < simulation cell. This method can create simultaneous temperature and
36 < velocity gradients and allow the measurement of friction transport
37 < properties at interfaces. We found the interfacial width for the basal and prismatic facets to be 3.2$\pm$0.4 \AA\ and 3.6$\pm$0.2 \AA\ with no momentum gradient present. The interfacial width was found to be independent of shear rate over the range of shear rates investigated, 0.6$\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} for the prismatic. The orientational time correlation function was evaluated at varying displacements from the interface, and fit to a triexponential decay with decay times $\tau_{short}$, $\tau_{middle}$, and $\tau_{long}$. We found that each decay time was independent of the shear rate, and that they each deviated from a bulk liquid value at the same displacement $d_{basal}$ and $d_{prismatic}$. We determined $d_{basal}$ and $d_{prismatic}$ to be approximately 15 \AA\ and 8 \AA\ . Both $\tau_{middle}$ and $\tau_{long}$ increased in value in decreasing displacements from $d_{basal}$ and $d_{prismatic}$, while we found that $\tau_{short}$ had the inverse correlation. Lastly we present calculations of the friction coefficients ($\lambda$) for the basal and prismatic facets. We have found $\lambda_{basal}$ to be (0.07$\pm$0.01) amu fs\textsuperscript{-1} and $\lambda_{prismatic}$ to be  (0.032$\pm$0.007) amu fs\textsuperscript{-1}.
28 >  We have investigated the structural and dynamic properties of the
29 >  basal and prismatic facets of an ice I$_\mathrm{h}$ / water
30 >  interface when the solid phase is being drawn through the liquid
31 >  (i.e. sheared relative to the fluid phase). To impose the shear, we
32 >  utilized a velocity-shearing and scaling (VSS) approach to reverse
33 >  non-equilibrium molecular dynamics (RNEMD).  This method can create
34 >  simultaneous temperature and velocity gradients and allow the
35 >  measurement of transport properties at interfaces.  The interfacial
36 >  width was found to be independent of relative velocity of the ice
37 >  and liquid layers over a wide range of shear rates.  Decays of
38 >  molecular orientational time correlation functions for gave very
39 >  similar estimates for the width of the interfaces, although the
40 >  short- and longer-time decay components of the orientational
41 >  correlation functions behave differently closer to the interface.
42 >  Although both facets of ice are in ``stick'' boundary conditions in
43 >  liquid water, the solid-liquid friction coefficient was found to be
44 >  different for the basal and prismatic facets of ice.
45   \end{abstract}
46  
47   \newpage
# Line 50 | Line 56 | Understanding the ice/water interface is essential for
56  
57   %Gay02: cites many other ice/water papers, make sure to cite them.
58  
59 < Understanding the ice/water interface is essential for explaining complex processes such as nucleartion and crystal growth\cite{Han92,Granasy95,Vanfleet95}, crystal melting\cite{Weber83,Han92,Sakai96,Sakai96B}, and biological interfacial processes, such as the antifreeze protein found in winter flounder\cite{Wierzbicki07, Chapsky97}. These processes have been studied at the fundamental level of the ice/water interface by several groups, including studying the structure and width of the interface. Haymet \emph{et al.} have done extensive work on ice Ih, the most common form of ice on Earth, including characterizing and determining the width of the ice/water interface for the SPC, SPC/E, CF1, and TIP4P models for water. \cite{Karim87,Karim88,Karim90,Hayward01,Bryk02,Hayward02,Gay02} More recently, Haymet \emph{et al.} have been investigating the effects cations and anions have on crystal nucleation\cite{Bryk04,Smith05,Wilson08,Wilson10}. Nada \emph{et al.} have also studied the ice/water interface\cite{Nada95,Nada00,Nada03,Nada12}. They have found that the different facets of ice Ih have different growth rates, primarily, that the prismatic facet grows faster than the basal facet due to the mechanism of the crystal growth being the reordering of the hydrogen bonding network\cite{Nada05}.
59 > Understanding the ice/water interface is essential for explaining
60 > complex processes such as nucleation and crystal
61 > growth,\cite{Han92,Granasy95,Vanfleet95} crystal
62 > melting,\cite{Weber83,Han92,Sakai96,Sakai96B} and some fascinating
63 > biological processes, such as the behavior of the antifreeze proteins
64 > found in winter flounder,\cite{Wierzbicki07, Chapsky97} and certain
65 > terrestrial arthropods.\cite{Duman:2001qy,Meister29012013} There has
66 > been significant progress on understanding the structure and dynamics
67 > of quiescent ice/water interfaces utilizing both theory and
68 > experiment.  Haymet \emph{et al.} have done extensive work on ice I$_\mathrm{h}$,
69 > including characterizing and determining the width of the ice/water
70 > interface for the SPC, SPC/E, CF1, and TIP4P models for
71 > water.\cite{Karim87,Karim88,Karim90,Hayward01,Bryk02,Hayward02,Gay02}
72 > More recently, Haymet \emph{et al.} have investigated the effects
73 > cations and anions have on crystal
74 > nucleation\cite{Bryk04,Smith05,Wilson08,Wilson10}. Nada \emph{et al.}
75 > have also studied ice/water
76 > interfaces,\cite{Nada95,Nada00,Nada03,Nada12} and have found that the
77 > differential growth rates of the facets of ice I$_\mathrm{h}$ are due to the the
78 > reordering of the hydrogen bonding network\cite{Nada05}.
79  
80 < Another complex process which requires investigation at the ice/water interface is the movement of water over ice, such as icebergs floating in the ocean. In addition to understanding the structure and width of the interface, it is pertinent to understand the friction caused by the shearing of water across the ice to understand this process. However, until recently, simulations of this nature were not possible.
81 < 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. In this paper, we investigate the width and the friction coefficient of the ice/water interface as the ice is sheared through the liquid under a weak thermal gradient.
80 > The movement of liquid water over the facets of ice has been less
81 > thoroughly studied than the quiescent surfaces. This process is
82 > potentially important in understanding transport of large blocks of
83 > ice in water (which has important implications in the earth sciences),
84 > as well as the relative motion of crystal-crystal interfaces that have
85 > been separated by nanometer-scale fluid domains.  In addition to
86 > understanding both the structure and thickness of the interfacial
87 > regions, it is important to understand the molecular origin of
88 > friction, drag, and other changes in dynamical properties of the
89 > liquid in the regions close to the surface that are altered by the
90 > presence of a shearing of the bulk fluid relative to the solid phase.
91  
92 + In this work, we apply a recently-developed velocity shearing and
93 + scaling approach to reverse non-equilibrium molecular dynamics
94 + (VSS-RNEMD). This method makes it possible to calculate transport
95 + properties like the interfacial thermal conductance across
96 + heterogeneous interfaces,\cite{Kuang12} and can create simultaneous
97 + temperature and velocity gradients and allow the measurement of
98 + friction and thermal transport properties at interfaces.  This has
99 + allowed us to investigate the width of the ice/water interface as the
100 + ice is sheared through the liquid, while simultaneously imposing a
101 + weak thermal gradient to prevent frictional heating of the interface.
102 + In the sections that follow, we discuss the methodology for creating
103 + and simulating ice/water interfaces under shear and provide results
104 + from both structural and dynamical correlation functions.  We also
105 + show that the solid-liquid interfacial friction coefficient depends
106 + sensitively on the details of the surface morphology.
107 +
108   \section{Methodology}
109  
110 < \subsection{Stable ice I$_\mathrm{h}$ / water interfaces}
110 > \subsection{Stable ice I$_\mathrm{h}$ / water interfaces under shear}
111  
112 < The structure of ice I$_\mathrm{h}$ is well understood; it
112 > The structure of ice I$_\mathrm{h}$ is very well understood; it
113   crystallizes in a hexagonal space group P$6_3/mmc$, and the hexagonal
114   crystals of ice have two faces that are commonly exposed, the basal
115   face $\{0~0~0~1\}$, which forms the top and bottom of each hexagonal
116   plate, and the prismatic $\{1~0~\bar{1}~0\}$ face which forms the
117   sides of the plate. Other less-common, but still important, faces of
118 < ice I$_\mathrm{h}$ are the secondary prism face, $\{1~1~\bar{2}~0\}$,
119 < and the prismatic face, $\{2~0~\bar{2}~1\}$.
118 > ice I$_\mathrm{h}$ are the secondary prism, $\{1~1~\bar{2}~0\}$, and
119 > the prismatic, $\{2~0~\bar{2}~1\}$, faces.  Ice I$_\mathrm{h}$ is
120 > normally proton disordered in bulk crystals, although the surfaces
121 > probably have a preference for proton ordering along strips of
122 > dangling H-atoms and Oxygen lone pairs.\cite{Buch:2008fk}
123  
71 Ice I$_\mathrm{h}$ is normally proton disordered in bulk crystals,
72 although the surfaces probably have a preference for proton ordering
73 along strips of dangling H-atoms and Oxygen lone
74 pairs.\cite{Buch:2008fk}
75
124   For small simulated ice interfaces, it is useful to have a
125   proton-ordered, but zero-dipole crystal that exposes these strips of
126 < dangling H-atoms and lone pairs.  Also, if we're going to place
127 < another material in contact with one of the ice crystalline planes, it
128 < is useful to have an orthorhombic (rectangular) box to work with.  A
129 < recent paper by Hirsch and Ojam\"{a}e describes how to create
130 < proton-ordered bulk ice I$_\mathrm{h}$ in alternative orthorhombic
83 < cells.\cite{Hirsch04}
126 > dangling H-atoms and lone pairs.  When placing another material in
127 > contact with one of the ice crystalline planes, it is useful to have
128 > an orthorhombic (rectangular) box.  A recent paper by Hirsch and
129 > Ojam\"{a}e describes how to create proton-ordered bulk ice
130 > I$_\mathrm{h}$ in alternative orthorhombic cells.\cite{Hirsch04}
131  
132   We are using Hirsch and Ojam\"{a}e's structure 6 which is an
133   orthorhombic cell ($P2_12_12_1$) that produces a proton-ordered
134 < version of ice Ih.  Table \ref{tab:equiv} contains a mapping between
134 > version of ice I$_\mathrm{h}$.  Table \ref{tab:equiv} contains a mapping between
135   the Miller indices in the P$6_3/mmc$ crystal system and those in the
136   Hirsch and Ojam\"{a}e $P2_12_12_1$ system.
137  
138   \begin{wraptable}{r}{3.5in}
139 +  \caption{Mapping between the Miller indices of four facets of ice in
140 +    the $P6_3/mmc$ crystal system to the orthorhombic $P2_12_12_1$
141 +    system in reference \protect\cite{Hirsch04}}
142 + \label{tab:equiv}
143   \begin{tabular}{|ccc|} \hline
144   & hexagonal & orthorhombic \\
145   & ($P6_3/mmc$) & ($P2_12_12_1$) \\
# Line 102 | Line 153 | molecules whose atoms reside at the following fraction
153  
154   Structure 6 from the Hirsch and Ojam\"{a}e paper has lattice
155   parameters $a = 4.49225$, $b = 7.78080$, $c = 7.33581$ and two water
156 < molecules whose atoms reside at the following fractional coordinates:
156 > molecules whose atoms reside at fractional coordinates given in table
157 > \ref{tab:p212121}.  To construct the basal and prismatic interfaces,
158 > these crystallographic coordinates were used to construct an
159 > orthorhombic unit cell which was then replicated in all three
160 > dimensions yielding a proton-ordered block of ice I$_{h}$. To expose
161 > the desired face, the orthorhombic representation was then cut along
162 > the ($001$) or ($100$) planes for the basal and prismatic faces
163 > respectively.  The resulting block was rotated so that the exposed
164 > faces were aligned with the $z$-dimension normal to the exposed face.
165 > The block was then cut along two perpendicular directions in a way
166 > that allowed for perfect periodic replication in the $x$ and $y$ axes,
167 > creating a slab with either the basal or prismatic faces exposed along
168 > the $z$ axis.  The slab was then replicated in the $x$ and $y$
169 > dimensions until a desired sample size was obtained.
170  
171   \begin{wraptable}{r}{3.25in}
172 < \begin{tabular}{|ccccc|}  \hline
173 < atom label & type & x & y & z \\ \hline
174 < O$_{a}$    & O & 0.75 & 0.1667 & 0.4375 \\
175 < H$_{1a}$ & H & 0.5735 & 0.2202 & 0.4836 \\
176 < H$_{2a}$ & H & 0.7420 & 0.0517 & 0.4836 \\
177 < O$_{b}$    & O & 0.25 & 0.6667 & 0.4375 \\
178 < H$_{1b}$ & H & 0.2580 & 0.6693 & 0.3071 \\
179 < H$_{2b}$ & H & 0.4265 & 0.7255 & 0.4756 \\ \hline
172 >  \caption{Fractional coordinates for water in the orthorhombic
173 >    $P2_12_12_1$ system for ice I$_\mathrm{h}$ in reference
174 >    \protect\cite{Hirsch04}}
175 > \label{tab:p212121}
176 > \begin{tabular}{|cccc|}  \hline
177 > atom type & x & y & z \\ \hline
178 > O & 0.75 & 0.1667 & 0.4375 \\
179 > H & 0.5735 & 0.2202 & 0.4836 \\
180 > H & 0.7420 & 0.0517 & 0.4836 \\
181 > O & 0.25 & 0.6667 & 0.4375 \\
182 > H & 0.2580 & 0.6693 & 0.3071 \\
183 > H & 0.4265 & 0.7255 & 0.4756 \\ \hline
184   \end{tabular}
185   \end{wraptable}
186  
119 To construct the basal and prismatic interfaces, the crystallographic
120 coordinates above were used to construct an orthorhombic unit cell
121 which was then replicated in all three dimensions yielding a
122 proton-ordered block of ice I$_{h}$. To expose the desired face, the
123 orthorhombic representation was then cut along the ($001$) or ($100$)
124 planes for the basal and prismatic faces respectively.  The resulting
125 block was rotated so that the exposed faces were aligned with the $z$-dimension normal to the exposed face.  The block was then cut along two
126 perpendicular directions in a way that allowed for perfect periodic
127 replication in the $x$ and $y$ axes, creating a slab with either the
128 basal or prismatic faces exposed along the $z$ axis.  The slab was
129 then replicated in the $x$ and $y$ dimensions until a desired sample
130 size was obtained.  
131
187   Although experimental solid/liquid coexistant temperature under normal
188   pressure are close to 273K, Haymet \emph{et al.} have done extensive
189   work on characterizing the ice/water
190   interface.\cite{Karim88,Karim90,Hayward01,Bryk02,Hayward02} They have
191 < found for the SPC/E water model,\cite{Berendsen87} which is also used
192 < in this study, the ice/water interface is most stable at
193 < 225$\pm$5K.\cite{Bryk02} To create a ice / water interface, a box of
194 < liquid water that had the same dimensions in $x$ and $y$ was
195 < equilibrated at 225 K and 1 atm of pressure in the NPAT ensemble (with
196 < the $z$ axis allowed to fluctuate to equilibrate to the correct
197 < pressure).  The liquid and solid systems were combined by carving out
198 < any water molecule from the liquid simulation cell that was within 3
199 < \AA\ of any atom in the ice slab.
191 > found that for the SPC/E water model,\cite{Berendsen87} which is also
192 > used in this study, the ice/water interface is most stable at
193 > 225$\pm$5K.\cite{Bryk02} Therefore, we created our ice / water
194 > interfaces, utilizing a box of liquid water that had the same
195 > dimensions in $x$ and $y$ was equilibrated at 225 K and 1 atm of
196 > pressure in the NPAT ensemble (with the $z$ axis allowed to fluctuate
197 > to equilibrate to the correct pressure).  The liquid and solid systems
198 > were combined by carving out any water molecule from the liquid
199 > simulation cell that was within 3 \AA\ of any atom in the ice slab.
200  
201   Molecular translation and orientational restraints were applied in the
202   early stages of equilibration to prevent melting of the ice slab.
203   These restraints were removed during NVT equilibration, well before
204 < data collection was carried out.
204 > data collection was carried out.  
205  
206   \subsection{Shearing ice / water interfaces without bulk melting}
207  
208 < As one drags a solid through a liquid, there will be frictional
209 < heating that will act to melt the interface.  To study the frictional
210 < behavior of the interface without causing the interface to melt, it is
211 < necessary to apply a weak thermal gradient along with the momentum
212 < gradient.  This can be accomplished with of the newly-developed
213 < approaches to reverse non-equilibrium molecular dynamics (RNEMD).  The
214 < velocity shearing and scaling (VSS) variant of RNEMD utilizes a series
215 < of simultaneous velocity exchanges between two regions within the
216 < simulation cell.\cite{Kuang12} One of these regions is centered within
217 < the ice slab, while the other is centrally located in the liquid phase
208 > As a solid is dragged through a liquid, there is frictional heating
209 > that will act to melt the interface.  To study the behavior of the
210 > interface under a shear stress without causing the interface to melt,
211 > it is necessary to apply a weak thermal gradient along with the
212 > momentum gradient.  This can be accomplished using he velocity
213 > shearing and scaling (VSS) variant of reverse non-equilibrium
214 > molecular dynamics (RNEMD), which utilizes a series of simultaneous
215 > velocity exchanges between two regions within the simulation
216 > cell.\cite{Kuang12} One of these regions is centered within the ice
217 > slab, while the other is centrally located in the liquid phase
218   region. VSS-RNEMD provides a set of conservation constraints for
219   simultaneously creating either a momentum flux or a thermal flux (or
220   both) between the two slabs.  Satisfying the constraint equations
221   ensures that the new configurations are sampled from the same NVE
222 < ensemble as previously.
222 > ensemble as before the VSS move.
223  
224   The VSS moves are applied periodically to scale and shift the particle
225   velocities ($\mathbf{v}_i$ and $\mathbf{v}_j$) in two slabs ($H$ and
# Line 219 | Line 274 | response to the applied flux.  In a bulk material it i
274  
275   As the simulation progresses, the VSS moves are performed on a regular
276   basis, and the system develops a thermal and/or velocity gradient in
277 < response to the applied flux.  In a bulk material it is quite simple
277 > response to the applied flux.  In a bulk material, it is quite simple
278   to use the slope of the temperature or velocity gradients to obtain
279 < the thermal conductivity or shear viscosity.
279 > either the thermal conductivity or shear viscosity.
280  
281   The VSS-RNEMD approach is versatile in that it may be used to
282   implement thermal and shear transport simultaneously.  Perturbations
# Line 231 | Line 286 | Here we are using this method primarily to generate a
286   map out the shear viscosity of SPC/E water over a wide range of
287   temperatures (90~K) with a single 1 ns simulation.\cite{Kuang12}
288  
289 < Here we are using this method primarily to generate a shear between
290 < the ice slab and the liquid phase, while using a weak thermal gradient
291 < to maintaining the interface at the 225K target value.  This will
292 < insure minimal melting of the bulk ice phase and allows us to control
293 < the exact temperature of the interface.
289 > For this work, we are using the VSS-RNEMD method primarily to generate
290 > a shear between the ice slab and the liquid phase, while using a weak
291 > thermal gradient to maintaining the interface at the 225K target
292 > value. This will insure minimal melting of the bulk ice phase and
293 > allows us to control the exact temperature of the interface.
294  
295   \subsection{Computational Details}
296 < All simulations were performed using OpenMD with a time step of 2 fs,
297 < and periodic boundary conditions in all three dimensions. The systems
298 < were divided into 100 artificial bins along the $z$-axis for the
299 < VSS-RNEMD moves, which were attempted every 50 fs. The gradients were
300 < allowed to develop for 1 ns before data collection was began. Once
301 < established, four successive 0.5 ns runs were performed for each shear rate. During these simulations, snapshots of the system were taken every 1 ps, and the
247 < average velocities and densities of each bin were accumulated every
248 < attempted VSS-RNEMD move.
296 > All simulations were performed using OpenMD,\cite{OOPSE,openmd} with a
297 > time step of 2 fs and periodic boundary conditions in all three
298 > dimensions.  Electrostatics were handled using the damped-shifted
299 > force real-space electrostatic kernel.\cite{Ewald} The systems were
300 > divided into 100 bins along the $z$-axis for the VSS-RNEMD moves,
301 > which were attempted every 50 fs.
302  
303 < %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.
304 < %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.
305 < %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.
306 < %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.
303 > The interfaces were equilibrated for a total of 10 ns at equilibrium
304 > conditions before being exposed to either a shear or thermal gradient.
305 > This consisted of 5 ns under a constant temperature (NVT) integrator
306 > set to 225K followed by 5 ns under a microcanonical integrator.  Weak
307 > thermal gradients were allowed to develop using the VSS-RNEMD (NVE)
308 > integrator using a a small thermal flux ($-2.0\times 10^{-6}$
309 > kcal/mol/\AA$^2$/fs) for a duration of 5 ns to allow the gradient to
310 > stabilize.  The resulting temperature gradient was less than 5K over
311 > the entire 1 nm box length, which was sufficient to keep the
312 > temperature at the interface within $\pm 1$ K of the 225K target.
313  
314 + Velocity gradients were then imposed using the VSS-RNEMD (NVE)
315 + integrator with a range of momentum fluxes.  These gradients were
316 + allowed to stabilize for 1 ns before data collection began. Once
317 + established, four successive 0.5 ns runs were performed for each shear
318 + rate.  During these simulations, snapshots of the system were taken
319 + every 1 ps, and statistics on the structure and dynamics in each bin
320 + were accumulated throughout the simulations.
321 +
322   \section{Results and discussion}
323  
324   \subsection{Measuring the Width of the Interface}
325 < \subsubsection{Tetrahedrality Order Parameter}
326 < Any parameter or function that varies across the interface from a bulk liquid value to a solid value can be used as a measure of the width of the interface. However, due to the VSS-RNEMD moves pertrurbing the momentum of the molecules, parameters such as the translational order parameter and the diffusion order parameter may be artifically skewed. A structural parameter such as the pairwise correlation function would not be influenced by the perturbations. Here, the local order tetraherdal parameter as described by Kumar\cite{Kumar09} and
327 < Errington\cite{Errington01} was used as a measure of the interfacial width.
325 > Any order parameter or correlation function that varies across the
326 > interface from a bulk liquid to a solid can be used as a measure of
327 > the width of the interface.  However, because VSS-RNEMD imposes a
328 > lateral flow, parameters that depend on translational motion of the
329 > molecules (e.g. the diffusion constant) may be artifically skewed by
330 > the RNEMD moves.  A structural parameter like a radial distribution
331 > function is not influenced by the RNEMD perturbations to the same
332 > degree. Here, we have used the local tetraherdal order parameter as
333 > described by Kumar\cite{Kumar09} and Errington\cite{Errington01} as a
334 > measure of the interfacial width.
335  
336 < The local tetrahedral order parameter, $q$, is given by
336 > The local tetrahedral order parameter, $q(z)$, is given by
337   \begin{equation}
338 < 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
338 > q(z) \equiv \int_0^L \sum_{k=1}^{N} \Bigg(1 -\frac{3}{8}\sum_{i=1}^{3}
339 > \sum_{j=i+1}^{4} \bigg[\cos\psi_{ikj}+\frac{1}{3}\bigg]^2\Bigg)
340 > \delta(z_{k}-z)\mathrm{d}z
341 > \label{eq:qz}
342   \end{equation}
343 < 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.
343 > where $\psi_{ikj}$ is the angle formed between the oxygen site on
344 > molecule $k$, and the oxygen sites on its two closest neighbors,
345 > molecules $i$ and $j$.  The local tetrahedral order parameter function
346 > has a range of (0,1), where the larger the value $q$ has the more
347 > tetrahedral the ordering of the local environment is.  A $q$ value of
348 > one describes a perfectly tetrahedral environment relative to it and
349 > its four nearest neighbors, and the parameter's value decreases as the
350 > local ordering becomes less tetrahedral.   Equation \ref{eq:qz}
351 > describes a $z$-binned tetrahedral order parameter in which the $z$
352 > coordinate of the central molecule is used to give a spatial
353 > description of the local orientational ordering.
354  
355 < %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.
355 > The system was divided into 100 bins along the $z$-dimension, and a
356 > cutoff radius for the neighboring molecules was set to 3.41 \AA\ .
357 > The $q_{z}$ values for each snapshot were then averaged to give a
358 > tetrahedrality profile of the system about the $z$-dimension. The
359 > profile was then fit with a hyperbolic tangent function given by
360  
270 The system was divided into 100 bins of length $L$ along the $z$-dimension, and a cutoff radius for the neighboring molecules was set to 3.41 \AA\ .  A $q_{z}$ value was then  determined by averaging the $q$ values for each molecule in the bin.
271 \begin{equation}
272 q_{z} \equiv \int_0^L \Bigg[1 -\frac{3}{8}\sum_{i=1}^{3} \sum_{j=i+1}^{4} \bigg[\cos\psi_{ikj}+\frac{1}{3}\bigg]^2\Bigg]\delta(z_{k}-z)\mathrm{d}z
273 \end{equation}
274 The $q_{z}$ values for each snapshot were then averaged to give a tetrahedrality profile of the system about the $z$-dimension. The profile was then fit with a hyperbolic tangent function given by
275
361   \begin{equation}\label{tet_fit}
362   q_{z} \approx q_{liq}+\frac{q_{ice}-q_{liq}}{2}\Bigg[\tanh\bigg(\frac{z-I_{L,m}}{w}\bigg)-\tanh\bigg(\frac{z-I_{R,m}}{w}\bigg)\Bigg]+\beta|(z-z_{mid})|
363   \end{equation}
# Line 349 | Line 434 | Here we have simulated the basal and prismatic facets
434  
435  
436   \section{Conclusion}
437 < 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 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$0.007) amu fs\textsuperscript{-1} for the basal and prismatic facets respectively. We believe the large difference between the two friction coefficients is due to the direction of the shear and the surface structure of the crystal facets.
437 > Here we have simulated the basal and prismatic facets of an SPC/E model of the ice I$_\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.
438  
439   \begin{acknowledgement}
440    Support for this project was provided by the National Science

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines