ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/iceWater2/iceWater3.tex
(Generate patch)

Comparing trunk/iceWater2/iceWater3.tex (file contents):
Revision 4244 by plouden, Tue Dec 9 20:45:02 2014 UTC vs.
Revision 4245 by gezelter, Wed Dec 10 20:49:40 2014 UTC

# Line 11 | Line 11
11  
12   %\usepackage{PNASTWOF}
13   \usepackage[version=3]{mhchem}
14 < %\usepackage[square, comma, sort&compress]{natbib}
14 > \usepackage[round,numbers,sort&compress]{natbib}
15 > \usepackage{fixltx2e}
16 > \bibpunct{(}{)}{,}{n}{,}{,}
17 > \bibliographystyle{pnas2011}
18  
19   %% OPTIONAL MACRO DEFINITIONS
20   \def\s{\sigma}
# Line 27 | Line 30
30  
31   \begin{document}
32  
33 < \title{Friction at Water / Ice-I$_\mathrm{h}$ interfaces: Do the
34 <  Different Facets of Ice Have Different Hydrophilicities?}
33 > \title{Friction at water / ice-I\textsubscript{h} interfaces: Do the
34 >  different facets of ice have different hydrophilicities?}
35  
36   \author{Patrick B. Louden\affil{1}{Department of Chemistry and Biochemistry, University of Notre Dame, Notre Dame,
37   IN 46556}
# Line 45 | Line 48 | of the United States of America}
48   \maketitle
49  
50   \begin{article}
51 < \begin{abstract}
52 < {In this follow up paper of the basal and prismatic facets of the
53 < Ice-I$_\mathrm{h}$/water interface, we present the
54 < pyramidal and secondary prismatic
55 < interfaces for both the quiescent and sheared systems. The structural and
56 < dynamic interfacial widths for all four crystal facets were found to be in good
57 < agreement, and were found to be independent of the shear rate over the shear
58 < rates investigated.
59 < Decomposition of the molecular orientational time correlation function showed
60 < different behavior for the short- and longer-time decay components approaching
61 < normal to the interface. Lastly we show through calculation of the interfacial
62 < friction coefficient and dynamic water contact angle measurement
63 < that the basal and pyramidal facets are more
64 < hydrophilic than the prismatic and secondary prismatic facets.}
51 >  \begin{abstract}
52 >    In this paper we present evidence that some of the crystal facets
53 >    of ice-I$_\mathrm{h}$ posess structural features that can halve
54 >    the effective hydrophilicity of the ice/water interface. The
55 >    spreading dynamics of liquid water droplets on ice facets exhibits
56 >    long-time behavior that differs substantially for the prismatic
57 >    $\{1~0~\bar{1}~0\}$ and secondary prism $\{1~1~\bar{2}~0\}$ facets
58 >    when compared with the basal $\{0001\}$ and pyramidal
59 >    $\{2~0~\bar{2}~1\}$ facets.  We also present the results of
60 >    simulations of solid-liquid friction of the same four crystal
61 >    facets being drawn through liquid water. Both simulation
62 >    techniques provide evidence that the two prismatic faces have an
63 >    effective surface area in contact with the liquid water of
64 >    approximately half of the total surface area of the crystal.  The
65 >    ice / water interfacial widths for all four crystal facets are
66 >    similar (using both structural and dynamic measures), and were
67 >    found to be independent of the shear rate.  Additionally,
68 >    decomposition of orientational time correlation functions show
69 >    position-dependence for the short- and longer-time decay
70 >    components close to the interface.
71   \end{abstract}
72  
73   \keywords{ice | water | interfaces | hydrophobicity}
74   \abbreviations{QLL, quasi liquid layer; MD, molecular dynamics; RNEMD,
75   reverse non-equilibrium molecular dynamics}
76  
77 < \dropcap{T}he quiescent ice-I$_\mathrm{h}$/water interface has been
78 < extensively studied using computer simulations over the past 30
79 < years. Haymet \emph{et al.} characterized and measured the width of
80 < these interfaces for the SPC,\cite{Karim90} SPC/E,\cite{Gay02,Bryk02},
81 < CF1,\cite{Hayward01,Hayward02} and TIP4P\cite{Karim88} water models in
82 < neat water and with solvated
83 < ions.\cite{Bryk04,Smith05,Wilson08,Wilson10} Nada and Furukawa have
84 < studied the width of basal- and prismatic-water
85 < interfaces\cite{Nada95} as well as crystal restructuring at
86 < temperatures approaching the melting point.\cite{Nada00}
77 > \dropcap{S}urfaces can be characterized as hydrophobic or hydrophilic
78 > based on the strength of the interactions with water. Hydrophobic
79 > surfaces do not have strong enough interactions with water to overcome
80 > the internal attraction between molecules in the liquid phase, and the
81 > degree of hydrophilicity of a surface can be described by the extent a
82 > droplet can spread out over the surface. The contact angle formed
83 > between the solid and the liquid depends on the free energies of the
84 > three interfaces involved, and is given by Young's
85 > equation.\cite{Young}
86 > \begin{equation}\label{young}
87 > \cos\theta = (\gamma_{sv} - \gamma_{sl})/\gamma_{lv} .
88 > \end{equation}
89 > Here $\gamma_{sv}$, $\gamma_{sl}$, and $\gamma_{lv}$ are the free
90 > energies of the solid/vapor, solid/liquid, and liquid/vapor interfaces
91 > respectively.  Large contact angles, $\theta > 90^{\circ}$, correspond
92 > to hydrophobic surfaces with low wettability, while small contact
93 > angles, $\theta < 90^{\circ}$, correspond to hydrophilic surfaces.
94 > Experimentally, measurements of the contact angle of sessile drops is
95 > often used to quantify the extent of wetting on surfaces with
96 > thermally selective wetting
97 > characteristics.\cite{Tadanaga00,Liu04,Sun04}
98  
99 + Nanometer-scale structural features of a solid surface can influence
100 + the hydrophilicity to a surprising degree.  Small changes in the
101 + heights and widths of nano-pillars can change a surface from
102 + superhydrophobic, $\theta \ge 150^{\circ}$, to hydrophilic, $\theta
103 + \sim 0^{\circ}$.\cite{CBW} This is often referred to as the
104 + Cassie-Baxter to Wenzel transition.  Nano-pillared surfaces with
105 + electrically tunable Cassie-Baxter and Wenzel states have also been
106 + observed.\cite{Herbertson06,Dhindsa06,Verplanck07,Ahuja08,Manukyan11}
107 + Luzar and coworkers have modeled these transitions on nano-patterned
108 + surfaces\cite{Daub07,Daub10,Daub11,Ritchie12}, and have found the
109 + change in contact angle is due to the field-induced perturbation of
110 + hydrogen bonding at the liquid/vapor interface.\cite{Daub07}
111 +
112 + One would expect the interfaces of ice to be highly hydrophilic (and
113 + possibly the most hydrophilic of all solid surfaces). In this paper we
114 + present evidence that some of the crystal facets of ice-I$_\mathrm{h}$
115 + have structural features that can halve the effective hydrophilicity.
116 + Our evidence for this comes from molecular dynamics (MD) simulations
117 + of the spreading dynamics of liquid droplets on these facets, as well
118 + as reverse non-equilibrium molecular dynamics (RNEMD) simulations of
119 + solid-liquid friction.
120 +
121 + Quiescent ice-I$_\mathrm{h}$/water interfaces have been studied
122 + extensively using computer simulations. Haymet \textit{et al.}
123 + characterized and measured the width of these interfaces for the
124 + SPC~\cite{Karim90}, SPC/E~\cite{Gay02,Bryk02},
125 + CF1~\cite{Hayward01,Hayward02} and TIP4P~\cite{Karim88} models, in
126 + both neat water and with solvated
127 + ions~\cite{Bryk04,Smith05,Wilson08,Wilson10}. Nada and Furukawa have
128 + studied the width of basal/water and prismatic/water
129 + interfaces~\cite{Nada95} as well as crystal restructuring at
130 + temperatures approaching the melting point~\cite{Nada00}.
131 +
132   The surface of ice exhibits a premelting layer, often called a
133 < quasi-liquid layer (QLL), at temperatures near the melting point.
134 < Molecular dynamics simulations of the facets of ice-I$_\mathrm{h}$
135 < exposed to vacuum have found QLL widths of approximately 10 \AA\ at 3
136 < K below the melting point.\cite{Conde08} Similarly, Limmer and
137 < Chandler have used the mW water model\cite{Molinero09} and statistical field
138 < theory to estimate QLL widths at similar temperatures to be about 3
86 < nm.\cite{Limmer14}
133 > quasi-liquid layer (QLL), at temperatures near the melting point.  MD
134 > simulations of the facets of ice-I$_\mathrm{h}$ exposed to vacuum have
135 > found QLL widths of approximately 10 \AA\ at 3 K below the melting
136 > point.\cite{Conde08} Similarly, Limmer and Chandler have used the mW
137 > water model~\cite{Molinero09} and statistical field theory to estimate
138 > QLL widths at similar temperatures to be about 3 nm.\cite{Limmer14}
139  
140 < Recently, Sazaki and Furukawa have developed a technique using
141 < laser confocal micrscopy combined with differential interference
142 < contrast microscopy
143 < that has sufficient spatial and temporal resolution to visulaize and
144 < quantitatively analyze QLLs on ice crystals at temperatures near
145 < melting.\cite{Sazaki10} They have found the width of the QLLs
146 < perpindicular to the surface at -2.2$^{o}$C to be 3-4 \AA\ wide. They
147 < have also seen the formation of two immiscible QLLs, which displayed
96 < different dynamics on the crystal surface.\cite{Sazaki12}
140 > Recently, Sazaki and Furukawa have developed a technique using laser
141 > confocal microscopy combined with differential interference contrast
142 > microscopy that has sufficient spatial and temporal resolution to
143 > visulaize and quantitatively analyze QLLs on ice crystals at
144 > temperatures near melting.\cite{Sazaki10} They have found the width of
145 > the QLLs perpindicular to the surface at -2.2$^{o}$C to be 3-4 \AA\
146 > wide.  They have also seen the formation of two immiscible QLLs, which
147 > displayed different dynamics on the crystal surface.\cite{Sazaki12}
148  
149 < There is significant interest in the properties of ice/ice and
150 < ice/water interfaces in the geophysics community.  Understanding the
151 < dynamics of solid-solid shearing mediated by a liquid
152 < layer\cite{Cuffey99, Bell08} will aid in understanding the macroscopic
153 < motion of large ice masses.\cite{Casassa91, Sukhorukov13, Pritchard12,
154 <  Lishman13}
149 > There is now significant interest in the \textit{tribological}
150 > properties of ice/ice and ice/water interfaces in the geophysics
151 > community.  Understanding the dynamics of solid-solid shearing that is
152 > mediated by a liquid layer\cite{Cuffey99, Bell08} will aid in
153 > understanding the macroscopic motion of large ice
154 > masses.\cite{Casassa91, Sukhorukov13, Pritchard12, Lishman13}
155  
156   Using molecular dynamics simulations, Samadashvili has recently shown
157   that when two smooth ice slabs slide past one another, a stable
158 < liquid-like layer develops between them.\cite{Samadashvili13} We have
159 < previously used reverse non-equilibrium molecular dynamics (RNEMD)
160 < simulations of ice-I$_\mathrm{h}$ / water interfaces to shear the
161 < solid phase relative to the surrounding liquid.\cite{Louden13} The
162 < computed solid-liquid kinetic friction coefficients displayed a factor
163 < of two difference between the basal $\{0001\}$ and prismatic
164 < $\{1~0~\bar{1}~0\}$ facets.  The friction was found to be independent
165 < of shear direction relative to the surface orientation.  We attributed
166 < facet-based difference in liquid-solid friction to the 6.5 \AA\
167 < corrugation of the prismatic face which reduces the effective surface
117 < area of the ice that is in direct contact with liquid water.
118 <
119 < Surfaces can be charactarized as hydrophobic or hydrophilic based on
120 < the strength of the interactions with water. Hydrophobic surfaces do
121 < not have strong enough interactions with water to overcome the
122 < internal attraction between molecules in the liquid phase.  Water on
123 < hydrophobic surfaces maintains a nearly-spherical droplet shape.
124 < Conversely, hydrophilic surfaces have strong solid-liquid interactions
125 < and exhibit droplets that spread over the surface.
158 > liquid-like layer develops between them.\cite{Samadashvili13} In a
159 > previous study, our RNEMD simulations of ice-I$_\mathrm{h}$ shearing
160 > through liquid water have provided quantitative estimates of the
161 > solid-liquid kinetic friction coefficients.\cite{Louden13} These
162 > displayed a factor of two difference between the basal and prismatic
163 > facets.  The friction was found to be independent of shear direction
164 > relative to the surface orientation.  We attributed facet-based
165 > difference in liquid-solid friction to the 6.5 \AA\ corrugation of the
166 > prismatic face which reduces the effective surface area of the ice
167 > that is in direct contact with liquid water.
168  
169 < The hydrophobicity or hydrophilicity of a surface can be described by
170 < the extent a droplet can spread out over the surface. The contact
171 < angle formed between the solid and the liquid, $\theta$, which relates
172 < the free energies of the three interfaces involved, is given by
173 < Young's equation.
174 < \begin{equation}\label{young}
133 < \cos\theta = (\gamma_{sv} - \gamma_{sl})/\gamma_{lv}
134 < \end{equation}
135 < Here $\gamma_{sv}$, $\gamma_{sl}$, and $\gamma_{lv}$ are the free energies
136 < of the solid/vapor, solid/liquid, and liquid/vapor interfaces respectively.
137 < Large contact angles ($\theta$ $\gg$ 90\textsuperscript{o}) correspond to low
138 < wettability and hydrophobic surfaces, while small contact angles
139 < ($\theta$ $\ll$ 90\textsuperscript{o}) correspond to high wettability and
140 < hydrophilic surfaces. Experimentally, measurements of the contact angle
141 < of sessile drops has been used to quantify the extent of wetting on surfaces
142 < with thermally selective wetting charactaristics\cite{Tadanaga00,Liu04,Sun04},
143 < as well as nano-pillared surfaces with electrically tunable Cassie-Baxter and
144 < Wenzel states\cite{Herbertson06,Dhindsa06,Verplanck07,Ahuja08,Manukyan11}.
145 < Luzar and coworkers have done significant work modeling these transitions on
146 < nano-patterned surfaces\cite{Daub07,Daub10,Daub11,Ritchie12}, and have found
147 < the change in contact angle to be due to the external field perturbing the
148 < hydrogen bonding of the liquid/vapor interface\cite{Daub07}.
169 > In the sections that follow, we outline the methodology used to
170 > simulate droplet-spreading dynamics using standard MD and tribological
171 > properties using RNEMD simulations.  These simulation methods give
172 > complementary results that point to the prismatic and secondary prism
173 > facets having roughly half of their surface area in direct contact
174 > with the liquid.
175  
150 In this paper we present the same analysis for the pyramidal and secondary
151 prismatic facets, and show that the differential interfacial friction
152 coefficients for the four facets of ice-I$_\mathrm{h}$ are determined by their
153 relative hydrophilicity by means of dynamics water contact angle simulations.
154
176   \section{Methodology}
177 + \subsection{Construction of the Ice / Water Interfaces}
178 + To construct the four interfacial ice/water systems, a proton-ordered,
179 + zero-dipole crystal of ice-I$_\mathrm{h}$ with exposed strips of
180 + H-atoms and lone pairs was constructed using Structure 6 of Hirsch and
181 + Ojam\"{a}e's set of orthorhombic representations for
182 + ice-I$_{h}$.\cite{Hirsch04} This crystal structure was cleaved along
183 + the four different facets being studied.  The exposed face was
184 + reoriented normal to the $z$-axis of the simulation cell, and the
185 + structures were and extended to form large exposed facets in
186 + rectangular box geometries.  Liquid water boxes were created with
187 + identical dimensions (in $x$ and $y$) as the ice, and a $z$ dimension
188 + of three times that of the ice block, and a density corresponding to
189 + $\sim$ 1 g / cm$^3$.  Each of the ice slabs and water boxes were
190 + independently equilibrated, and the resulting systems were merged by
191 + carving out any liquid water molecules within 3 \AA\ of any atoms in
192 + the ice slabs.  Each of the combined ice/water systems were then
193 + equilibrated at 225K, which is the liquid-ice coexistence temperature
194 + for SPC/E water.\cite{} Ref. \citealp{Louden13} contains a more
195 + detailed explanation of the construction of ice/water interfaces. The
196 + resulting dimensions, number of ice, and liquid water molecules
197 + contained in each of these systems can be seen in Table
198 + \ref{tab:method}.
199  
200 < \subsection{Water Model}
201 < Understanding ice/water interfaces inherently begins with the isolated
202 < systems. There has been extensive work parameterizing models for liquid water,
203 < such as the SPC\cite{Berendsen81}, SPC/E\cite{Berendsen87},
204 < TIP4P\cite{Jorgensen85}, TIP4P/2005\cite{Abascal05},
162 < ($\dots$), and more recently, models for simulating
163 < the solid phases of water, such as the TIP4P/Ice\cite{Abascal05b} model. The
164 < melting point of various crystal structures of ice have been calculated for
165 < many of these models
166 < (SPC\cite{Karim90,Abascal07}, SPC/E\cite{Baez95,Arbuckle02,Gay02,Bryk02,Bryk04b,Sanz04b,Fernandez06,Abascal07,Vrbka07}, TIP4P\cite{Karim88,Gao00,Sanz04,Sanz04b,Koyama04,Wang05,Fernandez06,Abascal07}, TIP5P\cite{Sanz04,Koyama04,Wang05,Fernandez06,Abascal07}),
167 < and the partial or complete phase diagram for the model has been determined
168 < (SPC/E\cite{Baez95,Bryk04b,Sanz04b}, TIP4P\cite{Sanz04,Sanz04b,Koyama04}, TIP5P\cite{Sanz04,Koyama04}).
200 > We used SPC/E Why?  Extensively characterized over a wide range of
201 > liquid conditions.  Well-studied phase diagram. Reasonably accurate
202 > crystalline free energies.  Mostly avoids spurious crystalline
203 > morphologies like ice-i and ice-B.  Most importantly, the use of SPC/E
204 > has been well characterized in previous ice/water interfacial studies.
205  
170 Haymet et al. have studied the quiescent Ice-I$_\mathrm{h}$/water interface
171 using the rigid SPC, SPC/E, TIP4P, and the flexible CF1 water models, and has seen good
172 agreement for structural and dynamic measurements of the interfacial
173 width. Given the expansive size of our systems of interest, and the
174 apparent independence of water model on interfacial width, we have chosen to use the rigid SPC/E
175 water model in this study.
206  
207 < \subsection{Pyramidal and secondary prismatic ice/water interface construction}
208 < To construct the pyramidal and secondary prismatic ice/water systems,
209 < first a proton-ordered zero dipole crystal of ice-I$_\mathrm{h}$ with exposed strips
210 < of H-atoms and lone pairs was constructed from Structure 6 of Hirsch
211 < and Ojam\"{a}e's recent paper\cite{Hirsch04}. The crystal was then cut
212 < along the plane of the desired facet, and reoriented so that the
213 < $z$-axis was perpdicular to the exposed face. Two orthoganol cuts were
214 < then made to the crystal such that perfect periodic replication could
215 < be perfromed in the $x$ and $y$ dimensions. The slab was then
216 < replicated along the $x$ and $y$ axes until the desired crystal size
217 < was obtained. Liquid water boxes were created having identical
218 < dimensions (in $x$ and$y$) as the ice blocks, and a $z$ dimension of
219 < three times that of the ice block. Each of the ice slabs and water
220 < boxes were independently equilibrated to 50K, and the resulting
221 < systems were merged by carving out any liquid water molecules within 3
222 < \AA\ of any atoms in the ice slabs. Each of the combined ice/water
223 < systems were then equilibrated to 225K, which was found to be a stable
224 < temperature for each of the interfaces over a 5 ns simulation.
225 < For a more detailed explanation of
226 < the ice/water systems construction, please refer to a previous
227 < paper\cite{Louden13}. The resulting dimensions, number of ice, and liquid water molecules
228 < contained in each of these systems can be seen in Table \ref{tab:method}.
229 < \subsection{Shearing simulations}
207 >
208 > There has been extensive work parameterizing good models for liquid
209 > water over a wide range of conditions.  The melting points of various
210 > crystal structures of ice have been calculated for many of these
211 > models (SPC\cite{Karim90,Abascal07},
212 > SPC/E\cite{Baez95,Arbuckle02,Gay02,Bryk02,Bryk04b,Sanz04b,Fernandez06,Abascal07,Vrbka07},
213 > TIP4P\cite{Karim88,Gao00,Sanz04,Sanz04b,Koyama04,Wang05,Fernandez06,Abascal07},
214 > TIP5P\cite{Sanz04,Koyama04,Wang05,Fernandez06,Abascal07}), and the
215 > partial or complete phase diagram for the model has been determined
216 > (SPC/E\cite{Baez95,Bryk04b,Sanz04b},
217 > TIP4P\cite{Sanz04,Sanz04b,Koyama04}, TIP5P\cite{Sanz04,Koyama04}).
218 >
219 >
220 > such as the SPC\cite{Berendsen81}, SPC/E\cite{Berendsen87},
221 > TIP4P\cite{Jorgensen85}, TIP4P/2005\cite{Abascal05}, ($\dots$), and
222 > more recently, models for simulating the solid phases of water, such
223 > as the TIP4P/Ice\cite{Abascal05b} model.
224 >
225 > Haymet et al. have studied the quiescent Ice-I$_\mathrm{h}$/water
226 > interface using the rigid SPC, SPC/E, TIP4P, and the flexible CF1
227 > water models, and has seen good agreement for structural and dynamic
228 > measurements of the interfacial width. Given the expansive size of our
229 > systems of interest, and the apparent independence of water model on
230 > interfacial width, we have chosen to use the rigid SPC/E water model
231 > in this study.
232 >
233 > \subsection{Shearing simulations (interfaces in bulk water)}
234   % Should we mention number of runs, sim times, etc. ?
235 < To perform the shearing simulations, the velocity shearing and scaling
235 > To perform the shearing simulations, the velocity shearing and scaling
236   varient of reverse nonequilibrium molecular dynamics (VSS-RNEMD) was
237   conducted. This method performs a series of simultaneous velocity
238   exchanges between two regions of the simulation cell, to
# Line 218 | Line 252 | To construct ice surfaces to perform water contact ang
252   secondary prismatic simulations were performed under the NVE ensamble.
253  
254   \subsection{Droplet simulations}
255 < To construct ice surfaces to perform water contact angle calculations
256 < on, ice crystals were created as described earlier (see Pyramidal and
257 < secondary prismatic ice/water interface construction). The crystals
258 < were then cut from the negative $z$ dimension, ensuring the remaining
259 < ice crystal was thicker in $z$ than the potential cutoff. The crystals
260 < were then replicated in $x$ and $y$ until a sufficiently large surface
261 < had been created. The sizes and number of molecules in each of the surfaces is given in Table
262 < \ref{tab:ice_sheets}. Molecular restraints were applied to the center of mass
263 < of the rigid bodies to prevent surface melting, however the molecules were
264 < allowed to reorient themselves freely. The water doplet contained 2048
265 < SPC/E molecules, which has been found to produce
266 < agreement for the Young contact angle extrapolated to an infinite drop
267 < size\cite{Daub10}. The surfaces and droplet were independently
268 < equilibrated to 225 K, at which time the droplet was placed  3-5 \AA\
269 < above the positive $z$ dimension of the surface at 5 unique
270 < locations. Each simulation was 5 ns in length and conducted in the NVE ensemble.  
271 <
255 > Ice interfaces with a thickness of $\sim 30 \AA$ were created as
256 > described above, but were not solvated in a liquid box. The crystals
257 > were then replicated along the $x$ and $y$ axes (parallel to the
258 > surface) until a large surface had been created.  The sizes and
259 > numbers of molecules in each of the surfaces is given in Table
260 > \ref{tab:ice_sheets}.  Weak translational restraining potentials with
261 > spring constants of XXXX were applied to the center of mass of each
262 > molecule in order to prevent surface melting, although the molecules
263 > were allowed to reorient freely. A water doplet containing 2048 SPC/E
264 > molecules was created separately. Droplets of this size can produce
265 > agreement with the Young contact angle extrapolated to an infinite
266 > drop size\cite{Daub10}. The surfaces and droplet were independently
267 > equilibrated to 225 K, at which time the droplet was placed 3-5 \AA\
268 > above the surface.  Five statistically independent simulations were
269 > carried out for each facet, and the droplet was placed at unique $x$
270 > and $y$ locations for each of these simulations.  Each simulation was
271 > 5 ns in length and conducted in the microcanonical (NVE) ensemble.
272  
273   \section{Results and discussion}
274   \subsection{Interfacial width}
# Line 563 | Line 597 | hydrophilicities.
597    provided by the Center for Research Computing (CRC) at the
598    University of Notre Dame.
599   \end{acknowledgments}
566
567 \newpage
600  
569 \bibliographystyle{pnas2011}
601   \bibliography{iceWater}
602   % *****************************************
603   % There is significant interest in the properties of ice/ice and ice/water

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines