ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/iceWater2/iceWater3.tex
Revision: 4261
Committed: Tue Dec 16 21:01:15 2014 UTC (9 years, 9 months ago) by gezelter
Content type: application/x-tex
File size: 43877 byte(s)
Log Message:
Latest version

File Contents

# Content
1 %% PNAStwoS.tex
2 %% Sample file to use for PNAS articles prepared in LaTeX
3 %% For two column PNAS articles
4 %% Version1: Apr 15, 2008
5 %% Version2: Oct 04, 2013
6
7 %% BASIC CLASS FILE
8 \documentclass{pnastwo}
9
10 %% ADDITIONAL OPTIONAL STYLE FILES Font specification
11
12 %\usepackage{PNASTWOF}
13 \usepackage[version=3]{mhchem}
14 \usepackage[round,numbers,sort&compress]{natbib}
15 \usepackage{fixltx2e}
16 \usepackage{booktabs}
17 \usepackage{multirow}
18 \usepackage{tablefootnote}
19
20 \bibpunct{(}{)}{,}{n}{,}{,}
21 \bibliographystyle{pnas2011}
22
23 %% OPTIONAL MACRO DEFINITIONS
24 \def\s{\sigma}
25 %%%%%%%%%%%%
26 %% For PNAS Only:
27 %\url{www.pnas.org/cgi/doi/10.1073/pnas.0709640104}
28 \copyrightyear{2014}
29 \issuedate{Issue Date}
30 \volume{Volume}
31 \issuenumber{Issue Number}
32 %\setcounter{page}{2687} %Set page number here if desired
33 %%%%%%%%%%%%
34
35 \begin{document}
36
37 \title{The different facets of ice have different hydrophilicities:
38 Friction at water / ice-I\textsubscript{h} interfaces}
39
40 \author{Patrick B. Louden\affil{1}{Department of Chemistry and Biochemistry, University of Notre Dame, Notre Dame,
41 IN 46556}
42 \and
43 J. Daniel Gezelter\affil{1}{}}
44
45 \contributor{Submitted to Proceedings of the National Academy of Sciences
46 of the United States of America}
47
48 %%%Newly updated.
49 %%% If significance statement need, then can use the below command otherwise just delete it.
50 \significancetext{Surface hydrophilicity is a measure of the
51 interaction strength between a solid surface and liquid water. Our
52 simulations show that the solid that is thought to be extremely
53 hydrophilic (ice) displays different behavior depending on which
54 crystal facet is presented to the liquid. This behavior is
55 potentially important in geophysics, in recognition of ice surfaces
56 by anti-freeze proteins, and in understanding how the friction
57 between ice and other solids may be mediated by a quasi-liquid layer
58 of water.}
59
60 \maketitle
61
62 \begin{article}
63 \begin{abstract}
64 We present evidence that the prismatic and secondary prism facets
65 of ice-I$_\mathrm{h}$ crystals posess structural features that can
66 reduce the effective hydrophilicity of the ice/water
67 interface. The spreading dynamics of liquid water droplets on ice
68 facets exhibits long-time behavior that differs substantially for
69 the prismatic $\{1~0~\bar{1}~0\}$ and secondary prism
70 $\{1~1~\bar{2}~0\}$ facets when compared with the basal $\{0001\}$
71 and pyramidal $\{2~0~\bar{2}~1\}$ facets. We also present the
72 results of simulations of solid-liquid friction of the same four
73 crystal facets being drawn through liquid water. These simulation
74 techniques provide evidence that the two prismatic faces have an
75 effective surface area in contact with the liquid water of
76 approximately half of the total surface area of the crystal. The
77 ice / water interfacial widths for all four crystal facets are
78 similar (using both structural and dynamic measures), and were
79 found to be independent of the shear rate. Additionally,
80 decomposition of orientational time correlation functions show
81 position-dependence for the short- and longer-time decay
82 components close to the interface.
83 \end{abstract}
84
85 \keywords{ice | water | interfaces | hydrophobicity}
86 \abbreviations{QLL, quasi liquid layer; MD, molecular dynamics; RNEMD,
87 reverse non-equilibrium molecular dynamics}
88
89 \dropcap{S}urfaces can be characterized as hydrophobic or hydrophilic
90 based on the strength of the interactions with water. Hydrophobic
91 surfaces do not have strong enough interactions with water to overcome
92 the internal attraction between molecules in the liquid phase, and the
93 degree of hydrophilicity of a surface can be described by the extent a
94 droplet can spread out over the surface. The contact angle, $\theta$,
95 formed between the solid and the liquid depends on the free energies
96 of the three interfaces involved, and is given by Young's
97 equation~\cite{Young05},
98 \begin{equation}\label{young}
99 \cos\theta = (\gamma_{sv} - \gamma_{sl})/\gamma_{lv} .
100 \end{equation}
101 Here $\gamma_{sv}$, $\gamma_{sl}$, and $\gamma_{lv}$ are the free
102 energies of the solid/vapor, solid/liquid, and liquid/vapor interfaces,
103 respectively. Large contact angles, $\theta > 90^{\circ}$, correspond
104 to hydrophobic surfaces with low wettability, while small contact
105 angles, $\theta < 90^{\circ}$, correspond to hydrophilic surfaces.
106 Experimentally, measurements of the contact angle of sessile drops is
107 often used to quantify the extent of wetting on surfaces with
108 thermally selective wetting
109 characteristics~\cite{Tadanaga00,Liu04,Sun04}.
110
111 Nanometer-scale structural features of a solid surface can influence
112 the hydrophilicity to a surprising degree. Small changes in the
113 heights and widths of nano-pillars can change a surface from
114 superhydrophobic, $\theta \ge 150^{\circ}$, to hydrophilic, $\theta
115 \sim 0^{\circ}$.\cite{Koishi09} This is often referred to as the
116 Cassie-Baxter to Wenzel transition. Nano-pillared surfaces with
117 electrically tunable Cassie-Baxter and Wenzel states have also been
118 observed~\cite{Herbertson06,Dhindsa06,Verplanck07,Ahuja08,Manukyan11}.
119 Luzar and coworkers have modeled these transitions on nano-patterned
120 surfaces~\cite{Daub07,Daub10,Daub11,Ritchie12}, and have found the
121 change in contact angle is due to the field-induced perturbation of
122 hydrogen bonding at the liquid/vapor interface~\cite{Daub07}.
123
124 One would expect the interfaces of ice to be highly hydrophilic (and
125 possibly the most hydrophilic of all solid surfaces). In this paper we
126 present evidence that some of the crystal facets of ice-I$_\mathrm{h}$
127 have structural features that can reduce the effective hydrophilicity.
128 Our evidence for this comes from molecular dynamics (MD) simulations
129 of the spreading dynamics of liquid droplets on these facets, as well
130 as reverse non-equilibrium molecular dynamics (RNEMD) simulations of
131 solid-liquid friction.
132
133 Quiescent ice-I$_\mathrm{h}$/water interfaces have been studied
134 extensively using computer simulations. Haymet \textit{et al.}
135 characterized and measured the width of these interfaces for the
136 SPC~\cite{Karim90}, SPC/E~\cite{Gay02,Bryk02},
137 CF1~\cite{Hayward01,Hayward02} and TIP4P~\cite{Karim88} models, in
138 both neat water and with solvated
139 ions~\cite{Bryk04,Smith05,Wilson08,Wilson10}. Nada and Furukawa have
140 studied the width of basal/water and prismatic/water
141 interfaces~\cite{Nada95} as well as crystal restructuring at
142 temperatures approaching the melting point~\cite{Nada00}.
143
144 The surface of ice exhibits a premelting layer, often called a
145 quasi-liquid layer (QLL), at temperatures near the melting point. MD
146 simulations of the facets of ice-I$_\mathrm{h}$ exposed to vacuum have
147 found QLL widths of approximately 10 \AA\ at 3 K below the melting
148 point~\cite{Conde08}. Similarly, Limmer and Chandler have used the mW
149 water model~\cite{Molinero09} and statistical field theory to estimate
150 QLL widths at similar temperatures to be about 3 nm~\cite{Limmer14}.
151
152 Recently, Sazaki and Furukawa have developed a technique using laser
153 confocal microscopy combined with differential interference contrast
154 microscopy that has sufficient spatial and temporal resolution to
155 visulaize and quantitatively analyze QLLs on ice crystals at
156 temperatures near melting~\cite{Sazaki10}. They have found the width of
157 the QLLs perpindicular to the surface at -2.2$^{o}$C to be 3-4 \AA\
158 wide. They have also seen the formation of two immiscible QLLs, which
159 displayed different dynamics on the crystal surface~\cite{Sazaki12}.
160
161 % There is now significant interest in the \textit{tribological}
162 % properties of ice/ice and ice/water interfaces in the geophysics
163 % community. Understanding the dynamics of solid-solid shearing that is
164 % mediated by a liquid layer~\cite{Cuffey99, Bell08} will aid in
165 % understanding the macroscopic motion of large ice
166 % masses~\cite{Casassa91, Sukhorukov13, Pritchard12, Lishman13}.
167
168 Using molecular dynamics simulations, Samadashvili has recently shown
169 that when two smooth ice slabs slide past one another, a stable
170 liquid-like layer develops between them~\cite{Samadashvili13}. In a
171 previous study, our RNEMD simulations of ice-I$_\mathrm{h}$ shearing
172 through liquid water have provided quantitative estimates of the
173 solid-liquid kinetic friction coefficients~\cite{Louden13}. These
174 displayed a factor of two difference between the basal and prismatic
175 facets. The friction was found to be independent of shear direction
176 relative to the surface orientation. We attributed facet-based
177 difference in liquid-solid friction to the 6.5 \AA\ corrugation of the
178 prismatic face which reduces the effective surface area of the ice
179 that is in direct contact with liquid water.
180
181 In the sections that follow, we describe the simulations of
182 droplet-spreading dynamics using standard MD and simulations of
183 tribological properties using RNEMD. These simulations give
184 complementary results that point to the prismatic and secondary prism
185 facets having roughly half of their surface area in direct contact
186 with the liquid.
187
188 \subsection{Construction of the Ice / Water interfaces}
189 To construct the four interfacial ice/water systems, a model of a
190 proton-ordered, zero-dipole crystal of ice-I$_\mathrm{h}$ was created.
191 This crystal structure was cleaved to expose each of the four facets
192 under investigation, and these were solvated in boxes of liquid water.
193 The procedure for creating the interfaces and equilibrating with a
194 liquid phase is described in more detail in the SI.
195
196 The SPC/E water model~\cite{Berendsen87} has been extensively
197 characterized over a wide range of liquid
198 conditions~\cite{Arbuckle02,Kuang12}, and its phase diagram has been
199 well studied~\cite{Baez95,Bryk04b,Sanz04b,Fennell:2005fk}, With longer
200 cutoff radii and careful treatment of electrostatics, SPC/E mostly
201 avoids metastable crystalline morphologies like
202 ice-\textit{i}~\cite{Fennell:2005fk} and ice-B~\cite{Baez95}. The
203 free energies and melting
204 points~\cite{Baez95,Arbuckle02,Gay02,Bryk02,Bryk04b,Sanz04b,Fennell:2005fk,Fernandez06,Abascal07,Vrbka07}
205 of various other crystalline polymorphs have also been calculated.
206 Haymet \textit{et al.} have studied quiescent Ice-I$_\mathrm{h}$/water
207 interfaces using the SPC/E water model, and have seen structural and
208 dynamic measurements of the interfacial width that agree well with
209 more expensive water models, although the coexistence temperature for
210 SPC/E is still well below the experimental melting point of real
211 water~\cite{Bryk02}. Given the extensive data and speed of this model,
212 it is a reasonable choice even though the temperatures required are
213 somewhat lower than real ice / water interfaces.
214
215 \section{Droplet Simulations}
216 Ice interfaces with a thickness of $\sim$~20~\AA\ were created as
217 described above, but were not solvated in a liquid box. The
218 crystals were then replicated along the $x$ and $y$ axes (parallel to
219 the surface) until a large surface ($>$ 126 nm\textsuperscript{2}) had
220 been created. The sizes and numbers of molecules in each of the
221 surfaces is given in Table S1. Weak translational restraining
222 potentials with spring constants of 1.5~$\mathrm{kcal\
223 mol}^{-1}\mathrm{~\AA}^{-2}$ (prismatic and pyramidal facets) or
224 4.0~$\mathrm{kcal\ mol}^{-1}\mathrm{~\AA}^{-2}$ (basal facet) were
225 applied to the centers of mass of each molecule in order to prevent
226 surface melting, although the molecules were allowed to reorient
227 freely. A water doplet containing 2048 SPC/E molecules was created
228 separately. Droplets of this size can produce agreement with the Young
229 contact angle extrapolated to an infinite drop size~\cite{Daub10}. The
230 surfaces and droplet were independently equilibrated to 225 K, at
231 which time the droplet was placed 3-5~\AA\ above the surface. Five
232 statistically independent simulations were carried out for each facet,
233 and the droplet was placed at unique $x$ and $y$ locations for each of
234 these simulations. Each simulation was 5~ns in length and was
235 conducted in the microcanonical (NVE) ensemble. Representative
236 configurations for the droplet on the prismatic facet are shown in
237 figure \ref{fig:Droplet}.
238
239 \section{Shearing Simulations (Interfaces in Bulk Water)}
240 To perform the shearing simulations, the velocity shearing and scaling
241 variant of reverse non-equilibrium molecular dynamics (VSS-RNEMD) was
242 employed \cite{Kuang12}. This method performs a series of simultaneous
243 non-equilibrium exchanges of linear momentum and kinetic energy
244 between two physically-separated regions of the simulation cell. The
245 system responds to this unphysical flux with velocity and temperature
246 gradients. When VSS-RNEMD is applied to bulk liquids, transport
247 properties like the thermal conductivity and the shear viscosity are
248 easily extracted assuming a linear response between the flux and the
249 gradient. At the interfaces between dissimilar materials, the same
250 method can be used to extract \textit{interfacial} transport
251 properties (e.g. the interfacial thermal conductance and the
252 hydrodynamic slip length).
253
254 The kinetic energy flux (producing a thermal gradient) is necessary
255 when performing shearing simulations at the ice-water interface in
256 order to prevent the frictional heating due to the shear from melting
257 the crystal. Reference \citealp{Louden13} provides more details on the
258 VSS-RNEMD method as applied to ice-water interfaces. A representative
259 configuration of the solvated prismatic facet being sheared through
260 liquid water is shown in figure \ref{fig:Shearing}.
261
262 The exchanges between the two regions were carried out every 2 fs
263 (i.e. every time step). Although computationally expensive, this was
264 done to minimize the magnitude of each individual momentum exchange.
265 Because individual VSS-RNEMD exchanges conserve both total energy and
266 linear momentum, the method can be ``bolted-on'' to simulations in any
267 ensemble. The simulations of the pyramidal interface were performed
268 under the canonical (NVT) ensemble. When time correlation functions
269 were computed, the RNEMD simulations were done in the microcanonical
270 (NVE) ensemble. All simulations of the other interfaces were carried
271 out in the microcanonical ensemble.
272
273 \section{Results}
274 \subsection{Ice - Water Contact Angles}
275 To determine the extent of wetting for each of the four crystal
276 facets, contact angles for liquid droplets on the ice surfaces were
277 computed using two methods. In the first method, the droplet is
278 assumed to form a spherical cap, and the contact angle is estimated
279 from the $z$-axis location of the droplet's center of mass
280 ($z_\mathrm{cm}$). This procedure was first described by Hautman and
281 Klein~\cite{Hautman91}, and was utilized by Hirvi and Pakkanen in
282 their investigation of water droplets on polyethylene and poly(vinyl
283 chloride) surfaces~\cite{Hirvi06}. For each stored configuration, the
284 contact angle, $\theta$, was found by inverting the expression for the
285 location of the droplet center of mass,
286 \begin{equation}\label{contact_1}
287 \langle z_\mathrm{cm}\rangle = 2^{-4/3}R_{0}\bigg(\frac{1-cos\theta}{2+cos\theta}\bigg)^{1/3}\frac{3+cos\theta}{2+cos\theta} ,
288 \end{equation}
289 where $R_{0}$ is the radius of the free water droplet.
290
291 A second method for obtaining the contact angle was described by
292 Ruijter, Blake, and Coninck~\cite{Ruijter99}, and is described in more
293 detail in the SI. Both methods provided similar estimates of the
294 dynamic contact angle, although the method described above is
295 significantly less prone to noise, and is the method used to compute
296 our contact angles.
297
298 Because the initial droplet was placed above the surface, the initial
299 value of 180$^{\circ}$ decayed over time (See figure S1 in the SI).
300 Each of these profiles were fit to a biexponential decay, with a
301 short-time contribution ($\tau_c$) that describes the initial contact
302 with the surface, a long time contribution ($\tau_s$) that describes
303 the spread of the droplet over the surface, and a constant
304 ($\theta_\infty$) to capture the infinite-time estimate of the
305 equilibrium contact angle,
306 \begin{equation}
307 \theta(t) = \theta_\infty + (180-\theta_\infty) \left[ a e^{-t/\tau_c} +
308 (1-a) e^{-t/\tau_s} \right]
309 \end{equation}
310 We have found that the rate for water droplet spreading across all
311 four crystal facets, $k_\mathrm{spread} = 1/\tau_s \approx$ 0.7
312 ns$^{-1}$. However, the basal and pyramidal facets produced estimated
313 equilibrium contact angles, $\theta_\infty \approx$ 35$^{\circ}$, while
314 prismatic and secondary prismatic had values for $\theta_\infty$ near
315 43$^{\circ}$ as seen in Table \ref{tab:kappa}.
316
317 These results indicate that, by traditional measures, the basal and
318 pyramidal facets are more hydrophilic than the prismatic and secondary
319 prism facets, and surprisingly, that the differential hydrophilicities
320 of the crystal facets is not reflected in the spreading rate of the
321 droplet.
322
323 % This is in good agreement with our calculations of friction
324 % coefficients, in which the basal
325 % and pyramidal had a higher coefficient of kinetic friction than the
326 % prismatic and secondary prismatic. Due to this, we beleive that the
327 % differences in friction coefficients can be attributed to the varying
328 % hydrophilicities of the facets.
329
330 \subsection{Solid-liquid friction of the interfaces}
331 In a bulk fluid, the shear viscosity, $\eta$, can be determined
332 assuming a linear response to a shear stress,
333 \begin{equation}\label{Shenyu-11}
334 j_{z}(p_{x}) = \eta \frac{\partial v_{x}}{\partial z}.
335 \end{equation}
336 Here $j_{z}(p_{x})$ is the flux (in $x$-momentum) that is transferred
337 in the $z$ direction (i.e. the shear stress). The RNEMD simulations
338 impose an artificial momentum flux between two regions of the
339 simulation, and the velocity gradient is the fluid's response. This
340 technique has now been applied quite widely to determine the
341 viscosities of a number of bulk fluids~\cite{XXXX}.
342
343 At the interface between two phases (e.g. liquid / solid) the same
344 momentum flux creates a velocity difference between the two materials,
345 and this can be used to define an interfacial friction coefficient
346 ($\kappa$),
347 \begin{equation}\label{Shenyu-13}
348 j_{z}(p_{x}) = \kappa \left[ v_{x}(liquid) - v_{x}(solid) \right]
349 \end{equation}
350 where $v_{x}(liquid)$ and $v_{x}(solid)$ are the velocities measured
351 directly adjacent to the interface.
352
353 The simulations described here contain significant quantities of both
354 liquid and solid phases, and the momentum flux must traverse a region
355 of the liquid that is simultaneously under a thermal gradient. Since
356 the liquid has a temperature-dependent shear viscosity, $\eta(T)$,
357 estimates of the solid-liquid friction coefficient can be obtained if
358 one knows the viscosity of the liquid at the interface (i.e. at the
359 melting temperature, $T_m$),
360 \begin{equation}\label{kappa-2}
361 \kappa = \frac{\eta(T_{m})}{\left[v_{x}(fluid)-v_{x}(solid)\right]}\left(\frac{\partial v_{x}}{\partial z}\right).
362 \end{equation}
363 For SPC/E, the melting temperature of Ice-I$_\mathrm{h}$ is estimated
364 to be 225~K~\cite{Bryk02}. To obtain the value of
365 $\eta(225\mathrm{~K})$ for the SPC/E model, a $31.09 \times 29.38
366 \times 124.39$ \AA\ box with 3744 water molecules in a disordered
367 configuration was equilibrated to 225~K, and five
368 statistically-independent shearing simulations were performed (with
369 imposed fluxes that spanned a range of $3 \rightarrow 13
370 \mathrm{~m~s}^{-1}$ ). Each simulation was conducted in the
371 microcanonical ensemble with total simulation times of 5 ns. The
372 VSS-RNEMD exchanges were carried out every 2 fs. We estimate
373 $\eta(225\mathrm{~K})$ to be 0.0148 $\pm$ 0.0007 Pa s for SPC/E,
374 roughly ten times larger than the shear viscosity previously computed
375 at 280~K~\cite{Kuang12}.
376
377 The interfacial friction coefficient can equivalently be expressed as
378 the ratio of the viscosity of the fluid to the hydrodynamic slip
379 length, $\kappa = \eta / \delta$. The slip length is an indication of
380 strength of the interactions between the solid and liquid phases,
381 although the connection between slip length and surface hydrophobicity
382 is not yet clear. In some simulations, the slip length has been found
383 to have a link to the effective surface
384 hydrophobicity~\cite{Sendner:2009uq}, although Ho \textit{et al.} have
385 found that liquid water can also slip on hydrophilic
386 surfaces~\cite{Ho:2011zr}. Experimental evidence for a direct tie
387 between slip length and hydrophobicity is also not
388 definitive. Total-internal reflection particle image velocimetry
389 (TIR-PIV) studies have suggested that there is a link between slip
390 length and effective
391 hydrophobicity~\cite{Lasne:2008vn,Bouzigues:2008ys}. However, recent
392 surface sensitive cross-correlation spectroscopy (TIR-FCCS)
393 measurements have seen similar slip behavior for both hydrophobic and
394 hydrophilic surfaces~\cite{Schaeffel:2013kx}.
395
396 In each of the systems studied here, the interfacial temperature was
397 kept fixed to 225~K, which ensured the viscosity of the fluid at the
398 interace was identical. Thus, any significant variation in $\kappa$
399 between the systems is a direct indicator of the slip length and the
400 effective interaction strength between the solid and liquid phases.
401
402 The calculated $\kappa$ values found for the four crystal facets of
403 Ice-I$_\mathrm{h}$ are shown in Table \ref{tab:kappa}. The basal and
404 pyramidal facets were found to have similar values of $\kappa \approx
405 6$ ($\times 10^{-4} \mathrm{amu~\AA}^{-2}\mathrm{fs}^{-1}$), while the
406 prismatic and secondary prism facets exhibited $\kappa \approx 3$
407 ($\times 10^{-4} \mathrm{amu~\AA}^{-2}\mathrm{fs}^{-1}$). These
408 results are also essentially independent of shearing direction
409 relative to features on the surface of the facets. The friction
410 coefficients indicate that the basal and pyramidal facets have
411 significantly stronger interactions with liquid water than either of
412 the two prismatic facets. This is in agreement with the contact angle
413 results above - both of the high-friction facets exhbited smaller
414 contact angles, suggesting that the solid-liquid friction is
415 correlated with the hydrophilicity of these facets.
416
417 \subsection{Structural measures of interfacial width under shear}
418 One of the open questions about ice/water interfaces is whether the
419 thickness of the 'slush-like' quasi-liquid layer (QLL) depends on the
420 facet of ice presented to the water. In the QLL region, the water
421 molecules are ordered differently than in either the solid or liquid
422 phases, and also exhibit distinct dynamical behavior. The width of
423 this quasi-liquid layer has been estimated by finding the distance
424 over which structural order parameters or dynamic properties change
425 from their bulk liquid values to those of the solid ice. The
426 properties used to find interfacial widths have included the local
427 density, the diffusion constant, and both translational and
428 orientational order
429 parameters~\cite{Karim88,Karim90,Hayward01,Hayward02,Bryk02,Gay02,Louden13}.
430
431 The VSS-RNEMD simulations impose thermal and velocity gradients.
432 These gradients perturb the momenta of the water molecules, so
433 parameters that depend on translational motion are often measuring the
434 momentum exchange, and not physical properties of the interface. As a
435 structural measure of the interface, we have used the local
436 tetrahedral order parameter, which measures the match of the local
437 molecular environments (e.g. the angles between nearest neighbor
438 molecules) to perfect tetrahedral ordering. This quantity was
439 originally described by Errington and Debenedetti~\cite{Errington01}
440 and has been used in bulk simulations by Kumar \textit{et
441 al.}~\cite{Kumar09} It has previously been used in ice/water
442 interfaces by by Bryk and Haymet~\cite{Bryk04b}.
443
444 To determine the structural widths of the interfaces under shear,
445 tetrahedrality profiles for each of the interfaces were computed (see
446 the SI for more details). The lower panels in figures S2-S5 in the SI
447 show tetrahedrality profiles (in circles) for each of the four
448 interfaces. These tetrahedrality profiles were fit using a hyperbolic
449 tangent function (see Eq. 6 in Reference \citealp{Louden13}) designed
450 to smoothly fit the bulk to ice transition while accounting for the
451 weak thermal gradient. This fit provides an estimate of the
452 structural width of the interface.
453
454 We find the interfacial width to be $3.2 \pm 0.2$ \AA\ (pyramidal) and
455 $3.2 \pm 0.2$ \AA\ (secondary prism) for the control systems with no
456 applied momentum flux. This is similar to our previous results for the
457 interfacial widths of the quiescent basal ($3.2 \pm 0.4$ \AA) and
458 prismatic systems ($3.6 \pm 0.2$ \AA).
459
460 Over the range of shear rates investigated, $0.4 \rightarrow
461 6.0~\mathrm{~m~s}^{-1}$ for the pyramidal system and $0.6 \rightarrow
462 5.2~\mathrm{~m~s}^{-1}$ for the secondary prism, we found no
463 significant change in the interfacial width. The mean interfacial
464 widths are collected in table \ref{tab:kappa}. This follows our
465 previous findings of the basal and prismatic systems, in which the
466 interfacial widths of the basal and prismatic facets were also found
467 to be insensitive to the shear rate~\cite{Louden13}.
468
469 The similarity of these interfacial width estimates indicate that the
470 particular facet of the exposed ice crystal has little to no effect on
471 how far into the bulk the ice-like structural ordering persists. Also,
472 it appears that for the shearing rates imposed in this study, the
473 interfacial width is not structurally modified by the movement of
474 water over the ice.
475
476 \subsection{Dynamic measures of interfacial width under shear}
477 The spatially-resolved orientational time correlation function,
478 \begin{equation}\label{C(t)1}
479 C_{2}(z,t)=\langle P_{2}(\mathbf{u}_i(0)\cdot \mathbf{u}_i(t))
480 \delta(z_i(0) - z) \rangle,
481 \end{equation}
482 provides local information about the decorrelation of molecular
483 orientations in time. Here, $P_{2}$ is the second-order Legendre
484 polynomial, and $\mathbf{u}_i$ is the molecular vector that bisects
485 the HOH angle of molecule $i$. The angle brackets indicate an average
486 over all the water molecules, and the delta function restricts the
487 average to specific regions. In the crystal, decay of $C_2(z,t)$ is
488 incomplete, while liquid water correlation times are typically
489 measured in ps. Observing the spatial-transition between the decay
490 regimes can define a dynamic measure of the interfacial width.
491
492 To determine the dynamic widths of the interfaces under shear, spatial
493 profiles of the orientational time correlation functions were computed
494 (see the SI for more details). Figures S6-S9 in the SI show the
495 $z$-coordinate profiles for the three decay constants, $\tau_{short}$,
496 $\tau_{middle}$, and $\tau_{long}$.
497
498 We have estimated the dynamic interfacial width $d_\mathrm{dyn}$ by
499 fitting the profiles of all the three orientational time constants
500 with an exponential decay to the bulk-liquid behavior (see SI). These
501 values are shown in table \ref{tab:kappa}. Because the spatial
502 granularity must be quite wide to obtain reasonable profiles of
503 $C_2(z,t)$, the error estimates for the dynamic widths of the
504 interface are significantly larger than for the structural widths.
505 However, all four interfaces exhibit dynamic widths that are
506 significantly below 1~nm, and are in reasonable agreement with the
507 structural width measured above.
508
509 \section{Conclusions}
510 In this work, we used MD simulations to measure the advancing contact
511 angles of water droplets on the basal, prismatic, pyramidal, and
512 secondary prism facets of Ice-I$_\mathrm{h}$. Although there was no
513 significant change in the \textit{rate} at which the droplets spread
514 over the surface, the long-time behavior indicates that we should
515 expect to see larger equilibrium contact angles for the two prismatic
516 facets.
517
518 We have also used RNEMD simulations of water interfaces with the same
519 four crystal facets to compute solid-liquid friction coefficients. We
520 have observed coefficients of friction that differ by a factor of two
521 between the two prismatic facets and the basal and pyramidal facets.
522 Because the solid-liquid friction coefficient is directly tied to the
523 hydrodynamic slip length, this suggests that there are significant
524 differences in the overall interaction strengths between these facets
525 and the liquid layers immediately in contact with them.
526
527 The agreement between these two measures have lead us to conclude that
528 the two prismatic facets have a lower hydrophilicity than either the
529 basal or pyramidal facets. One possible explanation of this behavior
530 is that the face presented by both prismatic facets consists of deep,
531 narrow channels (i.e. stripes of adjacent rows of pairs of
532 hydrodgen-bound water molecules). At the surfaces of these facets,
533 the channels are 6.35 \AA\ wide and the sub-surface ice layer is 2.25
534 \AA\ below (and therefore blocked from hydrogen bonding with the
535 liquid). This means that only 1/2 of the surface molecules can form
536 hydrogen bonds with liquid-phase molecules.
537
538 In the basal plane, the surface features are narrower (4.49 \AA) and
539 shallower (1.3 \AA), while the pyramidal face has much wider channels
540 (8.65 \AA) which are also quite shallow (1.37 \AA). These features
541 allow liquid phase molecules to form hydrogen bonds with all of the
542 surface molecules in the basal and pyramidal facets. This means that
543 for similar surface areas, the two prismatic facets have an effective
544 hydrogen bonding surface area of half of the basal and pyramidal
545 facets. The reduction in the effective surface area would explain
546 much of the behavior observed in our simulations.
547
548 \begin{acknowledgments}
549 Support for this project was provided by the National
550 Science Foundation under grant CHE-1362211. Computational time was
551 provided by the Center for Research Computing (CRC) at the
552 University of Notre Dame.
553 \end{acknowledgments}
554
555 \bibliography{iceWater}
556 % *****************************************
557 % There is significant interest in the properties of ice/ice and ice/water
558 % interfaces in the geophysics community. Most commonly, the results of shearing
559 % two ice blocks past one
560 % another\cite{Casassa91, Sukhorukov13, Pritchard12, Lishman13} or the shearing
561 % of ice through water\cite{Cuffey99, Bell08}. Using molecular dynamics
562 % simulations, Samadashvili has recently shown that when two smooth ice slabs
563 % slide past one another, a stable liquid-like layer develops between
564 % them\cite{Samadashvili13}. To fundamentally understand these processes, a
565 % molecular understanding of the ice/water interfaces is needed.
566
567 % Investigation of the ice/water interface is also crucial in understanding
568 % processes such as nucleation, crystal
569 % growth,\cite{Han92, Granasy95, Vanfleet95} and crystal
570 % melting\cite{Weber83, Han92, Sakai96, Sakai96B}. Insight gained to these
571 % properties can also be applied to biological systems of interest, such as
572 % the behavior of the antifreeze protein found in winter
573 % flounder,\cite{Wierzbicki07,Chapsky97} and certain terrestial
574 % arthropods.\cite{Duman:2001qy,Meister29012013} Elucidating the properties which
575 % give rise to these processes through experimental techniques can be expensive,
576 % complicated, and sometimes infeasible. However, through the use of molecular
577 % dynamics simulations much of the problems of investigating these properties
578 % are alleviated.
579
580 % Understanding ice/water interfaces inherently begins with the isolated
581 % systems. There has been extensive work parameterizing models for liquid water,
582 % such as the SPC\cite{Berendsen81}, SPC/E\cite{Berendsen87},
583 % TIP4P\cite{Jorgensen85}, TIP4P/2005\cite{Abascal05},
584 % ($\dots$), and more recently, models for simulating
585 % the solid phases of water, such as the TIP4P/Ice\cite{Abascal05b} model. The
586 % melting point of various crystal structures of ice have been calculated for
587 % many of these models
588 % (SPC\cite{Karim90,Abascal07}, SPC/E\cite{Baez95,Arbuckle02,Gay02,Bryk02,Bryk04b,Sanz04b,Gernandez06,Abascal07,Vrbka07}, TIP4P\cite{Karim88,Gao00,Sanz04,Sanz04b,Koyama04,Wang05,Fernandez06,Abascal07}, TIP5P\cite{Sanz04,Koyama04,Wang05,Fernandez06,Abascal07}),
589 % and the partial or complete phase diagram for the model has been determined
590 % (SPC/E\cite{Baez95,Bryk04b,Sanz04b}, TIP4P\cite{Sanz04,Sanz04b,Koyama04}, TIP5P\cite{Sanz04,Koyama04}).
591 % Knowing the behavior and melting point for these models has enabled an initial
592 % investigation of ice/water interfaces.
593
594 % The Ice-I$_\mathrm{h}$/water quiescent interface has been extensively studied
595 % over the past 30 years by theory and experiment. Haymet \emph{et al.} have
596 % done significant work characterizing and quantifying the width of these
597 % interfaces for the SPC,\cite{Karim90} SPC/E,\cite{Gay02,Bryk02},
598 % CF1,\cite{Hayward01,Hayward02} and TIP4P\cite{Karim88} models for water. In
599 % recent years, Haymet has focused on investigating the effects cations and
600 % anions have on crystal nucleaion and
601 % melting.\cite{Bryk04,Smith05,Wilson08,Wilson10} Nada and Furukawa have studied
602 % the the basal- and prismatic-water interface width\cite{Nada95}, crystal
603 % surface restructuring at temperatures approaching the melting
604 % point\cite{Nada00}, and the mechanism of ice growth inhibition by antifreeze
605 % proteins\cite{Nada08,Nada11,Nada12}. Nada has developed a six-site water model
606 % for ice/water interfaces near the melting point\cite{Nada03}, and studied the
607 % dependence of crystal growth shape on applied pressure\cite{Nada11b}. Using
608 % this model, Nada and Furukawa have established differential
609 % growth rates for the basal, prismatic, and secondary prismatic facets of
610 % Ice-I$_\mathrm{h}$ and found their origins due to a reordering of the hydrogen
611 % bond network in water near the interface\cite{Nada05}. While the work
612 % described so far has mainly focused on bulk water on ice, there is significant
613 % interest in thin films of water on ice surfaces as well.
614
615 % It is well known that the surface of ice exhibits a premelting layer at
616 % temperatures near the melting point, often called a quasi-liquid layer (QLL).
617 % Molecular dynamics simulations of the facets of ice-I$_\mathrm{h}$ exposed
618 % to vacuum performed by Conde, Vega and Patrykiejew have found QLL widths of
619 % approximately 10 \AA\ at 3 K below the melting point\cite{Conde08}.
620 % Similarly, Limmer and Chandler have used course grain simulations and
621 % statistical field theory to estimated QLL widths at the same temperature to
622 % be about 3 nm\cite{Limmer14}.
623 % Recently, Sazaki and Furukawa have developed an experimental technique with
624 % sufficient spatial and temporal resolution to visulaize and quantitatively
625 % analyze QLLs on ice crystals at temperatures near melting\cite{Sazaki10}. They
626 % have found the width of the QLLs perpindicular to the surface at -2.2$^{o}$C
627 % to be $\mathcal{O}$(\AA). They have also seen the formation of two immiscible
628 % QLLs, which displayed different stabilities and dynamics on the crystal
629 % surface\cite{Sazaki12}. Knowledge of the hydrophilicities of each
630 % of the crystal facets would help further our understanding of the properties
631 % and dynamics of the QLLs.
632
633 % Presented here is the follow up to our previous paper\cite{Louden13}, in which
634 % the basal and prismatic facets of an ice-I$_\mathrm{h}$/water interface were
635 % investigated where the ice was sheared relative to the liquid. By using a
636 % recently developed velocity shearing and scaling approach to reverse
637 % non-equilibrium molecular dynamics (VSS-RNEMD), simultaneous temperature and
638 % velocity gradients can be applied to the system, which allows for measurment
639 % of friction and thermal transport properties while maintaining a stable
640 % interfacial temperature\cite{Kuang12}. Structural analysis and dynamic
641 % correlation functions were used to probe the interfacial response to a shear,
642 % and the resulting solid/liquid kinetic friction coefficients were reported.
643 % In this paper we present the same analysis for the pyramidal and secondary
644 % prismatic facets, and show that the differential interfacial friction
645 % coefficients for the four facets of ice-I$_\mathrm{h}$ are determined by their
646 % relative hydrophilicity by means of dynamics water contact angle
647 % simulations.
648
649 % The local tetrahedral order parameter, $q(z)$, is given by
650 % \begin{equation}
651 % q(z) = \int_0^L \sum_{k=1}^{N} \Bigg(1 -\frac{3}{8}\sum_{i=1}^{3}
652 % \sum_{j=i+1}^{4} \bigg(\cos\psi_{ikj}+\frac{1}{3}\bigg)^2\Bigg)
653 % \delta(z_{k}-z)\mathrm{d}z \Bigg/ N_z
654 % \label{eq:qz}
655 % \end{equation}
656 % where $\psi_{ikj}$ is the angle formed between the oxygen sites of molecules
657 % $i$,$k$, and $j$, where the centeral oxygen is located within molecule $k$ and
658 % molecules $i$ and $j$ are two of the closest four water molecules
659 % around molecule $k$. All four closest neighbors of molecule $k$ are also
660 % required to reside within the first peak of the pair distribution function
661 % for molecule $k$ (typically $<$ 3.41 \AA\ for water).
662 % $N_z = \int\delta(z_k - z) \mathrm{d}z$ is a normalization factor to account
663 % for the varying population of molecules within each finite-width bin.
664
665
666 % The hydrophobicity or hydrophilicity of a surface can be described by the
667 % extent a droplet of water wets the surface. The contact angle formed between
668 % the solid and the liquid, $\theta$, which relates the free energies of the
669 % three interfaces involved, is given by Young's equation.
670 % \begin{equation}\label{young}
671 % \cos\theta = (\gamma_{sv} - \gamma_{sl})/\gamma_{lv}
672 % \end{equation}
673 % Here $\gamma_{sv}$, $\gamma_{sl}$, and $\gamma_{lv}$ are the free energies
674 % of the solid/vapor, solid/liquid, and liquid/vapor interfaces respectively.
675 % Large contact angles ($\theta$ $\gg$ 90\textsuperscript{o}) correspond to low
676 % wettability and hydrophobic surfaces, while small contact angles
677 % ($\theta$ $\ll$ 90\textsuperscript{o}) correspond to high wettability and
678 % hydrophilic surfaces. Experimentally, measurements of the contact angle
679 % of sessile drops has been used to quantify the extent of wetting on surfaces
680 % with thermally selective wetting charactaristics\cite{Tadanaga00,Liu04,Sun04},
681 % as well as nano-pillared surfaces with electrically tunable Cassie-Baxter and
682 % Wenzel states\cite{Herbertson06,Dhindsa06,Verplanck07,Ahuja08,Manukyan11}.
683 % Luzar and coworkers have done significant work modeling these transitions on
684 % nano-patterned surfaces\cite{Daub07,Daub10,Daub11,Ritchie12}, and have found
685 % the change in contact angle to be due to the external field perturbing the
686 % hydrogen bonding of the liquid/vapor interface\cite{Daub07}.
687
688 % SI stuff:
689
690 % Correlation functions:
691 % To compute eq. \eqref{C(t)1}, each 0.5 ns simulation was
692 % followed by an additional 200 ps NVE simulation during which the
693 % position and orientations of each molecule were recorded every 0.1 ps.
694
695
696 \end{article}
697
698 \begin{figure}
699 \includegraphics[width=\linewidth]{Droplet}
700 \caption{\label{fig:Droplet} Computational model of a droplet of
701 liquid water spreading over the prismatic $\{1~0~\bar{1}~0\}$ facet
702 of ice, before (left) and 2.6 ns after (right) being introduced to the
703 surface. The contact angle ($\theta$) shrinks as the simulation
704 proceeds, and the long-time behavior of this angle is used to
705 estimate the hydrophilicity of the facet.}
706 \end{figure}
707
708 \begin{figure}
709 \includegraphics[width=2in]{Shearing}
710 \caption{\label{fig:Shearing} Computational model of a slab of ice
711 being sheared through liquid water. In this figure, the ice is
712 presenting two copies of the prismatic $\{1~0~\bar{1}~0\}$ facet
713 towards the liquid phase. The RNEMD simulation exchanges both
714 linear momentum (indicated with arrows) and kinetic energy between
715 the central box and the box that spans the cell boundary. The
716 system responds with weak thermal gradient and a velocity profile
717 that shears the ice relative to the surrounding liquid.}
718 \end{figure}
719
720 % \begin{figure}
721 % \includegraphics[width=\linewidth]{ContactAngle}
722 % \caption{\label{fig:ContactAngle} The dynamic contact angle of a
723 % droplet after approaching each of the four ice facets. The decay to
724 % an equilibrium contact angle displays similar dynamics. Although
725 % all the surfaces are hydrophilic, the long-time behavior stabilizes
726 % to significantly flatter droplets for the basal and pyramidal
727 % facets. This suggests a difference in hydrophilicity for these
728 % facets compared with the two prismatic facets.}
729 % \end{figure}
730
731 % \begin{figure}
732 % \includegraphics[width=\linewidth]{Pyr_comic_strip}
733 % \caption{\label{fig:pyrComic} Properties of the pyramidal interface
734 % being sheared through water at 3.8 ms\textsuperscript{-1}. Lower
735 % panel: the local tetrahedral order parameter, $q(z)$, (circles) and
736 % the hyperbolic tangent fit (turquoise line). Middle panel: the
737 % imposed thermal gradient required to maintain a fixed interfacial
738 % temperature of 225 K. Upper panel: the transverse velocity gradient
739 % that develops in response to an imposed momentum flux. The vertical
740 % dotted lines indicate the locations of the midpoints of the two
741 % interfaces.}
742 % \end{figure}
743
744 % \begin{figure}
745 % \includegraphics[width=\linewidth]{SP_comic_strip}
746 % \caption{\label{fig:spComic} The secondary prismatic interface with a shear
747 % rate of 3.5 \
748 % ms\textsuperscript{-1}. Panel descriptions match those in figure \ref{fig:pyrComic}.}
749 % \end{figure}
750
751 % \begin{figure}
752 % \includegraphics[width=\linewidth]{Pyr-orient}
753 % \caption{\label{fig:PyrOrient} The three decay constants of the
754 % orientational time correlation function, $C_2(z,t)$, for water as a
755 % function of distance from the center of the ice slab. The vertical
756 % dashed line indicates the edge of the pyramidal ice slab determined
757 % by the local order tetrahedral parameter. The control (circles) and
758 % sheared (squares) simulations were fit using shifted-exponential
759 % decay (see Eq. 9 in Ref. \citealp{Louden13}).}
760 % \end{figure}
761
762 % \begin{figure}
763 % \includegraphics[width=\linewidth]{SP-orient-less}
764 % \caption{\label{fig:SPorient} Decay constants for $C_2(z,t)$ at the secondary
765 % prismatic face. Panel descriptions match those in \ref{fig:PyrOrient}.}
766 % \end{figure}
767
768
769 \begin{table}[h]
770 \centering
771 \caption{Structural and dynamic properties of the interfaces of
772 Ice-I$_\mathrm{h}$ with water.\label{tab:kappa}}
773 \begin{tabular}{r|cc|cc|cccc}
774 \toprule
775 \multirow{2}{*}{Interface} & \multicolumn{2}{c|}{Channel Size} &\multicolumn{2}{c|}{Droplet} & \multicolumn{4}{c}{Shearing\footnotemark[1]}\\
776 & Width (\AA) & Depth (\AA) & $\theta_{\infty}$ ($^\circ$) & $k_\mathrm{spread}$ (ns\textsuperscript{-1}) &
777 $\kappa_{x}$ & $\kappa_{y}$ & $d_\mathrm{struct}$ (\AA) & $d_\mathrm{dyn}$ (\AA) \\
778 \midrule
779 Basal $\{0001\}$ & 4.49 & 1.30 & $34.1(9)$ &$0.60(7)$
780 & $5.9(3)$ & $6.5(8)$ & $3.2(4)$ & $2(1)$ \\
781 Pyramidal $\{2~0~\bar{2}~1\}$ & 8.65 & 1.37 & $35(3)$ & $0.7(1)$ &
782 $5.8(4)$ & $6.1(5)$ & $3.2(2)$ & $2.5(3)$\\
783 Prismatic $\{1~0~\bar{1}~0\}$ & 6.35 & 2.25 & $45(3)$ & $0.75(9)$ &
784 $3.0(2)$ & $3.0(1)$ & $3.6(2)$ & $4(2)$ \\
785 Secondary Prism $\{1~1~\bar{2}~0\}$ & 6.35 & 2.25 & $43(2)$ & $0.69(3)$ &
786 $3.5(1)$ & $3.3(2)$ & $3.2(2)$ & $5(3)$ \\
787 \bottomrule
788 \end{tabular}
789 \begin{flushleft}
790 \footnotemark[1]\footnotesize{Liquid-solid friction coefficients ($\kappa_x$ and
791 $\kappa_y$) are expressed in 10\textsuperscript{-4} amu
792 \AA\textsuperscript{-2} fs\textsuperscript{-1}.} \\
793 \footnotemark[2]\footnotesize{Uncertainties in
794 the last digits are given in parentheses.}
795 \end{flushleft}
796 \end{table}
797
798 % Basal $\{0001\}$ & 4.49 & 1.30 & $34.1 \pm 0.9$ &$0.60 \pm 0.07$
799 % & $5.9 \pm 0.3$ & $6.5 \pm 0.8$ & $3.2 \pm 0.4$ & $2 \pm 1$ \\
800 % Pyramidal $\{2~0~\bar{2}~1\}$ & 8.65 & 1.37 & $35 \pm 3$ & $0.7 \pm 0.1$ &
801 % $5.8 \pm 0.4$ & $6.1 \pm 0.5$ & $3.2 \pm 0.2$ & $2.5 \pm 0.3$\\
802 % Prismatic $\{1~0~\bar{1}~0\}$ & 6.35 & 2.25 & $45 \pm 3$ & $0.75 \pm 0.09$ &
803 % $3.0 \pm 0.2$ & $3.0 \pm 0.1$ & $3.6 \pm 0.2$ & $4 \pm 2$ \\
804 % Secondary Prism $\{1~1~\bar{2}~0\}$ & 6.35 & 2.25 & $43 \pm 2$ & $0.69 \pm 0.03$ &
805 % $3.5 \pm 0.1$ & $3.3 \pm 0.2$ & $3.2 \pm 0.2$ & $5 \pm 3$ \\
806
807
808 \end{document}