Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2017 Oct 17;114(44):11627–11632. doi: 10.1073/pnas.1621058114

Dynamics and unsteady morphologies at ice interfaces driven by D2O–H2O exchange

Ran Drori a,b,c,1, Miranda Holmes-Cerfon d,1, Bart Kahr a,b, Robert V Kohn d, Michael D Ward a,b,1
PMCID: PMC5676873  PMID: 29042511

Significance

Freezing and melting of ice are one of the most common events on Earth. The dynamics of ice crystallization are relevant to climate research, mitigating frost damage in agriculture and construction, glacier dynamics, tissue and food preservation, and transportation. We describe the use of microfluidic devices, accompanied by precise temperature control, to examine the effect of H/D isotope exchange between liquid light water and solid heavy water on ice growth dynamics. These studies revealed unusual morphologies at the ice surface in contact with the liquid, including curious unsteady morphological features that give the appearance of oscillation due to complex interplay of H/D exchange, thermal gradients, and local surface curvature.

Keywords: ice growth, microfluidics, hydrogen–deuterium exchange, Raman microscopy, ice morphology

Abstract

The growth dynamics of D2O ice in liquid H2O in a microfluidic device were investigated between the melting points of D2O ice (3.8 °C) and H2O ice (0 °C). As the temperature was decreased at rates between 0.002 °C/s and 0.1 °C/s, the ice front advanced but retreated immediately upon cessation of cooling, regardless of the temperature. This is a consequence of the competition between diffusion of H2O into the D2O ice, which favors melting of the interface, and the driving force for growth supplied by cooling. Raman microscopy tracked H/D exchange across the solid H2O–solid D2O interface, with diffusion coefficients consistent with transport of intact H2O molecules at the D2O ice interface. At fixed temperatures below 3 °C, the D2O ice front melted continuously, but at temperatures near 0 °C a scalloped interface morphology appeared with convex and concave sections that cycled between growth and retreat. This behavior, not observed for D2O ice in contact with D2O liquid or H2O ice in contact with H2O liquid, reflects a complex set of cooperative phenomena, including H/D exchange across the solid–liquid interface, latent heat exchange, local thermal gradients, and the Gibbs–Thomson effect on the melting points of the convex and concave features.


Ice formation is arguably the most common crystallization process on Earth, ranging from the formation of atmospheric snow and ice to annual cycles of freezing and thawing of ground water. Ice has more solid phases than any other substance, with 17 polymorphs reported (1) and another predicted (2). The growth dynamics of common, hexagonal (Ih) ice are essential in climate modeling, mitigating structural damage due to freeze–thaw cycles (3, 4), glacier dynamics (5), cryobiology and cryopreservation (6), and the design of new materials through ice templating (7). Heavy water (D2O) also adopts the same Ih structure and crystal habit as light water (H2O) (8), but melts at higher temperatures Tm=3.8°C vs. Tm=0°C, respectively. D2O has been used as a solvent in the cold preservation of organs (9). The higher melting temperature and lower vapor pressure of D2O are essential for estimating millennial global temperatures through analysis of deuterium composition in ice cores (10). The fractionation of water isotopomers is a manifestation of preferential freezing of D2O with respect to H2O, which affects the distribution of environmental deuterium (11, 12). Isotopic fractionation also can occur through H/D exchange, a process that has been investigated at temperatures ranging from −178 °C to −2 °C (1320). H/D exchange at lower temperatures (up to −103 °C) has been investigated by measurement of changes in isotopic composition in thin stratified films of D2O–H2O–HDO (1–100 nm in thickness) (13, 14, 17, 19). The diffusion of tritium into bulk H2O ice has been measured over the range −35 °C to −2 °C (18, 20), with diffusion coefficients spanning seven orders of magnitude. Comparison of diffusion coefficients for ice films and bulk crystals from measurements performed over such a large range of temperatures may be problematic because (i) extrapolation of Arrhenius behavior may not be reliable over the temperature range; (ii) different diffusion mechanisms may be operative at lower temperatures compared with those at higher temperatures (interstitial ion migration vs. vacancy migration); (iii) thin ice films may be polycrystalline or even amorphous, compared with the single crystallinity of bulk ice; and (iv) thin ice films tend to have more defects than single crystals, affording larger-than-actual diffusion rates in films (17). Herein, we describe ice crystallization on the surface of a D2O crystal in liquid water within a microfluidic device for which temperature can be controlled with a precision of 0.001 °C and a stability of ±0.002 °C. At a constant temperature below the melting point of D2O, 3.8 °C, the crystal interface melted slowly. As the system was cooled continuously at a sufficiently high rate, the ice front advanced, but retreated immediately after cooling ceased, signaling exchange of H for D in the solid D2O that reduced the ice melting temperature. At a constant system temperature (the temperature that was measured by the thermistor), the liquid–solid interface exhibited cyclical and unsteady morphologies owing to the combined effects of H/D exchange, latent heat dissipation, local thermal gradients, and the Gibbs–Thomson effect.

D2O–H2O Phase Diagram

Several investigations of ice-containing mixtures of H2O and D2O (2123), and implicitly HOD, have been reported, relying on vibrational spectroscopic assays of HOD through the appearance of separate O-H and O-D vibrational bands. These reports, however, have not addressed the melting (or freezing) habits of H2O:D2O mixtures. Measurement of the melting points across a range of H2O:D2O compositions can permit construction of a phase diagram, which in turn can be used to deduce the equilibrium composition of a H2O–D2O liquid–solid interface during growth or melting. This was achieved by insertion of a small amount of premixed solutions (0.5 μL total volume) with various H2O:D2O compositions into an oil droplet resting on a sapphire disc placed on a custom-built cold stage (Fig. S1). The oil droplet was then topped with a glass coverslip, extruding the oil and leaving water trapped between the coverslip and the sapphire disc (24). The assembly was cooled to −20 °C and then heated slowly to melt the ice partially, affording a disk-shaped single crystal of ice. Optical microscopy was used to determine the melting temperature of the single crystal, enabling construction of a simple binary phase diagram, which displayed a linear relationship between the melting temperature and isotopic composition (Fig. 1).

Fig. S1.

Fig. S1.

(Top) Custom-built cold-stage setting. (Bottom) Stability of the cold-stage temperature, measured with a thermistor, at 3.0 °C and after cooling to 2.0 °C at 0.05 °C/5 s. Insets depict the temperature stability before and after cooling.

Fig. 1.

Fig. 1.

Binary phase diagram for H2O:D2O mixtures determined from visual observation of freezing/melting (red circles) and the predicted melting points (black line). The phase diagram follows a simple linear dependence between melting temperature and isotopic composition. Inset shows an ice crystal formed in a 75%-H2O–25%-D2O mixture at 0.996 °C. The disc-shaped morphology is typical of crystals grown across the isotopic range. The upper face of the disc corresponds to the 0001 plane. (Scale bar, 10 μm.)

H2O Crystal Growth on D2O Ice

Crystal growth resulting from ice deposition on a D2O crystal in liquid H2O was examined in a microfluidic device that initially contained neat D2O. The temperature of the cold stage was controlled to within ±0.001 °C. After lowering the temperature to −20 °C to freeze the liquid in the microfluidic channel, the resulting D2O ice was melted partially at 3.6 °C until a well-formed single crystal remained in the channel. The length of these crystals ranged from 80 μm to 800 μm. The channel dimensions of the primary compartment prescribed their width and height (150 μm and 20 μm, respectively). The liquid D2O surrounding the crystal was then replaced with H2O, and the temperature was lowered immediately to a value below 3 °C to limit melting of the crystal. The reliability of this protocol was assessed using an indicator dye that enabled visualization of the displacement of D2O by H2O. Injection of an H2O solution containing sunset yellow dye FCF (for coloring food; 1% wt/wt; CAS no. 2783-94-0) into the microfluidic channel confirmed plug flow displacement of the liquid D2O. All subsequent experiments were performed without sunset yellow, however. At constant system temperature below the D2O melting point, the D2O crystal interface melted, as observed by the retreat of the interface, signaling a continuous reduction in the melting point due to H/D exchange across the interface. For example, at 1 °C the solid–liquid interface retreated at a rate of ca. 0.5 μm/s. This exchange, which melts the interface, adds H to the D2O ice while enriching the solution near the solid interface with D.

When the temperature was decreased at a rate of 0.002–0.1 °C/s from an initial temperature between 2 °C and 0.2 °C, the growth front of a D2O crystal in contact with the added liquid H2O advanced, which can be attributed to freezing of liquid near the interface that contained D2O from prior melting of the interface. This new layer must have a composition governed by the phase diagram in Fig. 1. If the temperature was held constant after continuous cooling, this newly crystallized region melted over a few seconds. For example, lowering the temperature from 0.786 °C to 0.682 °C at 0.02 °C/s resulted in rapid growth of ice on the D2O crystal, accompanied by the formation of a distinct interface between the new H2O-rich ice and the D2O ice (Fig. 2 A and B). After the cooling was stopped and the cold stage stabilized at 0.699 °C, the H2O-rich ice melted rapidly at 2.2 μm/s to the original D2O crystal boundary, after which the remaining D2O crystal melted more slowly at 0.25 μm/s, limited by H/D exchange across the interface (Fig. 2 C and D). Above 0.002 °C/s, the growth rate of the ice front increased linearly with increasing cooling rate (Fig. S2). Raman microscopy of the newly H2O-rich ice performed on an area ∼10 μm from the D2O ice front revealed compositions near those expected from the phase diagram in Fig. 1 (Fig. S3).

Fig. 2.

Fig. 2.

(A) A D2O crystal in H2O liquid. (B) The crystal grew after lowering the temperature, revealing a distinct interface between the oval D2O crystal in the center and the surrounding H2O crystal. The dashed oval is a visual aid. (C) After raising the temperature, the newly formed ice retreated. (D) Raising the temperature melted the entire H2O crystal, leaving the D2O crystal with the original shape intact. (Scale bar, 20 μm.)

Fig. S2.

Fig. S2.

Growth rates of D2O ice fronts at various cooling rates.

Fig. S3.

Fig. S3.

(A and B) Comparison of Raman spectra of ice crystals with a known composition (black, data from Fig. S7) with spectra acquired after quenching new-formed ice (red) grown at (A) 1 °C and (B) 1.7 °C.

In another experiment in which the temperature was lowered from 0.745 °C at 0.01 °C/s, the D2O ice interface exhibited an unsteady scalloped morphology (Fig. 3 AE and Movie S1). The cellular features of this interface merged to form a flat ice front upon further cooling (Fig. 3 F and G). Like the crystal in Fig. 2, the solid–solid interface between the D2O crystal and the added H2O ice was readily apparent (Fig. 3G, dashed box). Upon cessation of cooling the ice growth stopped, and the ice front retreated again to the position of the former D2O crystal interface (Fig. 3H). The unsteady cellular features were not observed at cooling rates exceeding 0.02 °C/s (Fig. S4). Collectively, the cellular features in Fig. 3 suggest nonlinear behavior that results from a competition between the timescales for cooling and H/D exchange, the former promoting growth and the latter promoting melting of the ice interface. Both processes require exchange of latent heat. When the driving force (supercooling) is large (Fig. 3 D–G), growth becomes more competitive than H/D exchange, and the advancing growth front is flat, absent of cellular features. Notably, these features were not observed for D2O ice in contact with D2O liquid or H2O ice in contact with H2O liquid; thus any explanation must involve H/D exchange and diffusion as well as thermal effects.

Fig. 3.

Fig. 3.

(A) A single crystal of D2O ice after replacement of liquid D2O with H2O. (B–E) The D2O ice crystal revealed a cellular growth front as the temperature was decreased stepwise by 0.05 °C every 5 s (corresponding to 0.01 °C/s) with the number of cells increasing at first and then merging. (F and G) The ice front flattens and a faceted plane emerged, indicating the basal plane of the crystal (solid arrow, G). The original solid–solid interface is evident within the dashed rectangle. (H) After 36 s at near-constant temperature the newly formed ice completely retreated to the original position of the D2O ice interface. (Scale bar in A, 25 μm.) (Movie S1.)

Fig. S4.

Fig. S4.

(A) The liquid H2O–D2O interface at 1.264 °C. (B) Advancement of the ice growth front after cooling for 6 s at 0.02 °C/s, revealing a flat advancing interface. (C) An advancing ice front, exhibiting a linear facet, which is commonly assigned to the basal (0001) plane (solid arrow), observed at a later stage of growth. (D) Upon cessation of cooling, the newly formed ice melted and the solid–liquid interface retreated to the original location. (Scale bar, 25 μm.)

The interface between D2O and H2O crystals was examined during rapid lowering of the temperature of a D2O crystal in contact with liquid H2O from 0.998 °C to −0.211 °C. The liquid in the channel froze, creating a noticeable discontinuity between the D2O crystal and the new H2O ice (Fig. S5). The interface between the D2O and H2O crystals gradually blurred, becoming imperceptible after ∼20 min (Fig. S5 C–E). The discontinuity suggests a grain boundary between the old (D2O-rich) and the new (H2O-rich) ice, which may reflect a disordered liquid layer that persisted during freezing, possibly akin to quasi-liquid layers invoked for grain boundary melting (25). It is reasonable to suggest that the optical contrast across the width of the discontinuity stems from a different refractive index (η=1.24 if amorphous) compared with hexagonal ice (η=1.3) (26). The disappearance of this region likely reflects restructuring of the disordered region to match the lattice of the solid ice on either side. This process would require molecular motion, and it is reasonable to expect that the timescale would be comparable to that for diffusion. The blurring of the H2O–D2O discontinuity at the solid–solid interface formed as described above suggests a change in structure or composition due to H/D exchange, widening a mixed interface accompanied by the formation of HOD.

Fig. S5.

Fig. S5.

(A) A D2O ice crystal in liquid H2O. (B) After the temperature was lowered rapidly (ca. 0.5 °C over 3 s) a distinct solid–solid interface was formed. (C–E) At constant temperature the boundary remains distinct but begins to blur. (F) After raising the temperature to 0.943 °C, the H2O retreats to the former interface, consistent with the behavior observed in Figs. 2 and 3. (Scale bar in A, 20 μm.)

The changes in the composition of the D2O single ice crystal were measured by Raman microscopy at subfreezing temperatures in the range −0.2 °C to −0.5 °C, which were performed with a precision of ±0.001 °C and spatial resolution of ±0.5 μm. The Raman microscope was focused on the D2O side of the conjoined H2O–D2O ice crystal, 10 μm from the original D2O–H2O interface. The reduction in D2O content and the increase in HOD due to H/D exchange were evident from changes in the intensity and the position of the νO-D bands between 2200 cm−1 and 2600 cm−1 (Fig. S6). Initially, the spectrum revealed νOD stretching modes typical for pure D2O at 2337 cm−1 and 2483 cm−1 (23, 27). Over time, the intensity of the D2O peaks decreased concomitant with a shift of the 2337-cm−1 peak toward 2440 cm−1, the νOD stretching vibration for HOD (28). The amount of H/D was determined by measuring the reduction in the integrated intensity in the region of interest and by using a calibration curve of predetermined H2O:D2O compositions (Materials and Methods and Figs. S6 and S7). The diffusion coefficient of H+ (or H2O) into the D2O ice was determined using a standard semiinfinite slab model (Eq. 1), where Cs is the H2O fraction at the D2O–H2O interface (determined from the phase diagram), C is the H2O fraction at a distance x from the D2O–H2O interface, x is the distance from the D2O–H2O interface where the Raman measurement is positioned (x = 10 ± 0.5 μm), D is the diffusion coefficient, and t is time. A reasonable fit of Eq. 1 to the data afforded D = 0.59 ± 0.1 × 10−10 cm2s−1 (Fig. 4). Notably, the timescale for diffusion of H2O through a typical 3-μm-thick layer using this value of D is ∼25 min, comparable to the time for near disappearance of the interface region in Fig. S5E.

C=Cs[1erf(x2Dt)] [1]

The aforementioned diffusion coefficient compares favorably to the values of D = 1.0 ± 0.2 × 10−10 cm2s−1 reported for D2O diffusion in polycrystalline ice at −1.8 °C and D = 0.1 ± 0.1 ×10−10 cm2s−1 for T2O diffusion in single-crystal ice at −10 °C (18, 29). Using the reported value of D0=10.63 cm2/s measured for T2O diffusion in a single crystal of H2O ice (18), the data in Fig. 4 were consistent with an activation energy Ea = 58.7 ±0.2 kJ/mol. This value is comparable to that reported for T2O in ice, Ea = 59.8 ± 6.7 kJ/mol (18); diffusion of T2O into natural single ice crystals, Ea = 65.7 ± 8.3 kJ/mol (22); and the activation energy for desorption of H2O from 6-μm-thick ice films (58.1 ± 0.8 kJ/mol) (30). In contrast, the diffusion coefficient for HF, which may diffuse as an intact molecule or as ions, in single ice crystals at −10 °C was reported as D = 0.8 ± 0.5 × 10−6 cm2s−1, substantially larger that the values above (31). Collectively, these data implicate diffusion of intact water molecules through a vacancy mechanism for H/D exchange in the D2O crystal (8, 18, 32).

Fig. S6.

Fig. S6.

(A) Representative time-dependent Raman spectra of D2O ice acquired at 10 ± 0.5 μm from the D2O–H2O interface, collected at the time intervals indicated (minutes). The data were background subtracted and baseline corrected, and the integrated intensities between 2200 cm−1 and 2500 cm−1 (dashed lines) were normalized to the first spectrum collected at t=0 min and calibrated using the calibration curve (Fig. S7B). The first and last spectra acquired at 0 min and 540 min are shown as thick solid lines for clarity. (B) The O-D concentration calculated from the spectra in A and four additional Raman measurements. These data points were averaged and binned (Fig. 4).

Fig. S7.

Fig. S7.

(A) Raman spectra of single crystals of ice at various D2O–H2O compositions (%D2O in H2O), measured just below their melting points. The data were background subtracted and baseline corrected, after which the spectra of pure H2O were subtracted from each spectrum. (B) Dependence of the integrated intensity for the various D2O–H2O compositions in A. Each intensity value was normalized to the integrated intensity at 100% D2O. The slope of this plot was used to calibrate the integrated intensity measured during H/D exchange. The difference in the values of the diffusion coefficient with and without this calibration curve was only 5%.

Fig. 4.

Fig. 4.

The decrease of D2O concentration in a D2O crystal due to H/D exchange between the D2O crystal and an adjoining H2O crystal, measured at x = 10 ± 0.5 μm from the H2O–D2O interface. The data were fitted to a semiinfinite model (Eq. 1). Data points represent an average of at least three measurements. Error bars correspond to the SE. See Fig. S6 for details.

Unsteady Interface Morphologies

As described above, when the temperature of the microfluidic channel containing a D2O crystal in contact with liquid H2O was lowered quickly from ca. 2.6 °C to a constant temperature in the range 2.6–0.6 °C, the flat ice front grew, but then retreated, with lower rates of retreat at lower temperatures. Cooling the device from 2.6 °C to below 0.6 °C also resulted in an initial period of growth followed by retreat of the H2O-rich ice front. Once the ice front retreated to the original D2O crystal interface, unsteady morphologies like those in Fig. 3 were observed, growing and melting in an unsteady, almost oscillatory, manner across the interface (Fig. 5 and Movie S2). These unsteady morphologies were not observed for D2O ice in contact with D2O liquid or H2O ice in contact with H2O liquid, arguing against impurities as a major factor in their formation. The interface was decorated with convex features having radii of curvature ranging from 5 μm to 15 μm, alternating with deep concave features having large negative curvature. Under constant system temperature, the convex and concave features cycled every 20–120 s, persisting during a slow retreat of the average position of the D2O crystal interface, which typically required 8 h to traverse the length of the microfluidic chamber (3.6 mm, Movie S3). These unsteady morphologies vanished if the temperature was raised by >0.025 °C (Movie S2). In a few trials, the convex and concave features of the interface coincide in successive cycles to give the appearance of wavelike patterns, denoted by the vertical dashed lines in Fig. 5. Most experiments, however, reveal an unsteady cycling of the interface morphology, usually with new convex features emerging out of regions of sharp negative curvature where shrinking convex features intersect the interface. The evolution of the curved features was not a consequence of crystal anisotropy (Fig. S8).

Fig. 5.

Fig. 5.

Unsteady morphologies oscillating between positive (convex) and negative (concave) curvature at the solid–liquid interface during melting of a D2O ice crystal in liquid H2O under isothermal conditions (0.506 °C). In this example, the unsteady morphologies oscillated between convex and concave at fixed positions throughout, as denoted by the dashed lines in the corresponding panels on the right. The dashed lines intersect regions of negative curvature in A and E, and regions of positive curvature in C and G. Bottom Center panel illustrates the relationship between the interface temperatures and the bulk melting temperature as well as the gradients of D and H in the solid and liquid phases. The solid–liquid interface here is defined by x = 0. Local thermal gradients due to local nonisothermal conditions accompanying melting and freezing of the convex and concave features also are expected (not shown).

Fig. S8.

Fig. S8.

Convex and concave features (A and D) were observed regardless of the growth direction of the D2O ice crystal, as indicated in C and B.

These observations suggest that the growth and dynamics of the unsteady morphologies depend on a complex set of cooperative phenomena as described below, including H/D exchange from the H2O-rich liquid into the D2O-rich ice crystal, latent heat exchange, local thermal gradients, and the Gibbs–Thomson effect on the melting points of the convex and concave features. The Gibbs–Thomson effect (33) predicts that the melting temperature of a curved interface, Ti, can be calculated using Eq. 2, where Tm is the melting temperature of the bulk ice (which is nearly equal to the temperature measured by the thermistor mounted in a copper plate beneath the microfluidic device; Fig. S1), γ is the energy of the solid–liquid interface, κ is the curvature or reciprocal radius of the curved feature, and L is the latent heat of ice). The κ term is positive when the feature is convex (34). Convex features (positive curvature) are susceptible to melting because Ti>Tm, whereas concave features (negative curvature) can promote ice growth because Ti<Tm. Using the range of curvature radii observed here, a surface tension γ=0.033 Jm−2 and L = 3.34 × 10−8 Jm−3, the magnitude of the Ti values for the convex and concave features is anticipated to be in the range 0.001 ≤ |Ti| ≤ 0.005 °C. This small difference is within the precision of temperature control and measurement in the microfluidic device (as measured at the thermistor), which is viewed as essential for the observation of the unsteady morphologies that otherwise would go unnoticed.

TmTi=1(γκL) [2]

Considering the above factors, a preliminary framework for understanding the growth and dynamics of the unsteady interface morphology was constructed. Ice grows initially upon cooling because of the presence of D2O near the interface, which diffuses into the liquid phase, creating a supercooled condition. The ice then retreats when H2O diffuses into the solid phase, inducing melting as Tm decreases. During this growth and retreat, the cellular features form on the interface, possibly because of the nonequilibrium conditions created by the requirement for effective exchange of latent heat during freezing and melting following the temperature step. Once the unsteady morphologies appear, melting of the convex features, aided by the Gibbs–Thomson effect, enriches the concentration of D2O above the concave region, creating again a supercooled condition that favors freezing in this region. Conversely, the melting point of the adjacent concave feature is lower than Tm, which would induce ice growth in this region. This growth would consume the D2O molecules in the liquid above it due to preferential freezing of D2O at the lower temperature. Heat transfer between the convex and concave regions may then occur, aided by the difference in thermal diffusivities. The thermal diffusivity of ice (1.50 × 10−2 cm2s−1) is 10 times greater than that of liquid water or poly(dimethylsiloxane) (PDMS) (1.50 × 10−3 cm2s−1 and 1.00 × 10−3 cm2s−1, respectively) and 3 times greater than the glass bottom of the microfluidic chamber (5.50 × 10−3 cm2s−1).

Latent heat absorbed by melting of the convex features could therefore lower the temperature and hence supercool the concave interface. Considering the latent heat of fusion and the heat capacity of solid ice, and assuming an idealized adiabatic condition, a convex feature modeled as a hemicylinder with a 5-μm radius and a 20-μm height absorbs latent heat that would lower the temperature of an adjacent hemicylinder of equal volume in the adjacent solid region by 0.16 °C. Notably, local irradiation of a region of the unsteady ice interface at a constant system temperature with a 980-nm IR laser resulted in ice growth at neighboring (not irradiated) regions (Movie S4), which is consistent with the release of D2O into the liquid near the ice interface upon melting and the supercooling effect. Moreover, the expected content of D2O at x = 5 μm and t = 20 s, based on the diffusivity in the liquid (D = 10−5 cm2s−1), is consistent with the composition of mixed ice expected at the constant system temperature (Fig. 1).

This heuristic framework demands a more quantitative understanding of the growth and dynamics of the unsteady morphologies. A typical starting point is a linear stability analysis of the ice front. It is well known that a flat interface may be linearly unstable under certain conditions, notably when subject to undercooling in the liquid, which gives rise to the well-known Mullins–Sekerka instability (34, 35). The observations here cannot be explained by a Mullins–Sekerka instability, which would predict a melting front to be stable. Moreover, nonplanar morphologies for H2O ice contacting H2O liquid or D2O ice contacting D2O liquid were never observed. Therefore, both temperature and concentration diffusion must be included in a linear stability analysis, but such an analysis appears to be analytically intractable here.

Consequently, numerical simulations were attempted to validate pieces of the heuristic framework and to explore the values of parameters that might give rise to unsteady morphologies. Phase-field simulations (36) that included thermal diffusion, concentration diffusion, and surface energy effects (Supporting Information) were performed, allowing for different diffusion constants in the solid and liquid phases and a concentration-dependent melting temperature (with respect to the amount of H in D2O ice). Numerical limitations prevented simulations using the measured diffusivities or melting temperatures; therefore, simulations were aimed toward capturing the dynamics qualitatively. The initial configuration for the simulations included pure D2O solid at one end of the domain separated by a sharp, possibly curved interface from a domain of pure liquid H2O. The temperature at both ends of the domain was fixed at the same constant value (see Table S1 for the parameter values used in the phasefield simulations).

Table S1.

Parameter values used in the simulations shown in Figs. S9 and S10

Parameter Fig. S10 Fig. S9
a0 1.2 1.38
b0 0.4a0 = 0.48 0.4a0 = 0.55
ϵΦ 0.03 0.03
L0 0.2 0.2
J1 10 20
J2 1 4
DS(T) 0.2 0.05
DL(T) 0.02 0.02
DS(C) 0.02 0.01
DL(C) 0.2 0.4

Simulations of a flat interface upon imposing a constant temperature between the Tm values for D2O and H2O usually exhibited either melting or freezing, but with a transition region at intermediate temperatures where the interface froze initially and then melted (Fig. S9). This change in direction was due to a high initial flux of D2O into the liquid phase, creating a supercooled condition that causes the interface to freeze, followed by a slower flux of H2O into the solid phase, lowering the melting temperature and causing melting. The timescales and temperature profile in the simulations differed from those in the experiment, which also exhibited freezing followed by melting. Nevertheless, the simulations support the premise that the direction of interface motion reflects the difference in the timescales of the fluxes of D2O and H2O between liquid and solid and the subsequent coupling to the melting temperature.

Fig. S9.

Fig. S9.

A simulation that began with pure D2O in the solid phase (y<0.5) and pure H2O in the liquid phase (y>0.5) and evolved at constant temperature near the transition between overall melting and freezing. (A) Concentration of H2O as a function of y at various times. The location of the interface (contour ϕ=0.5) is shown as a red cross in B. (B) Speed of the interface as a function of time. The dynamics are as follows: D2O initially spills rapidly into the liquid phase (B, orange curve), raising the melting temperature, creating a supercooled condition and causing the interface to freeze (A, positive speed). H2O then diffuses into the solid phase (B, yellow and purple curves), lowering the melting temperature and causing the interface to melt (A, negative speed.) The initial condition was T(x,y,0)=1, ϕ(x,y,0)=c(x,y,0)=1/(1+e16(yLy/2)) so the interface was initially flat and stayed flat thereafter. Other parameters for the simulations are shown in Table S1.

The evolution of an interface that began with a nonflat (sinusoidal) shape also was explored. Simulations clearly reveal that the temperature of a convex feature is higher than the temperature of the surroundings (i.e., the constant system temperature), which provokes its disappearance through melting. Conversely, the adjacent concave feature is cooler, which will provoke freezing in this region and a lower D2O concentration above it due to preferential freezing of D2O at the lower temperature (Fig. S10B). This preliminary modeling did not find evidence of a linear instability, however, and it could not replicate the proposed feedback between the convex and concave features. Instead, all interfaces eventually decayed to a flat morphology, and the locations of the convex and concave features never moved (Figs. S10C and S11B, for example.)

Fig. S10.

Fig. S10.

(A) Initial conditions for phase, temperature, and concentration. The solid phase is at y=0 and the liquid phase is at y=Ly=3. (B) The same fields at nondimensional time 0.4. The concentration of H2O is lower in the concave region, indicating a higher concentration of D2O there. The temperature is cooler on the liquid side of the interface and hottest in the solid where the convex feature is melting. (C) Location of the interface (contour ϕ=0.5) as a function of horizontal coordinate x, at different times shown in the key. A sinusoidal interface decays to nearly flat by time 0.9. Nondimensional parameters for this simulation are shown in Table S1.

Fig. S11.

Fig. S11.

This is similar to Fig. S10, but for the model in SI Numerical Phase-Field Model, Alternative Phase-Field Model. (A) Phase, temperature, and concentration fields at nondimensional time 0.4. Phase and concentration are indistinguishable from the original model, but temperature differs slightly; notably the convex region is colder than the concave region. (B) Location of the interface (contour ϕ=0.5) as a function of horizontal coordinate x, at different times shown in the key. The initial condition was the same as in Fig. S10, so is not shown. Nondimensional parameters for this simulation are shown in Table S1.

These preliminary numerical investigations are limited by numerous factors—most notably grid resolution, numerical stiffness that makes use of the experimental parameter values impractical, and a 10-dimensional parameter space that has not yet been examined exhaustively. The discrepancy between the numerical results and experimental observations also may reflect other, as yet unidentified physics. Perhaps defects that form during the continuous oscillatory-like cycling of melting and freezing at the interface play a role; or perhaps the morphologies are oscillating between nonplanar, quasi-steady ice fronts, which are known to exist in simpler systems (34). Although the modeling reported here was numerical, it would be useful to explore whether there is a parameter regime in which the planar interface is linearly unstable, and it still seems possible that improved theory could predict unsteady dynamics like those seen in the experiments (37).

Collectively, the observations described herein reveal that ice crystallization can still surprise, generating behavior that cannot be explained readily by nonlinear models. Liquid exchange around a single crystal of ice in a microfluidic device, combined with highly precise temperature control, created conditions in which the competing influences of heat transfer, surface curvature, and H/D exchange afforded unusual and persistent features that melted and froze cyclically on the ice interface as it retreated due to melting. Preliminary numerical simulations captured the freezing and melting of a flat interface and the relationship between thermal and concentration fluxes at the interface, as well as a dependence of the fluxes on surface curvature. Further modeling is required, however, to capture the sensitive coupling and feedback between temperature and concentration fluxes, as well as other physical factors, which produce the oscillatory behavior. Although oscillating growth fronts have been reported during condensation of water on thin metal films that imposed a temperature gradient (38), to our knowledge, the unsteady morphologies observed here have not been observed previously on ice interfaces, and they appear to be unique to D2O solid-H2O liquid interfaces. The behavior is somewhat reminiscent of oscillations observed during vapor–liquid–solid growth of sapphire nanowires, however (39). Moreover, the microfluidic device configuration enabled measurement of self-diffusion of water in single crystals of ice, using Raman microscopy. The methodology described here promises utility for investigations of other crystalline materials, and we anticipate it can lead to further discoveries in ice crystallization, including the effects of ice crystallization inhibitors.

Materials and Methods

Microfluidic Solution Exchange Experiments.

A temperature-controlled cold stage that was previously described in ref. 24 was placed on an inverted microscope (DMIRE2; Leica Microsystems Inc.). The temperature of the cold stage could be adjusted with a precision of ±0.001 °C and stability of ±0.002 °C (Fig. S1). A sCMOS (Zyla 5.5; Andor) camera was used to acquire images. A sapphire disc (2.5 cm diameter) was placed between the cold stage and the microfluidic device to minimize temperature gradients. Immersion oil was introduced to the surface of the sapphire disc, and the microfluidic chip was then placed on top of the sapphire disc. The microfluidic channels were filled with 99.96% D2O (Cambridge Isotopic Laboratories) and the temperature decreased to ca. −20 °C to freeze the D2O in the channel. The temperature was then increased to melt the ice blocking the inlet and outlet and to form a D2O single ice crystal, with a distinct boundary. The D2O liquid in the channels was then exchanged with doubly distilled H2O to create a D2O–H2O solid–liquid interface. A 980-nm IR laser (Wuhan Laserlands Laser Equipment Co.) was used for local ice melting.

Raman Shift Measurements.

The H/D composition of ice crystals and H2O/H+ diffusion into D2O ice were measured with a Raman microscope (DXR Raman microscope; Thermo Fisher Scientific), using a 532-nm excitation laser operating at 10 mW, with a 2-cm−1 resolution and slit width of 50 μm. Due to the upright configuration of the microscope, the cold stage was mounted upside down so that the channel was accessible to the Raman excitation beam. The exchange of D2O with H2O was performed in a manner similar to that described above. After the D2O–H2O solid–liquid interface was created, the temperature was decreased to −0.2 °C to −0.5 °C to freeze the liquid H2O in the channel. Raman measurements were collected on a region of the D2O ice crystal located 10 ± 0.5 μm from the original D2O–H2O solid–solid interface. Data were collected at regular intervals using an exposure time of 15 s, and the final intensity for each interval was calculated from the sum of 15 measurements. The Raman spectra of the time-lapse measurements were background subtracted and baseline corrected before determination of the total intensity. The intensities used to evaluate the diffusion coefficient were calculated from the integral of the intensity between 2200 cm−1 and 2500 cm−1, the region that captures the bands attributable to νO-D for both D2O and HOD (Fig. S6). The intensity data were corrected using a calibration curve for various compositions of D2O:H2O mixtures (Fig. S7), although this did not substantially affect the fit of the data or the calculated diffusion constant (Eq. 1).

SI Materials and Methods

The protocol for the microfluidic device fabrication was described previously (24). The microfluidic device was designed using AutoCAD2015 (Autodesk, Inc.), and transparency films (Artnetpro) were used as masks. Molds for fabricating the devices were created by spin-coating (Cost Effective Equipment, Model 100, Rolla, MO) SU8-2025 on a silicon wafer (3-inch diameter; University Wafer) at 500 rpm for 10 s and then 3,000 rpm for 60 s. The wafers were baked at 65 °C for 2 min and 95 °C for 5 min, followed by UV exposure for 25 s through a mask, using an MJB3 contact mask aligner (Karl Suss). After a postexposure bake at 65 °C for 1 min and then 95 °C for 3 min, the mold was treated with SU8 developer (Microchem) for 4.5 min and then washed with isopropanol and water. Microfluidic chips were fabricated by carefully pouring a degassed 10:1 mixture of PDMS elastomer (SYLGARD 184; Dow Corning) and curing agent on the mold, followed by baking at 80 °C for 60 min. After the PDMS layer solidified, it was removed from the mold and inlet and outlet holes were punched. A glass coverslip and PDMS microfluidic chip were cleaned for 50 s with an oxygen plasma cleaner (model PDC-001; Harrick Plasma Cleaner) and then bonded together by placing the PDMS chip on top of the coverslip.

SI Raman Measurements Near Ice Growth Fronts

The compositions of newly formed ice (upon cooling) near the liquid–solid interface were measured using Raman microscopy. A typical growth sequence lasted on the order of minutes, and H/D exchange caused retreat of the ice front within minutes if the cooling rate was slow or if cooling was stopped at T>0°C. This timescale is not sufficiently long for Raman measurements with sufficient signal-to-noise ratio. Therefore, Raman spectra were acquired after quenching the system to T = −0.1 °C immediately after the ice front retreated to the original D2O boundary, freezing all liquid in the channel so that a Raman spectrum of the region of the previously frozen H2O-rich ice could be acquired. As expected from the phase diagram (Fig. 1), the Raman spectra acquired after quenching from 1 °C compared favorably with that obtained for ice containing 25% D2O whereas the Raman spectra acquired after quenching from 1.7 °C compared favorably with that obtained for ice containing 50% D2O in H2O (Fig. S3). These results substantiate that Raman spectra acquired near the interface accurately reflect the expected thermodynamic compositions.

SI Numerical Phase-Field Model

The direct numerical simulation of complex geometries created by phase transformations is a challenging problem with a long history. One flexible and widely used approach is the phase-field method, which uses a diffuse-interface approximation of the liquid–solid interface (36). Designing and implementing a phase-field model in the present setting requires making numerous modeling choices. A preliminary model is presented in this section.

Model.

The goal of this model is to simulate the joint evolution of temperature T(x,y,t), concentration c(x,y,t) of H2O, and an order parameter (hereafter “phase”) ϕ(x,y,t). Here t[0,) is time, and (x,y)[0,Lx]×[0,Ly] represents the spatial variables, with x being the horizontal, along-interface direction, and y varying from solid at y=0 to liquid at y=Ly. The phase is a continuous function that takes the value ϕ(x,y)=1 in the solid, ϕ(x,y)=0 in the liquid, and ϕ(x,y)(0,1) in the transition region between solid and liquid. As the width of the transition region tends to zero, the model captures the well-known characteristics of phase interfaces, for example the importance of curvature through the Gibbs–Thomson effect. The free energy of the system is assumed to be of the form

F[ϕ,c,T]=(12|HWϕϕ|2+f(ϕ,c,T))dx. [S1]

The first term in the integrand, which is related to the surface energy of the interface, depends on parameters Wϕ governing the width of the interface and H governing the height of the energy barrier between the phases. The second term in the integrand is the bulk free energy, which is modeled as

f(ϕ,c,T)=Hg(ϕ)+cp2Tsys(TTm(c))2+R0(cc)2. [S2]

The first term in the bulk free energy models the free energy as a function of phase. Consistent with the literature on phase-field modeling (36) it chosen to be a double-well potential g(ϕ)=ϕ2(1ϕ)2, with a barrier of height H between the wells. The second term in Eq. S2 is a quadratic approximation of the free energy about the melting temperature Tm(c). It depends on the heat capacity cp, assumed to be independent of concentration, and the system temperature Tsys. Although the factor in the denominator is usually Tm, the error in replacing it by Tsys is presumed to be small enough to be neglected. The melting temperature is taken to be an affine function of concentration as suggested by the experimental measurements (Fig. 1)

Tm(c)=abc, [S3]

where a is the melting temperature of pure D2O, and ab is the melting temperature of pure H2O.

The final term in Eq. S2 measures the entropy of mixing. It depends on parameter R0=RTsys, where R is the gas constant, and on the concentration c to which the system relaxes; the numerical value of c is immaterial in this model. Note that the true mixing entropy is RT((1c)log(1c)+clogc); here again there is a quadratic approximation, ignoring any higher-order nonlinear effects.

Next consider the dynamical evolution of ϕ, T, and c. It is important to remember that there are two conserved quantities: mass and (thermal) energy. As mass is conserved, the evolution of c should have the form ct+Jc=0, where Jc is a suitably defined vector of mass flux. The analogous calculation for energy involves the enthalpy, which is represented by

γ=TLcpP(ϕ), [S4]

where L is the latent heat (assumed to be independent of concentration) and P(ϕ)=(32ϕ)ϕ2 is an interpolation function, which increases monotonically from P(0)=0 to P(1)=1. To ensure conservation of energy, the evolution of γ should have the form γt+Jγ=0, where Jγ is a suitably defined vector of energy flux. The phase ϕ is of course not conserved; its evolution should have the property that the free energy decreases. With these considerations in mind, it is useful to view the free energy as a function of ϕ, c, and γ,

F[ϕ,c,γ]=(12|HWϕϕ|2+f(ϕ,c,γ))dx, [S5]

where the bulk term is now

f(ϕ,c,γ)=Hg(ϕ)+cp2Tsys(γTm(c)+LcpP(ϕ))2+R0(cc)2. [S6]

Since P(0)=P(1)=1, the function ϕf(ϕ,c,γ) has local minima at ϕ=0 and ϕ=1 for all c and γ.

The evolution law can now be specified: It is

τϕt=1HδFδϕ [S7]
γt=(MT(ϕ)δFδγ) [S8]
ct=(MC(ϕ)δFδc). [S9]

Here τ is a timescale, which must ultimately be related to the small parameter Wϕ since it should be chosen so that the interface moves with a speed of order 1. The functions MT(ϕ), MC(ϕ) are the temperature and concentration mobilities (diffusivities), respectively, which are assumed to depend on phase via ϕ. The logic here is familiar: ϕ does steepest-descent dynamics for the free energy F, while for the conserved variables γ and c the associated fluxes are proportional to the gradients of the “chemical potentials” δFδγ and δFδc.

To be more explicit, after absorbing some of the constants into the diffusivities, the evolution equations are

τϕt=Wϕ22ϕg(ϕ)LHTsys(γTm(c)+LcpP(ϕ))P(ϕ) [S10]
γt=(MT(ϕ)(γTm(c)+LcpP(ϕ))) [S11]
ct=(MC(ϕ)(c+bcpR0Tsys(γTm(c)+LcpP(ϕ)))). [S12]

It is convenient to nondimensionalize the model S10S12 above, using the following characteristic scales:

  • Length Lx = horizontal width of container

  • Temperature Tsys

  • Time τ.

The choice of time nondimensionalization does not come from any physical characteristic timescale, but rather is used for convenience because time then enters only in the choice of mobilities. The nondimensional variables are

γ=γTsys,t=tτ,x=xLx,a=aTsys,b=bTsys,α=τLx2α,MT(c)=τLx2MT(c).

Substituting into [S10S12] and removing all of the tildes affords the following equations:

ϕt=ϵ2ϕϕg(ϕ)J1(γa0+b0c+L0P(ϕ))P(ϕ) [S13]
γt=(MT(ϕ)(γ+b0c+L0P(ϕ))) [S14]
ct=(MC(ϕ)(c+J2(γ+b0c+L0P(ϕ)))). [S15]

Eqs. S13S15 form the model that is simulated. There are six nondimensional parameters:

ϵϕ = Wϕ2Lx2

L0 = LcpTsys

J1 = LH

J2 = bcpR0

a0 = a/Tsys

b0 = b/Tsys.

In addition, the modeler must prescribe the mobilities. Since the aim is to model the interplay between reduced concentration diffusivity and increased temperature diffusivity in the solid, the simulated model uses mobilities which interpolate continuously between constant diffusivities in the solid, DS(T),DS(C) for temperature and concentration, respectively, and constant diffusivities in the liquid, DL(T),DL(C), using P(ϕ) as the interpolation function:

MT(ϕ)=DS(T)P(ϕ)+DL(T)(1P(ϕ)),MC(ϕ)=DS(C)P(ϕ)+DL(C)(1P(ϕ)). [S16]

Including the mobilities, the simulated model contains a total of 10 nondimensional parameters. Finally, the model needs boundary conditions, which are chosen as follows:

  • In the x direction: periodic in all variables.

  • Concentration: c=0 at y=0, c=1 at y=Ly.

  • Phase: ϕ=1 at y=0, ϕ=0 at y=Ly.

  • Temperature: γ=1L0 at y=0, γ=1 at y=Ly. This comes from the boundary condition T=Tsys at y=0,Ly, combined with the boundary conditions for ϕ.

Model in the temperature variable.

The model S13S15 expressed in terms of T is

ϕt=ϵ2ϕϕg(ϕ)J1(TTm(c))P(ϕ) [S17]
Tt=(MT(ϕ)(TTm(c)))+L0P(ϕ)ϕt [S18]
ct=(MC(ϕ)(c+J2(TTm(c)))). [S19]

A rough description of the dynamics is as follows: (i) The phase evolves (melts or freezes) in places near the front where the local temperature differs from the local melting temperature; (ii) when freezing (melting) happens, latent heat is released (absorbed), causing a local increase (decrease) of temperature on the side of the front that is changing phase; (iii) meanwhile both temperature and concentration diffuse to make the local temperature equal the melting temperature.

Alternative Phase-Field Model.

A slightly different model was explored wherein the temperature and concentration diffuse independently. This model in nondimensional form is

ϕt=ϵ2ϕϕg(ϕ)J1(TTm(c))P(ϕ) [S20]
Tt=(MT(ϕ)T)+L0P(ϕ)ϕt [S21]
ct=(MC(ϕ)c). [S22]

By defining γ as in Eq. S11 and rewriting Eq. S21 as an equation for γt, one gets an equation equivalent to Eq. S21 with no time derivative on the right-hand side.

Preliminary simulations with this model produced results that differed in small ways from the first model, but with no significant qualitative differences. Notably, no oscillations, instabilities, or nonlinear growth patterns were observed.

Numerical Scheme.

Eqs. S13S15 were solved semiimplicitly, using finite differences to calculate derivatives on the right-hand sides of Eqs. S14 and S15, and treating the Laplacian ϵ2ϕϕ in Eq. S13 implicitly in Fourier space.

We first construct an equally spaced grid as

x=(Δx,2Δx,,NxΔx),y=(Δy,2Δy,,NyΔy).

The grid spacings Δx, Δy and number of points Nx, Ny are related by NxΔx=Lx (= 1), (Ny+1)Δy=Ly. The relationship is different for the x and y variables because we have periodic boundary conditions in x, so we assume u(0,y)=u(Lx,y) is assumed for any field u (and similarly for any derivatives of u), but there are Dirichlet boundary conditions in y so u(x,0),u(x,Ly) are fixed, known values.

A fixed time step Δt was chosen as Δt=μmin((Δx)2,(Δy)2)min(DS(C),DS(T),DL(C),DL(T)), where μ is the Courant number, set to μ=0.4. The fields γ,c are updated as

un+1=un+Δt(Aunjun+hun), [S23]

where un is either γ or c at the nth time step, jun is the quantity in the innermost brackets [either γ+b0c+L0P(ϕ) for γ or c+J2(γ+b0c+L0P(ϕ)) for c] at time-step n, Aun is the numerical discretization of the operator (Mu(ϕ)) and time-step n, and hun is any extra contribution from the boundary terms. In the numerical implementation, un, jun, and hun are vectors of length Np=Nx×Ny, and Aun is an Np×Np sparse matrix.

The operator (Mu(ϕ)) was discretized using a second-order scheme based on finite differences (40). For a given field v(x,y) evaluated at grid points vi=v(iΔx,y), the x component of the operator was discretized as

x(Mu(ϕ)xv)Mu(ϕi1/2)vi1(Mu(ϕi1/2)+Mu(ϕi+1/2))vi+Mu(ϕi+1/2)vi+1 [S24]

and similarly for the y component of the operator. Fields were evaluated at half grid points, using the usual averaging operator ui+1/2=(ui+ui+1)/2 and similarly for ui1/2. When a particular index extends beyond the set {1,2,,Nx(y)}, a boundary condition is substituted: uNx+1=u1 for the x part of the operator and the known values of u0, uNy+1 for the y part of the operator. This implies including a vector of boundary values hun for the y operator, since Eq. S24 will be an affine function of v, not a linear function.

The field ϕ is updated as

(1ϵϕΔt2)ϕn+1=ϕnΔt(g(ϕn)+J1(γna0+b0cn+L0P(ϕn))). [S25]

The right-hand side rn is calculated explicitly at each grid point. Its fast Fourier transform r^n is then calculated, so ϕ^n+1 can be solved for in Fourier space, and finally the inverse Fourier transform gives ϕn+1.

The fast Fourier transform works best for periodic functions. To account for the Dirichlet boundary conditions in y, all functions were shifted by an affine function of y so the boundary conditions are 0 on both sides of the domain. The resulting function was reflected antisymmetrically about y=Ly, and the final transformed function is denoted with a tilde. For example, the right-hand side is transformed as

rn(x,y)={rn(x,y)(1yLy)y[0,Ly)rn(x,Lyy)+(1LyyLy)y[Ly,2Ly). [S26]

The transformed right-hand side rn(x,y) is used to solve for the transformed phase ϕn+1(x,y) (y[0,2Ly)), and the transformations are then reversed to obtain ϕn+1. This transformation ensures the phase and its derivative are continuous functions on the interval [0,2Ly) with periodic boundary conditions, and therefore derivatives calculated in Fourier space are better behaved. The transformation does not ensure the second derivative of ϕn is continuous over the periodic interval, however, so calculating the Laplacian induces Gibbs phenomena (oscillations on the scale of the grid). These oscillations can be made negligible by choosing an initial condition whose second derivative is nearly zero for y{0,Ly}.

Examples.

A variety of parameter choices were explored, and results from two simulations are illustrated here. Both simulations had numerical parameters Lx=1, Ly=3, Ny=61, and Nx=24. The nondimensional parameters are provided in Table S1.

Supplementary Material

Supplementary File
Download video file (55.2MB, mp4)
Supplementary File
Download video file (20.3MB, mp4)
Supplementary File
Download video file (7MB, mp4)
Supplementary File
Download video file (28.8MB, mp4)

Acknowledgments

The authors thank Victor Yashunsky for his assistance with the design of the cold stage and the temperature-control program and Alexander Shtukenberg and Takuji Adachi for helpful discussions. This work was supported primarily by the New York University Materials Research Science and Engineering Center Program of the National Science Foundation (NSF) under Award DMR-1420073 and NSF Grant DMS-1311833. M.H.-C. was partially supported by Department of Energy Grant DE-SC0012296.

Footnotes

The authors declare no conflict of interest.

This article is a PNAS Direct Submission.

This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1621058114/-/DCSupplemental.

References

  • 1.Falenty A, Hansen TC, Kuhs WF. Formation and properties of ice XVI obtained by emptying a type sII clathrate hydrate. Nature. 2014;516:231–233. doi: 10.1038/nature14014. [DOI] [PubMed] [Google Scholar]
  • 2.Huang YY, et al. A new phase diagram of water under negative pressure: The rise of the lowest-density clathrate s-III. Sci Adv. 2016;2:e1501010. doi: 10.1126/sciadv.1501010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Liu Z, Hansen W. Freeze–thaw durability of high strength concrete under deicer salt exposure. Construct Build Mater. 2016;102:478–485. [Google Scholar]
  • 4.Vico G, Hurry V, Weih M. Snowed in for survival: Quantifying the risk of winter damage to overwintering field crops in northern temperate latitudes. Agric Forest Meteorol. 2014;197:65–75. [Google Scholar]
  • 5.Prior DJ, et al. Insight into the phase transformations between ice Ih and ice II from electron backscatter diffraction data. Scripta Mater. 2012;66:69–72. [Google Scholar]
  • 6.Lewis JK, et al. The grand challenges of organ banking: Proceedings from the first global summit on complex tissue cryopreservation. Cryobiology. 2016;72:169–182. doi: 10.1016/j.cryobiol.2015.12.001. [DOI] [PubMed] [Google Scholar]
  • 7.Deville S, Viazzi C, Guizard C. Ice-structuring mechanism for Zirconium acetate. Langmuir. 2012;28:14892–14898. doi: 10.1021/la302275d. [DOI] [PubMed] [Google Scholar]
  • 8.Hobbs PV. Ice Physics. Clarendon; Oxford: 1974. [Google Scholar]
  • 9.Wakayama K, et al. Successful transplantation of rat hearts subjected to extended cold preservation with a novel preservation solution. Transpl Int. 2012;25:696–706. doi: 10.1111/j.1432-2277.2012.01469.x. [DOI] [PubMed] [Google Scholar]
  • 10.Epstein S, Sharp RP, Gow AJ. Antarctic ice sheet - stable isotope analyses of Byrd station cores and interhemispheric climatic implications. Science. 1970;168:1570–1572. doi: 10.1126/science.168.3939.1570. [DOI] [PubMed] [Google Scholar]
  • 11.Tranter M. Encyclopedia of Earth Sciences Series. Springer Nature; New York: 2014. Isotopic fractionation of freezing water; pp. 668–669. [Google Scholar]
  • 12.Jouzel J, et al. More than 200 meters of lake ice above subglacial lake Vostok, Antarctica. Science. 1999;286:2138–2141. doi: 10.1126/science.286.5447.2138. [DOI] [PubMed] [Google Scholar]
  • 13.Everest MA, Pursell CJ. Isotope exchange of D2O on H2O ice: Surface versus bulk reactivity. J Chem Phys. 2001;115:9843–9847. [Google Scholar]
  • 14.Kang H. Chemistry of ice surfaces. Elementary reaction steps on ice studied by reactive ion scattering. Acc Chem Res. 2005;38:893–900. doi: 10.1021/ar0501471. [DOI] [PubMed] [Google Scholar]
  • 15.Kim JH, Shin T, Jung KH, Kang H. Direct observation of segregation of sodium and chloride ions at an ice surface. Chemphyschem. 2005;6:440–444. doi: 10.1002/cphc.200400429. [DOI] [PubMed] [Google Scholar]
  • 16.Lamberts T, Ioppolo S, Cuppen HM, Fedoseev G, Linnartz H. Thermal H/D exchange in polar ice - deuteron scrambling in space. Mon Not R Astron Soc. 2015;448:3820–3828. [Google Scholar]
  • 17.Oxley SP, Zahn CM, Pursell CJ. Diffusion of HDO in pure and acid-doped ice films. J Phys Chem A. 2006;110:11064–11073. doi: 10.1021/jp062270v. [DOI] [PubMed] [Google Scholar]
  • 18.Ramseier RO. Self-diffusion of tritium in natural and synthetic ice monocrystals. J Appl Phys. 1967;38:2553–2556. [Google Scholar]
  • 19.Livingston FE, George SM. Effect of HNO3 and HCl on HDO diffusion on crystalline D2O ice multilayers. J Phys Chem B. 1999;103:4366–4376. [Google Scholar]
  • 20.Itagaki K. Self-diffusion in single crystals of ice. J Phys Soc Jpn. 1964;19:1081. [Google Scholar]
  • 21.Perakis F, Borek JA, Hamm P. Three-dimensional infrared spectroscopy of isotope-diluted ice Ih. J Chem Phys. 2013;139:014501. doi: 10.1063/1.4812216. [DOI] [PubMed] [Google Scholar]
  • 22.Scherer JR, Snyder RG. Raman intensities of single crystal ice Ih. J Chem Phys. 1977;67:4794–4811. [Google Scholar]
  • 23.Iglev H, Schmeisser M, Simeonidis K, Thaller A, Laubereau A. Ultrafast superheating and melting of bulk ice. Nature. 2006;439:183–186. doi: 10.1038/nature04415. [DOI] [PubMed] [Google Scholar]
  • 24.Drori R, et al. A supramolecular ice growth inhibitor. J Am Chem Soc. 2016;138:13396–13401. doi: 10.1021/jacs.6b08267. [DOI] [PubMed] [Google Scholar]
  • 25.Dash JG, Rempel AW, Wettlaufer JS. The physics of premelted ice and its geophysical consequences. Rev Mod Phys. 2006;78:695–741. [Google Scholar]
  • 26.Berland BS, Brown DE, Tolbert MA, George SM. Refractive index and density of vapor-deposited ice. Geophys Res Lett. 1995;22:3493–3496. [Google Scholar]
  • 27.Shi L, Gruenbaum SM, Skinner JL. Interpretation of IR and Raman line shapes for H2O and D2O ice Ih. J Phys Chem B. 2012;116:13821–13830. doi: 10.1021/jp3059239. [DOI] [PubMed] [Google Scholar]
  • 28.Scherer JR, Go MK, Kint S. Raman spectra and structure of water from -10 to 90.deg. J Phys Chem. 1974;78:1304–1313. [Google Scholar]
  • 29.Kuhn W, Thurkauf M. Separation of isotopes on freezing of water and the diffusion coefficients of D and O18 in ice. Helv Chim Acta. 1958;41:938–971. [Google Scholar]
  • 30.Smith JA, Livingston FE, George SM. Isothermal desorption kinetics of crystalline H2O, (H2O)-O-18, and D2O ice multilayers. J Phys Chem B. 2003;107:3871–3877. [Google Scholar]
  • 31.Kopp M, Barnaal DE, Lowe IJ. Measurement by Nmr of diffusion rate of Hf in ice. J Chem Phys. 1965;43:2965–2971. [Google Scholar]
  • 32.Noguchi N, Kubo T, Durham WB, Kagi H, Shimizu I. Self-diffusion of polycrystalline ice I under confining pressure: Hydrogen isotope analysis using 2-D Raman imaging. Phys Earth Planet Int. 2016;257:40–47. [Google Scholar]
  • 33.Thomson W. On the equilibrium of vapour at a curved surface of liquid. Philos Mag. 1871;42:448–452. [Google Scholar]
  • 34.Langer JS. Instabilities and pattern-formation in crystal-growth. Rev Mod Phys. 1980;52:1–28. [Google Scholar]
  • 35.Mullins WW, Sekerka RF. Stability of a planar interface during solidification of a dilute binary alloy. J Appl Phys. 1964;35:444–451. [Google Scholar]
  • 36.Provatas N, Elder K. Phase-Field Methods in Materials Science and Engineering. Wiley; Hoboken, NJ: 2011. [Google Scholar]
  • 37.Turing AM. The chemical basis of morphogenesis. Philos Trans R Soc Lond B Biol Sci. 1952;237:37–72. doi: 10.1098/rstb.2014.0218. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Myagkov VG. Oscillations of the crystallization front of adsorbed water. JETP Lett. 2000;72:4–6. [Google Scholar]
  • 39.Oh SH, et al. Oscillatory mass transport in vapor-liquid-solid growth of sapphire nanowires. Science. 2010;330:489–493. doi: 10.1126/science.1190596. [DOI] [PubMed] [Google Scholar]
  • 40.Iserles A. A First Course in the Numerical Analysis of Differential Equations. Cambridge Univ Press; New York: 2009. [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary File
Download video file (55.2MB, mp4)
Supplementary File
Download video file (20.3MB, mp4)
Supplementary File
Download video file (7MB, mp4)
Supplementary File
Download video file (28.8MB, mp4)

Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES