You are on page 1of 17

REVIEWS

Surface premelting of water ice


Ben Slater   1* and Angelos Michaelides   2*
Abstract | Frozen water has a quasi-​liquid layer at its surface that exists even well below the bulk
melting temperature; the formation of this layer is termed premelting. The nature of the
premelted surface layer, its structure, thickness and how the layer changes with temperature
have been debated for over 160 years, since Faraday first postulated the idea of a quasi-​liquid
layer on ice. Here, we briefly review current opinions and evidence on premelting at ice surfaces,
gathering data from experiments and computer simulations. In particular, spectroscopy ,
microscopy and simulation have recently made important contributions to our understanding of
this field. The identification of premelting inhomogeneities, in which portions of the surface are
quasi-​liquid-like and other parts of the surface are decorated with liquid droplets, is an intriguing
recent development. Untangling the interplay of surface structure, supersaturation and surface
defects is currently a major challenge. Similarly , understanding the coupling of surface structure
with reactivity at the surface and crystal growth is a pressing problem in understanding the
behaviour and formation of ice on Earth.

Surface premelting of ice influences a multitude of phe- water molecule must accept and donate two hydro-
nomena on Earth, such as the genesis and lifetime of gen bonds, but this is not possible at the surface unless
clouds, the retreat of pack ice in the Arctic and Antarctic, reconstruction of the surface occurs. Additionally, if
the electrification of thunderclouds1 and, on a more the orientation of all the undercoordinated molecules
mundane level, potholes in roads. Premelting of ice is identical, this would lead to an intrinsically unstable
was first publicly reported by Faraday in the 1850s2 to polar surface. Hence, there is an expectation that the sur-
explain his now famous regelation experiment, which face molecules will also show orientational non-​random
he started a few years earlier. However, it was not until disorder under the naturally achievable temperatures
a century later that the first experiments appeared that and pressures on Earth.
could probe and quantify surface structural changes with An excellent summary of the physical and thermody-
temperature, and relatively recently there have been sev- namic foundations of premelting is presented by Dash
eral profound discoveries that enrich our understanding and co-​workers1 and Elbaum and colleagues7. As a general
of ice surface premelting. definition, the onset of premelting depends on the bulk
Premelting occurs on the surface and proceeds into melting temperature (Tm), with an increase in the thick-
the bulk ice crystal (Fig. 1), because the surface layers are ness of the premelting layer as T − Tm approaches 0 K.
more weakly bound than those in the bulk. The struc- It is also important to consider whether the surface is
ture of ice surfaces depends on the morphology of ice completely wetted or whether it would be decorated with
particulates. The dendritic hexagonal ice (Ih) crystal is droplets. At equilibrium, the Young expression relates
one of a panoply of distinct morphologies that are sum- the substrate (s), liquid (l) and vapour (v) surface free
marized in Fig. 2. The manifold morphologies arise from energies (γ) as γsv = γsl + γlv cosθ, where θ is the contact
1
Department of Chemistry, the interplay of temperature and supersaturation, which angle between the substrate and liquid phase; an angle
University College London,
London, UK.
influences the crystal growth mechanism and hence the of 0° indicates complete wetting. For a comprehensive
crystal habit3–5. Seeding ice formation with, for example, account of the thermodynamics of wetting, see refs8,9.
2
Thomas Young Centre,
Department of Physics and AgI and kaolinite can afford complex structural features Because water is reasonably volatile, the equilibrium of
Astronomy, and London such as gaps in the bulk crystal, which are known as hop- the liquid and vapour is interdependent at any given
Centre for Nanotechnology, per crystal morphologies. Most of the crystals depicted temperature and pressure, and thus the vapour pressure
University College London, in Fig. 2 expose the basal (0001) and prismatic (1010) is a key ingredient in establishing whether the ice sur-
London, UK.
planes, although, intriguingly, recent studies suggest that face is decorated by droplets or covered by a liquid film.
*e-​mail: b.slater@ucl.ac.uk;
the secondary (1120) prism face may be the most stable More generally, Stranski10 made an important contribu-
angelos.michaelides@
ucl.ac.uk face under laboratory conditions6. A subtle feature of all tion, reasoning that the surface energy of a crystal (of
https://doi.org/10.1038/ crystalline ice surfaces is that exposed interfacial H2O any composition) can be lowered by wetting the surface
s41570-019-0080-8 molecules are undercoordinated. By definition, each with its melt11. Frenken12 suggested that the presence of

Nature Reviews | Chemistry


Reviews

b c
0.37 nm

a
c

a c
b a
0.39 nm

d b

Fig. 1 | Crystal and surface structures of ice. A unit cell of hexagonal ice (Ih) (part a) and side views of a surface structure
of the basal face (part b), primary prism face (part c) and secondary prismatic face (part d) of Ih. The top of the surface is
indicated by an eye motif, and crystallographic axes labels are given to aid orientation. The bilayers for the basal face and
primary prism face are highlighted by the green arrows, and their heights are 0.37 nm and 0.39 nm for the basal and prism
faces, respectively. The unit cell depicts one possible arrangement of the water molecules possible in the P63/mmc space
group. Pauling proposed that there are (3/2)N possible arrangements of N water molecules in ice189 that obey the Bernal–
Fowler ice rules176 and that these would be isoenergetic. In fact, at 72 K , a hydrogen ordering transition can in principle
result in a distinct ferroelectric Cmc21 phase180, although a fully ordered phase has yet to be produced in the laboratory.
Density functional calculations have shown that the ordering transition is driven by purely electrostatic forces69,71.

a premelted layer can contrast nucleation near the melt- thickness of the QLL to be approximately 1–100 nm
ing temperature. A liquid requires a ‘nucleus’ of solid (refs16,18–20). Figure 3 vividly illustrates a central problem
to trigger freezing, which requires supercooling below in premelting: the disparity between the onset tempera-
the melting temperature, while it was argued that the ture and thickness of premelting layers between differ-
presence of the premelting layer on the solid means that ent techniques. This problem leads to questions such as
there is no nucleation barrier to forming the liquid from which technique captures the correct phenomenology
the solid. and what are the consequences of premelting from a
Traditionally, the big unsolved questions in ice pre- physicochemical point of view? We now step through
melting under ambient pressure are related to the tem- a timeline of key observations stemming from different
perature at which premelting occurs and the spatial techniques and piece the evidence together to highlight
extent of the premelted ice layers — the quasi-​liquid the state-​of-the-​art knowledge and the holes or fuzzy
layer (QLL). A range of techniques has been used to regions in our understanding of this topic. To aid navi-
investigate the onset temperature of premelting and gation, a timeline with some of the key developments in
thickness of the premelted layer on ice, and the out- the field is shown in Fig. 4.
come of measurements by a small subset of techniques In this brief Review, we summarize some of the key
are highlighted in Fig.  3. The temperature at which developments in the field and highlight outstanding
premelting begins has been estimated to be anywhere problems that are ripe for investigation through experi-
between ~200 K and 260 K (refs13–17), and the maximal mental and computer modelling approaches. We cannot

www.nature.com/natrevchem
Reviews

a b
Columns and plates Plates Columns Plates
0.3

Needles
Dendrites
Supersaturation (gwater m–3)

0.2
Sectored Dendrites
plates
Hollow
columns

Columns
0.1 Plates
Thin plates

Plates Solid plates Solid


prisms

0
238 243 248 253 258 263 268 273
Temperature (K)

Fig. 2 | the varied morphologies of hexagonal ice (ih). a | A survey of the relationship between the crystal habit and
supersaturation. See ref.4 for a comprehensive recent survey. The six-​fold symmetry of snow crystals has been noted
throughout history , a particularly early example being the poetry of Han Ying in circa 135 bc, who commented that “snow
flowers have six points”190, potentially depicted in 1555 (ref.191) and more certainly sketched in around 1600 by Dominic
Cassini192. b | Despite the dramatic variation in aspect ratio and crystal shape, the morphology , for example, is dominated
by the basal face (shown in blue), primary prismatic face (shown in green) and secondary prismatic face (shown in red); see
Fig. 1 for the atomic scale structure. gwater, grams of water. Part a adapted with permission from ref.193, AMS. Part b adapted
with permission from ref.194, AIP.

summarize 160 years of work towards understanding above 238 K, which was interpreted as the channels
the QLL of ice, and we also note that there is substan- being blocked along the c axis as a result of strong ther-
tial literature on premelting in other materials, such as mal vibrations and partial disordering. The disordered
metals12. Hence, we now discuss some important staging layer was found to increase in thickness at 213 K and to
points in the literature in roughly chronological order, reach ~80 nm at 271 K (which was extrapolated to be
with the perspective of trying to bracket what tempera- 95 nm at 272 K, corresponding to ~260 bilayers). By stark
ture premelting initiates at and the thickness of the pre- contrast, surface conductivity measurements27 detected
melting layer at a given temperature. We emphasize that anomalous increases in conductance at only 262–267 K,
we focus on ice premelting in contact with the vapour which were presumed to signal the formation of a liquid-​
phase but recognize that premelting of ice occurs on a like layer. Presumably, this anomaly signals the com-
host of substances and direct readers to the excellent plete wetting of the surface, which leads to substantial
review of Dash et al.1 for an introduction to that aspect mean-​free path lengths that could give rise to increased
of premelting. We also emphasize that our aim here is for conductivity on the microscopic scale.
a critical review of recent work in particular rather than Kouchi et al. reported28 X-​ray diffraction of single-
an exhaustively detailed and complete historical account crystal and polycrystalline ice surfaces in the 263–272 K
of the literature. Several authors have reviewed work on temperature regime. The ice crystals were prepared
the QLL, and we encourage readers to delve into this carefully using double-​d istilled water under slow-​
literature for further information1,11,21–26. crystal-growth conditions, and a microtome was used
to prepare the (0001) and (1010) surfaces. The mea­
Early characterization surements showed an absence of long-​range struc-
Specific studies on premelting of ice started to appear ture around 272 K, with no appreciable difference in
in the late 1950s, and here we focus on a selected set of the ratio between the characteristic (0001) and (1010)
experimental and theoretical studies that sought to pin secondary peak intensities over the 9.5 K window
down answers to the basic questions of at what temper- explored. These results indicated that premelting, as
ature does the QLL originate and how thick is the QLL at measured by the X-​ray penetration depth, was equiv-
a given temperature. Golecki and Jaccard13 reported on alent for both the basal and prismatic planes for both
proton backscattering and channelling experi­ments single-​crystal and polycrystalline samples. However,
on the (0001) surface between 143.15 K and 271.35 K. at ~271 K, there was a marked decrease in the inten-
They observed a marked increase in backscattering sity of the first 2θ peak, which can be interpreted as

Nature Reviews | Chemistry


Reviews

100 minimal defects that are molecularly smooth and rela­


X-ray diffraction tively free from steps. Ellipsometry detects changes
X-ray absorption in the polarity of light impinging surfaces and was used in
Ellipsometry
Molecular dynamics this study to measure the refractive index and QLL thick-
Monte Carlo ness of a single crystal of ice. The refractive index of the

QLL thickness (nm)


10
surface transition QLL was found to be 1.33 for both
the (0001) and (1010) surfaces, which can be compared
with 1.3327 for bulk water and 1.3079 for ice, indicat-
1
ing more liquid-​like than ice-​like properties. The onset
of premelting was reported to be 271 K and 269–272 K
for the (0001) and (1010) surfaces, respectively, and the
QLL thicknesses at 272 K were approximately 20 nm and
0
30 nm for the (0001) and (1010) surfaces, respectively
(estimated from a redrawn figure31). A curiosity of this
233 243 253 263 273
study is that the QLL on the prismatic face at 273.05 K is
Temperature (K)
several hundreds of nanometres thinner than the QLL
Fig. 3 | A selective overview of estimates of observed at the lower temperature. Earlier ellipsom-
approximate QLL thickness obtained from physical etry experiments by Beaglehole and Nason19 reported
measurement. X-​ray diffraction15, X-​ray absorption20, an opposite trend in the estimated QLL thicknesses:
ellipsometry30 and two computer simulation approaches: 100 nm for the prismatic face at 253 K and 0 nm for
molecular dynamics18 (using the TIP4P-​ice model) and the basal face at the same temperature. Furukawa et al.
grand canonical Monte Carlo with molecular dynamics101 attributed discrepancies between their observations and
(using the coarse-​grained model of water). The dotted other studies to potential sample purity problems
yellow line indicates the approximate boundaries for the and other technical aspects. Elbaum et al.32 estimated the
quasi-​liquid layer (QLL), estimated using two different
QLL on the basal plane at ~273.05 K to be 18 nm thick
simulation approaches; see ref.101 for a detailed account.
The different experimental probes produce radically (ice crystallites grown from and in water vapour) and
different estimates of what temperature the QLL appears accompanied by macroscopic and mobile droplets (that
at and the largest depth of QLL measured. Computer did not fully wet the surface) with a contact angle of 0.6°.
simulations with well-​validated potential models produce When air was admitted to the sample, a much thicker
a much narrower estimate of the maximum QLL thickness, macroscopic QLL of ~200 nm was formed, indicating
and the rates of premelting from the two independent complete wetting or surface melting. This study prefig-
simulations are in close accord. Interestingly , only ures work by Sazaki and co-​workers33,34, who explored
photoelectron spectroscopy measurements appear to be the relationship between saturation and the QLL more
qualitatively similar to the molecular simulations. For a than 20 years later.
more complete illustration of the disparity in the literature
An atomic force microscopy (AFM) study, carried
between techniques, see figure 7 in ref.24.
out with a controlled humidity of 85 ± 5%, put the QLL
at 11 nm at 263 K (ref.35) and 32 nm at 272 K (ref.35),
pronounced melting at the surface. A more recent study whereas Volta effect measurements placed the premelt-
by Dosch et al.15 showed substantial differences in the ing onset at 243 K (ref.36), with no direct estimate of the
thickness of the QLL on the (0001) and (1010) surfaces QLL depth. Mizuno and Hanafusa reported NMR14
but only marginal differences in the premelting onset spectra of powdered ice that was formed from droplets
temperature. Premelting onset was estimated to be of <150 μm in diameter to attain a high surface-​to-bulk
260 K for the (0001) surface and 261 K for the (1010) signal ratio. In addition, to provide information about
surface. However, the QLL thickness was estimated to sample purity and kinetic influences on crystalliza-
be ~50 nm at 272.85 K for the (0001) surface, ~15 nm tion, the surface-​to-bulk signal ratio in NMR spectra
at 272 K for the (1010) facet and approximately 44 nm measurements can be, in principle, related to the QLL
for {1120} at 273 K. Importantly, Dosch et al.15 estab- thickness, although this was not estimated in this study.
lished a logarithmic dependence of the QLL with tem- The onset of the formation of the QLL was not distinct,
perature and cleverly avoided scattering from nascent but the barriers for rotational and translation motions
and/or non-​equilibrium vicinal surfaces arising from at ~220 K were found to be 0.29 eV and 0.25 eV, respec-
evaporation by using a small-​angle or glancing-​angle tively, which are larger than those in water (~0.19 eV)
approach. Dosch et al.29 also performed an extremely and substantially smaller than those found for bulk ice
delicate study revealing hydrogen Bragg reflections (~0.65 eV), reflecting distinct phase properties for
at the surface. A Bjerrum-​defect-rich (Box 1) 2–3 nm the surface molecules at that temperature. No discon-
region (approximately 6–8 bilayers) with disrupted tinuity in the activation barriers was detected up to
hydrogen bonding within a 25 nm sample depth was 268–271.5 K, indicating that no clear bulk liquid signal
observed at 260 K. The presence of point-​defect-rich was detectable even in the outermost bilayer. Similarly,
surface layers and their contribution to premelting the self-​diffusion coefficient of the water molecules
behaviour are areas deserving of more study. on the surface was found to be ~4% of that of liquid
Furukawa et al.30 performed ellipsometry exper- water, even up to 271.5 K.
iments on a negative crystal of ice exposing (0001) In one of the earliest examples of the conceptual
and (1010) surfaces — precisely oriented surfaces with modelling of the ice surface structure, Weyl speculated

www.nature.com/natrevchem
Reviews

Conde et al.
report the first DFT studies show
quantitative pristine ice surfaces to
Bernal and Fowler Materer et al. show through estimate of QLL be amorphous-like102
publish the ice rules174 LEED and theory that ice is depth using a
Faraday’s regelation fully bilayer terminated51 TIP4P potential
model that gives GCMD with
experiment and the Pauling Sazaki et al. controlled vapour
airing of the postulates the the correct bulk
Elbaum and Schick predict observe pressure shows
liquid-like layer residual entropy melting
coexistence of droplets and steps at the that QLL is
concept concept187 temperature18
wetting layers45,46 ice surface33 inhomogeneous101

1850 1930 1950 1980 1990 2000 2010 2018

First determination Crystal structure NMR shows Elbaum et al. Glancing-angle Onset of premelting
of lattice parameters of hexagonal ice putative evidence observe droplets diffraction linked to the
due to Bragg determined by of premelting at and wetting determines onset disruption of two
X-ray diffraction 173 K (REF.14) layers7,32 of premelting bilayers by FTIR
temperature15 and MD16
Fletcher predicts
proton ordering at Sazaki et al. observe inhomogeneous
the ice surface40 premelting and advocate the α-QLL
Theory Experiment and β-QLL via optical spectroscopy112

Fig. 4 | timeline of some of the key developments in unravelling the nature of the QLL of ice, as selected by the
authors. For a more complete summary of the evolution of the field, readers are encouraged to refer to other reviews of
the literature1,22–24,195. DFT, density functional theory ; FTIR , Fourier-​transform infrared spectroscopy ; GCMD, grand
canonical molecular dynamics; LEED, low-​energy electron diffraction; MD, molecular dynamics; QLL , quasi-​liquid layer.

that water molecules on the surface would be oriented prismatic planes and taking account of the hexagonal
with the hydrogen atoms pointing towards the bulk symmetry, he studied the optimal arrangement of water
owing to electrostatic and hydrogen bonding considera- molecules on a surface to minimize the electrostatic
tions37. He reasoned that the presence of dangling hydro- dipolar repulsion between molecules. Fletcher argued
gen bonds should be minimized because they would that an ordered, striped surface configuration, in which
lead to an unstable arrangement. At the time Weyl’s each molecule in the outer layer is surrounded by six
arguments were made, water molecule orientation nearest neighbours with four antiparallel and two parallel
and arrangement in the bulk crystal were not certain; dipole moments, would minimize electrostatic interac-
nevertheless, the arguments presented helped frame tions on the basal plane. On the prismatic plane, pairs of
the question of the surface structure of ice. Probably the antiparallel molecules were predicted to stripe the sur-
most important contribution towards discrimina­ting face. Note that the proposed basal and prism configura-
the surface structure of ice, especially the QLL, and tions were both microscopically apolar in the direction
its influence on properties was made by Fletcher38. He perpendicular to the surface, an important consideration
proposed the first quantitative model prediction of the in minimizing the surface energy. Fletcher predicted that
onset of premelting to be just above 243 K on the basis these ordered surface configurations would be adopted at
of free energy, hydrogen bonding, Bjerrum defects and very low temperatures: 30 K for the basal plane and 70 K
electrostatic considerations and stated that the thickness for the prism plane. These models were not quantitatively
of the QLL would increase nonlinearly towards the bulk tested until more than a decade later, when they were
melting point Tm. A revised model39 was proposed that demonstrated to be favourable low-​energy structures41–44.
took into account the quadrupole moment of water. This Elbaum and Schick45,46 applied the Dzyaloshinskii,
and more sophisticated models39 predicted premelting Lifshitz and Pitaevskii (DLP) dispersion model47 to
temperatures between 267 K and 270 K, depending on predict that the surface melting of ice would be incom-
the Bjerrum L-​defect (see Box 1) formation energy, plete at the triple point. Most importantly, their model
and a QLL thickness of ~1 nm at 267 K and between predicted the coexistence of complete wetting layers
3 nm and 4 nm at 272 K. Fletcher’s model substantially and water droplets, in line with their own experimental
underestimated the QLL thickness according to the observations32. Using the DLP model and the experi-
measurement reported above, but it predicted an onset mental data reported by Elbaum and Schick45, the liquid
temperature in fair agreement with experimental obser- thickness at the free energy minimum (see the Young
vations. This model predates computer modelling of ice, equation and ref.45) is between 2 nm and 4 nm, which
but these predictions of QLL thickness are actually in is in marked contrast to the experimental thicknesses
very good agreement with state-​of-the-​art predictions reported above. Furthermore, the model predicted that
performed with classical potential-​based models some the rates of premelting at equilibrium and away from
40 years later. equilibrium are not equivalent48.
Fletcher made another important contribution to the The selection of data presented here (more extensive
field40: considering the crystal structure of the basal and summaries and a critique can be found elsewhere21–24)

Nature Reviews | Chemistry


Reviews

Box 1 | Basic crystallography of bulk ice emergence of atomistic modelling. From around the
1980s onwards, UHV approaches reached a state of
although crystalline ice has many phases, we typically experience only one form of ice: maturity and experienced widespread use in fields such
the hexagonal phase of ice, known as ih. in the crystal bulk, the oxygen atoms of the as surface catalysis49. To contribute to the QLL ques-
molecules are arranged in the same way that silicon stacks are in the tridymite mineral tions of structure and premelting temperature, it was
and carbon in lonsdaleite. Lonsdale summarized the key developments172,173 in the story
necessary to perform extremely well-​controlled exper-
of the structure solution, and the position of the hydrogen atoms (actually D2O) was
resolved unambiguously using neutron diffraction in 1957 by Peterson and Levy174, after iments, with meticulous sample preparation to mini-
the first neutron study by Wollan et al.175. the bulk structure is shown in the left panel of mize the possibility of surface contamination, which
Fig. 1, along with the key surface structures for hexagonal ice (right panel). could affect the measurements. Additionally, computer
the molecules are arranged in bilayers that stack in an aBaBaB repeat sequence simulations were required to model the structure and
in the [0001] direction shown in Fig. 1. each molecule in an infinite crystal of ice ih predict the stability of liquid water and crystalline ice
obeys the Bernal–Fowler ice rules176, which state that each oxygen atom is bonded with sufficient accuracy. By the 1990s, surface science
to two hydrogen atoms and that each oxygen hydrogen-​bonds to two other oxygen approaches, models (classical and quantum chemical,
atoms so that only one hydrogen atom is located between oxygen nearest neighbours. especially density functional theory (DFT)) and com-
each water molecule has a tetrahedral geometry, but there are six possible molecular
puter architecture had come of age, opening up the pos-
orientations that obey the ice rules, giving rise to orientational disorder. One
sibility of answering basic questions, such as what is the
consequence of this orientational (and hydrogen) disorder is that ih is paraelectric.
upon cooling to ~72 K (ref.177), the water molecules in principle align to adopt a distinct surface structure of pristine crystalline Ih and what is
hydrogen ordered analogue of ice ih, namely, the ice Xi178–180 phase, which is the structure of the QLL?
ferroelectric. A landmark demonstration of the synergistic ben-
the ice rules can be violated by the formation of orientational defects, termed efits of allying surface science techniques and model-
Bjerrum defects181; L defects (from the German leere, meaning empty), which have a ling approaches was reported by Materer et al.50 (see
missing hydrogen atom between two neighbouring oxygen atoms; and D defects (from also ref.51), who used low-​energy electron diffraction
the German doppeltbesetzte, meaning doubly occupied), in which two hydrogen atoms (LEED), classical molecular dynamics (MD) and
lie along the vector connecting two neighbouring oxygen atoms. additionally, charged DFT calculations to study the (0001) ice surface. In
ionic defects can form from autoionization, for example, yielding H3O+ (hydronium) and this work, a Pt(111) substrate was used to grow an ice
OH− (hydroxide). a self-​interstitial defect is another common point defect, in which a
film at 140 K and the LEED pattern acquired at 90 K.
molecule displaces off its ideal lattice site (see refs22,182).
Co-​author Kroes built upon their TIP4P study at
190–250 K (ref.52) (see below) to simulate the (0001)
highlights considerable variation and therefore uncer- surface at 90 K. According to LEED, hexagonal ice (Ih)
tainty in the thickness of the QLL, its onset and whether rather than cubic ice (Ic)(Box 2) was resolved with rea-
particular faces of the ice crystallites are more prone to sonable confidence, but the measurements did not help
premelt at lower temperatures than others. Suffice to say, distinguish whether the ice surface was terminated by a
variation in these data arises from the effects of thermal full bilayer or a half-​bilayer. Intuitively, the full bilayer
vibrations on the surface structure, the potential influ- model would be expected to be more stable because
ence of impurities (extrinsic defects), intrinsic point and the half-​bilayer or single-​layer models expose twice as
line defects, and sample preparation artefacts and vapour many dangling hydrogen bonds as does the full bilayer
pressure. More fundamentally, the connection between model. Indeed, the DFT results discussed within this
the atomic scale structure, the ideal surface structure paper appear to confirm this speculation, and the clas-
and surface properties under any relevant conditions sical MD suggested that only the bilayer model is com-
was limited by the absence of local or average structure patible with the structure and amplitudes of vibration
information about the surface. Furthermore, the scatter measured. Notably, according to LEED measurements,
in the reported QLL onset temperatures and QLL thick- the amplitude of vibration of the upper layer of the first
ness measurements is in part due to the probe depth and bilayer (which has one dangling hydrogen bond) at a
the intrinsic resolving power of the experimental tech- chilly 90 K is large enough that it cannot be directly
niques. A key issue in the experiments mentioned in this discerned from the lower layer of the same bilayer. The
section was the control and equilibration of the vapour key experimental evolution and findings in this field
pressure, as vividly shown in the work of Elbaum et al.32. are summarized by Li and Somorjai25. However, LEED
In that case, it was reported that mobile water droplets could shed no light on the nature of the arrangement
under water vapour did not fully wet the surface but that of the protons at the surface and the structure of the
an ~200 nm QLL was formed at the same temperature premelting layer.
and ambient air pressure. Given the extreme sensitivity Bluhm et al.20 performed an important set of X-​ray
of the nature of the ice surface to vapour pressure, a route photoelectron spectroscopy (XPS) and near-​e dge
to disambiguation would be to eliminate vapour and X-​ray absorption fine structure (NEXAFS) experi­
perform the measurements under vacuum conditions. ments. Bluhm and Salmeron developed a novel
This required progression in instrumentation through near-​ambient-pressure XPS experiment that enables
the development of surface science approaches as well as determination of atomic-​level information at well-​
modelling approaches to resolve the structure of the QLL. defined solid–liquid interfaces. Their study revealed that
the onset of a discernible QLL occurred above 250 K and
High vacuum and modelling studies that, at just below the melting temperature (271 K), the
A major change in the refinement of our knowledge of QLL was ~2 nm thick. Their result is notable not least
premelting came from the advent of ultra-​high vacuum because it marks the lowest experimental estimation of
(UHV) approaches in surface science combined with the QLL depth just below Tm.

www.nature.com/natrevchem
Reviews

From around the 1980s, more or less contemporane- determined to be crystalline, evidenced by no self-​
ous with developments in surface science, classical mod- diffusion, whereas appreciable self-​diffusion, rotational
elling approaches based on force field methods began disorder and translational disorder comparable to those
to be instructive in water ice studies53. A key devel- of bulk water were observed above 230 K. Increased
opment in this field was the advent of the rigid body polarization of the surface was noted as the target tem-
TIP4P54 four-​site water potential. The three-​site single perature was elevated, and the protons of the water
point charge (SPC)55 and TIP3P55 models had already molecules at the vacuum–ice interface layer were noted
appeared, but TIP4P had proved to yield accurate repro- to point into the surface rather than out of the surface.
duction of the water radial and partial radial distribution Kroes carefully pointed out that the nonpolarizable
functions and the structure of ice. Another important nature of the TIP4P model and the use of fixed layers
development was reported by Hayward and Reimers56, in the simulation precluded him from commenting on
who examined the random orientation of water mole- whether surface melting was complete. Importantly,
cules within ice structures subject to polarity constraints it has subsequently been shown that the melting tem-
to produce ice configurations that contained intrinsically perature of ice with the TIP4P model is approximately
apolar bilayers. It is important to include apolar bilayers 230 K (ref.60), so it is likely that full melting occurred
if one wants to model apolar surfaces without recon- at 230 K and above but was tempered by the fixed ice
struction, despite the small energetic penalty for violat- substrate. Nevertheless, the disordering observed in the
ing the constraint. The advent of well-​behaved density regime between 210 K and 230 K was evidence of pre-
functionals and the availability of supercomputers are melting well below Tm for this model. Similar findings
other crucial developments in quantum chemistry and were reported by Bolton and Pettersson using TIP4P61.
computer hardware that enabled the study of water Nada and Furukawa59 also used the TIP4P model and
and ice at the quantum mechanical level on a tractable compared the structural responses of (0001) and (1010)
timescale. For a contemporary survey of the performance ice–water interfaces. The study considered 100 ps MD
of DFT for water and ice, the reader is referred to the simulations again at 230 K, now known to be the bulk
review by Gillan et al.57, and for a greater emphasis on melting temperature of TIP4P ice. The important result
classical models, to the review by Cisneros et al.58. was that the (0001) surface was shown to be more
Building on the advent of the TIP4P model, two disordered than (1010), with a larger disordered layer
important modelling contributions to the evolution thickness on the basal plane hinting at strongly ani-
of the ice premelting came about: the first was made sotropic surface properties and structural responses.
by Kroes52, and the second was made by Nada and Notwithstanding the uncertainty over the effects of heat-
Furukawa59. Kroes considered the MD of a periodic-​ ing model slabs beyond the bulk melting temperature
oriented (0001) slab of ice in the 190–250 K temper­ of the TIP4P model, this study was interpreted as pro-
ature range for up to 260 ps. Below 210 K, the surface was viding evidence of no major distinction in the melting
responses of these two surfaces.
The arrangement of protons at the ice surface was
Box 2 | Cubic ice and stacking disorder
proposed by Buch et al.41, building on two sets of exper-
there are two additional metastable phases of ice that are important for understanding imental data. The first set of data was obtained from
the physicochemistry of ice on earth: cubic ice, ic (ref.183) (aBCaBC stacking of the sum frequency experiments on the surface of ice, which
bilayers), and stacking disordered ice, isd (refs88,184), which contains a randomly mixed showed orientational disordering of surface water
and varied proportion of ic and ih sequences. at the time of writing, there is no molecules at 200 K (ref.17) that was first interpreted as
unambiguous evidence that ideal ic crystals have been prepared in the laboratory a proxy for premelting and later confirmed in another
(though a scanning tunnelling microscopy (stM) image of individual layers of ic has
independent study62. The second set of data was obtai­
been reported115), although nature may well have achieved this feat185. in 1629,
Christophe scheiner noted a halo around the sun at approximately 28° owing to the
ned from helium scattering studies63 that previously
scattering of solar rays by ice crystals in high-​altitude clouds. a purely cubic ice crystal could not help discern whether the basal surface of ice
could have a perfect octahedral morphology (completely distinct from documented showed ordered arrangements of protons, such as
forms, see Fig. 2) that would give rise to refraction at approximately 28°, while Fletcher’s phase40, or a disordered arrangement in which
hexagonal columnar ih ice crystals refract rays at 22°. the 28° signature of pure, ideal the Fletcher phase is not dominant, as was suggested
cubic ice is incredibly rare, being reported less than ten times since 1629. by more recent measurements64. Buch used dynamical
Despite the dearth of evidence for the existence of ideal cubic ice, use of ic in the and Monte Carlo simulations based on TIP4P-​ice65 to
literature is still ubiquitous, although isd is likely a more accurate classification of this show that the Fletcher striped phase of protons was
phase type. isd and ic are metastable with respect to ih. isd or ic transforms to ih at ~150 K, most stable, and in fact, this arrangement was found
but detailed analysis carried out by Malkin et al.88,186 suggests that cubic sequences are
to be stable until surface reconstruction occurs above
evident up to 257 K. above ~176 K, ic or isd transforms irreversibly to ih (ref.187), yielding
–46 ± 3 J mol−1 at 200 K, but the lifetimes of ic and isd are non-​negligible, certainly in the
180 K. Comparison of the models with previous sum
context of the reactivity of ice particulates in clouds. Despite the tiny energy difference frequency generation (SFG) studies23,66 provided addi-
between isd, ic and ih, Kuhs et al.145 suggest lifetimes of isd crystals of several days (by tional evidence to support the hypothesis that protons
extrapolation) at 175 K and several hours at 185–190 K, and advocate that the isd phase adopt ordered arrangements at the surface. Very shortly
persists up to at least 240 K. after this work was published, Pan et al.42–44 demon-
Computer simulations by Lupi et al.188 show that nucleation of isd at 230 K is three strated through a DFT assessment that ordered proton
orders of magnitude faster than nucleation of ih, underlining its importance in the arrangements, including those proposed by Fletcher,
context of premelting and recrystallization phenomena. Hence, the role of ic and isd in were indeed the lowest energy arrangements of protons
premelting needs to be considered, along with the dynamics of crystallization and at the surface. In addition, the lowest energy configu-
surface recrystallization under different vapour saturation conditions.
rations were found to be stable beyond the premelting

Nature Reviews | Chemistry


Reviews

and bulk melting temperatures, suggesting that there bilayer, partially ordered, a structure that is qualitatively
is a clear driving force for the surface of ice to display consistent with the findings of Conde et al. Note that
ordered hydrogen configurations rather than random simulated surface energies are incredibly delicate quan-
distributions. The latter prediction may be important tities because they measure the difference between bulk
in understanding uptake and chemical reactivity on ice and surface stability and have been found in remarkably
particles in high-​altitude clouds, for example, in which good accord, ~30–35 mJ m−2, using a variety of poten-
ordering will lead to patches of the surface that are more tial models and different techniques, including the cap-
favourable for binding electrophiles and others that are illary wave method and mould integration83,84. We note
more attractive for nucleophilic centres67. It is noted that in passing that in the context of the proposal of Frenken
potential models such as the TIP4P family generally fail and Stranski, the function of the QLL is to passivate
to reproduce the correct ground-​state bulk crystal struc- the solid as the bulk melt temperature is approached.
tures for hydrogen ordered ices (such as ice XI), while McBride et al.85 determined that the TIP4P model of
DFT has an excellent track record of predicting ground-​ ice can be superheated to around 80 K (ref.86) beyond its
state hydrogen ordered structures68–76. Other surface bulk melting temperature when all the free surfaces are
ordered configurations were also found with internal removed, which is just the crystal bulk simulated under
energies that were considered to be isoenergetic with the 3D periodic boundary conditions.
Fletcher striped phase within the uncertainties associ- We note that for many properties of the QLL, the
ated with DFT. Pan et al.42 also discerned that the surface coarse-​g rained model (mW)87 of water, which has
energies of the basal and prism faces in their optimal been used extensively to study homogeneous88–93 and
arrangement of protons were essentially identical within heterogenous94–97 ice nucleation, yields results that are
the accuracy of DFT. generally consistent with those from atomistic models
Although the terminating structure of ice at low tem- of water87,91,98–101. Because of its greatly reduced compu-
perature seemed to have been resolved, the premelting tational cost compared with that of standard atomistic
temperature and QLL thickness remained open ques- models, it is possible to examine larger system sizes (see
tions for theory. Abascal and Vega performed crucial below) and perform more extensive analyses of the sen-
work laying the foundation for more accurate modelling sitivity of the QLL thickness to temperature. Notably,
of the QLL by reparameterizing TIP4P to reproduce the Limmer and Chandler used mW to examine the thick-
melting temperature of ice60,65,77 and key features of ness of the QLL at T up to Tm − 0.25 K in what is to date
the phase diagram for water for the first time. The TIP4P- the most detailed and probably best converged study
2005 (Tm = 249 ± 3 K) and TIP4P-​ice (Tm = 271 ± 3 K) close to the melting temperature, finding logarithmic
models were used by Conde et al.18 to determine the evolution of the QLL from approximately Tm −5 K. The
thickness of the QLL and the onset temperature for latter study was performed on cubic ice (not stacking
premelting. Both models predicted a similar thickness disordered ice; Box 2); however, it remains to be iden-
of the QLL, but a qualitative difference was found in tified whether the same qualitative behaviour is seen in
the respective QLL depths of the basal plane compared hexagonal ice. Likewise, it would be interesting to see if
with the primary and secondary prism faces. TIP4P-​ice the same logarithmic dependence is recovered with fully
gave the largest QLL on the secondary prism face rather atomistic water models.
than the primary prism or basal face, whereas TIP4P- All the work described thus far focused on pris-
2005 gave the smallest QLL on the secondary prism face. tine ice surfaces, but in reality, intrinsic defects will be
Focusing on the TIP4P-​ice results, which reproduced the present, of which vacancies and interstitial defects are
Tm more accurately, the onset of premelting was found expected to be the most prevalent22. In 2011, Watkins
to occur at ~170 K for the basal face, ~190 K for the pri- et al.102 reported a DFT study of the basal plane of the
mary prism face and ~200 K for the secondary prism ice surface at 0 K that revealed a surprising finding:
face. These estimates are broadly in line with the helium the surface of perfect crystalline ice is amorphous-like.
scattering experiments of Suter et al.78, who detected It was found that vacancy formation energy (in this case,
an anomalous response from the surface at 180 K, and the energetic cost of removing a water molecule from the
compatible with the expected premelting temperature crystal and placing it in a bulk vacancy) varies by up to
according to Tammann’s postulate79 (~180 K). ~0.7 eV depending on whether a molecule is removed
The results reported by Conde and co-​workers are from the outer layers or the crystal bulk. The outermost
also comparable to the XPS and NEXAFS results of layer shows a remarkably large variation of ~0.35 eV
Bluhm et al.20, which yielded an estimate of the QLL depending on the molecular environment of the mol-
thickness of ~2 nm at 271 K, just below the melting tem- ecule that is removed, as shown in Fig. 5. The average
perature. However, Bluhm’s estimate of the premelting vacancy formation energy in the outermost layer is
onset was above 250 K, in marked disagreement with the very comparable to the activation energy for translation
study of Conde et al. Similar QLL depths were obtained reported by Mizuno and Hanafusa using NMR14 and
by Carignano et al.80 using a six-​site water model81 at viscosity measurements that show molecular transport
275 K, which at the time was thought to be above Tm, to be mediated by vacancies103. Ordinarily, crystalline
but a subsequent study determined Tm for the six-​site materials are expected to show a single value for the
TIP6P model to be 287 K (ref.82). Simulations performed vacancy energy in bulk, which is the case for bulk ice.
by Bishop et al.44 at 285 K using TIP6P show figuratively At the surface, the vacancy formation energy is expected
that only the top two bilayers have liquid character, the to be less than that in the bulk, and attenuation between
outermost layer being disordered, while the second the bulk and surface vacancy formation energy is also

www.nature.com/natrevchem
Reviews

expected. It was found that while the average vacancy investigation, it was found that the formation energies
formation energy actually decreases the closer to the were correlated with the dipole moment of the mol-
surface the removed molecule is, a wide variation was ecules that were being removed, which varied from
observed, especially in the lower part of the first bilayer 2.9 D to 4.4 D, as opposed to the bulk solid, in which
and the upper part of the second bilayer. Upon closer the dipole moment was uniform at 3.5 D (ref.102). By the
third bilayer, the dipole moment distribution was more
or less like the crystal bulk, so just the two outer bilayers
a b were qualitatively different from the underlying part of
– –
the crystal102. The outer two layers of the crystal have
molecules that sit essentially on their ideal lattice sites,
+ – but the heterogeneity of the dipole moments (which
arise because of the orientational disorder at the sur-
face) means that these layers have qualitatively distinct
+ + properties from those of the crystal bulk. Bizarrely, some
of the water molecules in the outermost two bilayers are
more strongly bound to the crystal than molecules in
+
the crystal bulk102. These computational results were

corroborated in a separate study by Moreira et al.104.
Recent SFG measurements105,106 showing that the surface
– + of ice has properties resembling supercooled water argue
that there are two key temperatures in the premelting
regime: ~245 K, which signals the melting of the outer-
+ – most bilayer, and ~270 K, which signals the melting of
the second bilayer. The interpretation is supported by
c insight from the work of Sánchez et al.16, who showed,
0.9 using MD with a polarizable water model, that the first
bilayer melts around 235 K and the second around 270 K.
0.8
Studies by Smit et al.105 and Sánchez et al.16 suggest that
0.7 around 270 K, only the outer two layers are melted and
form a film ~1 nm thick. This interpretation is further
1
Vacancy energy (eV)

0.6 2 supported by two IR studies by Sadchentko and Ewing,


3 in which they used a cold finger setup107,108. The study by
0.5 4 Watkins et al.102 highlights the importance of molecular
5 polarization and defects that need to be considered in
0.4
6
simulations of premelting and that there is an explicit
0.3 6
5 heterogeneity102 in ice surfaces that is present before
4 premelting.
0.2
3
0.1 2 Heterogeneity at the surface
1 A major experimental breakthrough has come from the
0.0 work of Gen Sazaki and co-​workers33,34,109–111. Over
0 1 2 3 4 5 6
Layer number the past few years, Sazaki et al. have pioneered the use
of laser confocal microscopy combined with differential
Fig. 5 | inhomogeneity in the outer two bilayers of ice. a | The variation in the dipole interference contrast microscopy to study the surface
moment of water molecules projected onto coloured spheres. Weakly bound water of ice. One of the many reasons that this work is par-
molecules with small dipole moments are represented in red, strongly bound water ticularly important is related to the use of a direct local
molecules with large molecular dipole moments are represented in blue and molecules
probe to elucidate structural features, providing unprec-
exhibiting intermediate binding and dipole moments are in grey. b | The orientationally
partially disordered arrangement of molecules at the surface gives rise to dipolar fields edented insight into features of ice crystals. Sazaki et al.33
(indicated in light green) that reduce (top panel) or increase (bottom panel) the dipole reported direct visualization of steps on the basal and
moment of the central water molecule. c | Vacancy formation energies computed using prism surfaces of an ice crystal. Their work showed
density functional theory by Watkins102. The work by Watkins et al. showed that the four the time-​evolved structure at the surface, revealing a
layers in the first two bilayers of ice of the inset basal slab are more weakly bound than 2D crystal growth mechanism and that the step heights
the layers in the centre of the slab. The centre of the six-​bilayer slab exhibits uniform on both the basal and prism faces corresponded to
vacancy formation energies equating to almost degenerate binding energies. the height of a bilayer, thus indicating that incomplete
Conversely , a fraction of molecules in the first bilayer (in particular) and in the second bilayers are metastable. Images were collected at ~263 K
bilayer are relatively weakly bound and presumably more susceptible to premelting. and showed step-​bunching and regular step separations
The defect-​free crystalline ice surface has an amorphous-​like quality , which may be
revealing that, despite various spatially and temporally
related to the evolution of the quasi-​liquid layer. Intriguingly , a sum frequency
generation and molecular dynamics study by Sánchez et al.16 showed clear melting averaged techniques indicating the existence of a liquid-​
of the first two bilayers but no clear melting of the third bilayer, which may be expected like character, the surface has crystalline, sharp features
to be a consequence of the distinct polarization character of the outer two layers of suggesting that short-​range and long-​range orders are
ice106. Inset depicts a six-​bilayer structure and the corresponding bilayer numbering. present at these temperatures. The same study also
Parts a, b and c adapted from ref.102, Springer Nature Limited. showed growth spirals (dislocations) that were seen to

Nature Reviews | Chemistry


Reviews

form under the conditions sampled. Growth spirals were that of α-​QLLs, indicating that the wettability of β-​QLLs
also imaged on the basal face in an earlier ellipsometry is higher than that of α-​QLLs and that the interaction
study30, and the microphotograph depicted therein may energy of β-​QLLs with ice is more favourable than that
also show droplets. Sazaki et al. observed crystal growth of α-​QLLs. α-​QLLs and β-​QLLs were seen to coexist,
on the prism face at ~271 K, again indicating that crys- suggesting distinct phases, and be immiscible, further
talline order is maintained at these temperatures, despite supporting the idea that these phases are distinct. In pre-
SFG measurements suggesting premelting of two bilay- vious experiments, these distinct heterogeneous phases
ers at this temperature. Next, Sazaki et al.112 imaged the were not observed, perhaps owing to the lower spatial
ice surface in the range of 258–272.9 K and performed a resolution or kinetic artefacts. These two forms of QLL
delicate investigation of the surface in the 271–272.9 K were initially observed on the basal plane of ice but then
range, collecting measurements every 0.1 K steps at a subsequently observed on the prism and other faces of
heating rate of 0.02 K per minute. The crucial finding an ice crystal111. In a related paper113, it was shown that
was the observation of water droplets (of approximately α-​QLLs occur at screw dislocations, while β-​QLLs are
micrometre scale), which were termed α-​QLLs (bulk associated with microdefects (pits on the surface), and
liquid droplets), at the surface below 273 K. These drop- in both cases, a strain field promotes the formation of
lets were found between 271.5 K and 272.6 K depending these distinct QLLs. The detailed surface structure
on the particular experiment (see Fig. 6). The observa- of screw dislocations has been assessed114 at 270 K, and
tions were consistent with previous optical reflection it was found that the dislocation core was immediately
(interference microscopy and ellipsometry) measure- enveloped by a double bilayer, such that the hexagonal
ments by Elbaum et al.32, who also observed droplets on ABABAB stacking sequence was perpetuated. Similar
the surface of ice close to the melting temperature double-​bilayer structures were observed on the prism
(sensitive to the vapour pressure), in line with previous and basal plane. Similar features and unambiguous sig-
predictions45. natures of cubic ice spirals were also identified by AFM
The droplets in the Sazaki experiment112 were seen to and scanning tunnelling microscopy (STM)115 on an ice
coalesce until they formed a film coating the crystal. The sample imaged at 145 K.
film was determined to be considerably thicker than a The thickness of the β-​QLLs has been determined to
bilayer, but its depth could not be quantified directly or be 9 ± 3 nm (ref.116) at 271.4 K, remarkably close to the
with any certainty. These droplets were seen to diffuse on values obtained by ellipsometry30. The next major devel-
a timescale that was incompatible with crystal growth, opment came from theory, which argued that two differ-
leading to the conclusion that the droplets were indeed ent regimes can be identified for the formation of QLLs:
of liquid-​like character. When the droplets coalesced to one in which surface condensation is associated with
form a continuous film, this QLL was termed a β-​QLL. supersaturation and another in which surface melting is
These β-​QLLs formed around 272.5 K, and they were associated with undersaturation34. Indeed, the α-​QLLs
observed to disappear, forming separated beads of water and β-​QLLs were shown to be triggered by supersatura-
(α-​QLLs), at 272 K. β-​QLLs have a flatter profile than tion conditions109 and essentially kinetically controlled,

a b
Round liquid-like
droplet (α-QLL)

Thin liquid-like
Growing layer (β-QLL)

20 μm
Elementary step

Fig. 6 | the emergence of two distinct and immiscible QLLs over an ice surface. a | Bulk-​water-like droplets form with
diameters much greater than 1,000 nm under supersaturated conditions on growth spirals. b | Optical microscopy images
taken at 272.5 K of a high-​spatial-density region festooned with screw dislocations (growth spirals), where the droplets are
seen to be located. The droplets have non-​zero contact angles, indicating distinct viscosities of the droplets from the
underlayer113. A thin liquid-​like layer wets the surface that exhibits a different viscosity from the droplet. The observation of
these three phases is a direct visualization of a quadruple point. Aside from the different depths probed by different
experimental techniques, which undoubtedly influence the range of reported quasi-​liquid layers (QLLs), the direct
observation of droplets with distinct physical properties from the wetting β-​QLL layer places another question mark
against a range of experimental measurements. Atomic force microscopy (AFM) measurements35 that have reported a
32 nm QLL depth at 272 K could be affected by the presence of droplets, which could have led to AFM tip rastering
artefacts. Sample preparation and extrinsic defects (impurities) may also affect the emergence of these droplets. Parts a
and b adapted with permission from ref.113, ACS.

www.nature.com/natrevchem
Reviews

profoundly affected by the timescale of the experiments 100 Å


and other aspects of the experimental conditions. The
equilibrium between wetting and coalescence has previ-
ously been predicted7,117. It has also been shown that the
condition of the surface (the presence of impurities, for
example) may strongly affect the onset and character of
the QLLs. Intriguingly, the structure of the surface has a
strong effect on the wettability and the mode of crystal
growth118. A greater understanding of the molecular pro-
cesses at work could lead to insights into how to control
ice growth and hence anti-​icing materials.
Evidence for inhomogeneities and partial crystal- Fig. 7 | heterogeneity in the QLL obtained through
linity in the QLL can be found in the excellent work of grand canonical molecular dynamics. The large slab
shows the top two bilayers of an ~26 nm x 31 nm slab
Hudait et al.98 and Bishop et al.44. A recent modelling
obtained at 270 K with a coarse-​grained model of water,
study by Pickering and co-​workers101 provides unprec- in which water molecules within 3.5 Å are connected by a
edented insight into the formation of QLLs, which bond. Liquid water is shown in blue, while ice is shown in
potentially may help explain some of the observations red in the first bilayer (top left) and yellow in the second
from optical microscopy made by Sazaki and various bilayer (bottom left). Even just below the bulk melting
colleagues. Pickering et al. performed comprehensive temperature (Tm), part of the outermost bilayer is still ice-​
and landmark grand canonical MD calculations (that like, while the second bilayer is mostly ice-​like. The first
is, considering the grand canonical statistical ensemble bilayer corresponds to β-​QLL , but the inhomogeneous
μVT, where μ is the chemical potential, V is the volume nature of the quasi-​liquid layer (QLL) raises questions about
of the system and T is the temperature) using the mW119 distinguishing the depth of the QLL using experimental
probes. Adapted with permission from ref.101, ACS.
potential. Simulations were performed on both the basal
and prismatic planes on slabs with surfaces as large as
~26 nm x 31 nm, making sure that the vapour pressure is extremely rough. Moreover, the fluctuations in the QLL–
equilibrated through a barostat, in light of recent exper- vapour and QLL–crystal dynamic interfaces are inde-
imental results. Several important findings stem from pendent of one another. It would be interesting to see
this work, but perhaps the finding of greatest importance whether the magnitude of the fluctuations observed by
is that the QLL is composed of crystalline portions and Pickering using the mW model and those observed
liquid-​like regions and that the QLL formation can be by Benet et al. using TIP4P-2005 are quantitatively
attributable to incomplete premelted layers. The other consistent.
new insight is that the QLL is dynamic or fluxional,
meaning that regions of the QLL that are solid were seen Chemistry at the ice–air interface
to become liquid. The connection with the homogene- We now turn to contemplate the consequences of pre-
ous nucleation of ice is clearly relevant here and merits melting on atmospheric chemistry. In particular, we
deeper investigation. The authors argue that the defi- focus on a selection of recent studies on the uptake of
nition of the thickness of the QLL becomes somewhat atmospherically relevant compounds on the surface
arbitrary because some component of the QLL is solid. of ice. Several recent excellent reviews on this topic by
At 270 K, ~70% of the first bilayer is liquid water, while Abbatt121, Dash et al.1, Kang122 and Bartels-​Rausch et al.24
only 6–15% of the second bilayer is liquid water on both exist, to which the interested reader is referred.
the prismatic and basal faces, as shown in Fig. 7. The Surface temperatures on Earth at the poles and in
QLL thickness was also seen to be measurably sensitive their vicinity vary between ~180 K and well above the
to the vapour pressure. At 260 K, Pickering et al.101 found bulk transition temperature (occasionally) in Antarctica
that 50% of the first basal bilayer is melted, in contrast to and (regularly) in the Arctic. Cloud temperatures vary
the 38% reported by Hudait; the discrepancy was sug- tremendously between 150 K and 340 K (refs123), the
gested to arise from the settings used to determine which former being representative of lower bounds for polar
molecules are liquid. Notwithstanding the ambiguity stratospheric clouds and of particular note because these
associated with defining the QLL thickness, Pickering conditions are expected to foster stacking disordered ice
et al. reported a QLL thickness of ~5–8 Å at 270 K, which (Box 2). Nevertheless, ice particles within large fractions
is clearly much smaller than the experimental results. of cloud cover expose a complete or incomplete pre-
Conde et al.18 reported a thickness in the same order melted ice layer. Recent laboratory experiments indicate
of magnitude at 270 K using their TIP4P-​ice poten- that below 160 K (154 ± 5 K, as reported elsewhere124),
tial. These data are particularly useful, as the melting amorphous solid water is the dominant phase125 of ice
temperature of the mW119 and TIP4P-​ice65 models are rather than any crystalline form and that at 180 K, con-
essentially in direct agreement with experiments within version of cubic ice (actually stacking disordered ice126)
the uncertainty of estimating the melting temperature. to hexagonal ice is slow (taking longer than one hour124).
Another hidden heterogeneity at the ice surface Heterogenous agents such as dust and organic mat-
has been revealed by Benet et al.120, who performed ter facilitate nucleation of ice at low temperatures in
MD simu­lations using the TIP4P-2005 model. Their general127, though recently it has been demonstrated
work has shown that although the QLL–vapour inter- that adding ammonium salts promotes nucleation128.
face is smooth, the underlying QLL–crystal interface is Notwithstanding the uncertainties associated with the

Nature Reviews | Chemistry


Reviews

phase of ice present in clouds and their interface satura- boundary layer. The salt-​rich interfacial region causes a
tion levels, cloud-​bound ice and surface ice present some reduction in the freezing point1,128, so the surface layer
fraction of premelted ice, and so the question of how this is expected to show an increased tendency to premelt.
changes the chemistry of the surface arises. The uptake properties of these layers with trace gases
Changes in pH at the ice surface, which clearly can under controlled, reproducible conditions remains a
influence reactivity and uptake properties, have been substantial experimental challenge.
measured to be rather small129. The surface of pristine CO2 uptake on ice has been shown to be substan-
ice has been argued to be acidic, with a pH <4.8 (ref.130) tial140 and comparable to that of SO2. Strongly enhanced
(using IR spectroscopy, isotopic labelling and combined gas uptake has been attributed to the onset of premelt-
classical and ab initio DFT MD studies, although mod- ing and the adsorption of gas in the QLL. It is there-
elling was performed on liquid water). Watkins et al.131 fore rather surprising that no recent measurements of
showed, through a comparative assessment of the surface CO2 uptake have been performed given concern over
segregation of hydronium and hydroxide in the presence the rising levels of atmospheric CO2 levels and the
of Bjerrum defects, that the trapping energies of both possible role of snow and ice in the carbon budget.
species were essentially not distinguishable. However, Again, there is a relative dearth of modelling studies
on the basis of transport barrier arguments, it was sug- in this area, and measurements of uptake over a range
gested that hydronium ions were more likely to segregate of temperatures of geophysical importance and, in
to the surface than hydroxide, whereas reactive sputter- particular, those approaching Tm would offer valuable
ing studies have detected enhanced concentrations of cross-comparisons.
hydroxide at the surface132. One consequence of the amorphous nature of the
Uptake of alkanes on the ice surface has been known pristine ice surface identified by Watkins et al.102 is that
to be anomalous, and this was attributed to the premelt- this comprises molecules that exhibit a range of polar-
ing layer133. Experimental studies at 258 K (for example, ities and binding energies even below those of pre-
in the presence of a QLL) using glancing-​angle, laser-​ melting. Watkins et al.102 reported that H2O molecules
induced fluorescence134 and classical MD and DFT sim- weakly bound at the surface of crystalline ice can be
ulations135 have shown that organic pollutants such as exothermally displaced by H2S, HCl and HF so that the
polyaromatic hydrocarbons (PAHs) accumulate on ice latter molecules are trapped on the ice surface and can
surfaces. According to the fluorescence studies, PAHs potentially undergo fast dissociative reactions141.
such as naphthalene aggregate more efficiently on ice–air While there are many examples of modelling stud-
interfaces than on water–air interfaces. The molecular-​ ies of molecules on pristine crystalline ice, the role
scale interpretation for such a preference offered by the of the premelting layer in surface chemistry has yet
modelling studies is that the disordered and denser water to be addressed comprehensively. Recently, Hudait
layer has a stronger affinity for the delocalized electron et al.98 performed a very instructive set of simulations
density of naphthalene than does ice. Photolysis of naph- examining the influence of ions and glyoxal mole-
thalene was found to be an order of magnitude faster on cules on ice surfaces at premelting temperatures. On
ice than on liquid water at 272 K and 274 K, respectively. an annual basis, 45 × 109 kg of this simple dicarbonyl
Intriguingly, the rates of photolysis on ice at 258 K and is released in the atmosphere, along with 140 × 109 kg
272 K were identical (within experimental uncertainty), of methylglyoxal142, quantities that simulations sug-
and the same was observed on liquid water at 274 K and gest equate to 10–100 parts per trillion. The major
297 K. These results highlight the need to understand conclusions from this work were that ions increase
the nature of the QLL to rationalize why there is such a the liquidity of the surface of ice over a range of tem-
marked discrepancy in rates for a 2 K difference in the peratures. The ions perturb their coordination shell
ice and water samples. One possibility is that the naph- of waters, promoting melting; however, the ions aggre-
thalene aggregates on ice, and these aggregates persist gate, and hence the correlation between ion concentra-
even on the essentially pure liquid outer layer until just tion and the fraction of liquid molecules is less than unity.
above 272 K, when these become unstable with respect The ions show diffusivity that is approximately a third
to separation on the entirely liquid layer. of that in bulk water. Conversely, the interaction between
Measurable quantities of H2O2 have been detected in the glyoxal and the ice surface is extremely weak. In
natural snow. Such a H2O2 concentration is smaller than fact, the diffusivity of the glyoxal is approximately
that detected in rainwater but large enough that its pres- 25 times larger than in liquid water, resulting in the
ence cannot be ignored in rationalizing the properties molecules surfing across the surface without perturbing
of ice. Indeed, H2O2 provides an efficient way of scav- the underlying solid or liquid structure. Clearly, this is
enging SO2 from the atmosphere to yield H2SO4 (ref.136), an area that deserves more attention, in particular from
and SO2 uptake has been measured for ice spheres137,138. the modelling community to link with the findings of
More generally, H2O2 is essential to the Fenton mecha- Sazaki et al.34.
nism of hydroxyl radical formation (see, for example,
ref.139), so understanding the uptake of this and other Outlook
environmentally important gases is an area that deserves Over the past 5–10 years, key findings have emerged from
more attention from the modelling community. single and aggregated experimental studies and an array
For sea ice, the presence of salt may be important. of modelling approaches revealing new and fundamental
Ice is very hostile to any impurities and expunges salt aspects to the now more than 160-year-old conundrum
from sea water in a superconcentrated solution as an ice of the nature of the QLL of ice. Despite clear progress

www.nature.com/natrevchem
Reviews

and, in particular, the watershed publications of Sazaki and and dynamics and have various widely discussed short-
co-workers and Elbaum and co-workers 7,32, many comings, such as incorrect prediction of dynamics and
open questions remain to be answered. The following long-​range behaviour and the neglect of polarization
discussion describes some of the most urgent topics. effects58. First-​principles techniques, such as DFT, pro-
vide a more sophisticated approach in which the full
Resolve discrepancies on QLL thickness. Despite con- electronic structure of the system is taken into account
siderable effort, differences in measured QLL thick- and polarization effects are explicitly accounted for. As
nesses remain. There are differences between distinct noted, DFT studies of ice surfaces have emerged, and
experimental techniques and disparities between experi- there is considerable scope for the application of DFT in
ment and simulation. The former is no doubt influenced this area. However, the computational cost of DFT cal-
by the distinct experimental techniques probing differ- culations (several orders of magnitude greater than that
ent features and ‘skin’ depths of the surfaces, as well as of classical force fields) means that system sizes and
different levels of impurities and in-​plane inhomoge- simulation timescales are severely limited. In addition,
neity, such as α-​QLLs. The discrepancy between the for certain properties of pure water and ice, the approx-
simulated QLL thickness and that observed by a variety imate nature of the exchange–correlation functional
of experimental techniques is much larger than could that underlies practical DFT calculations means that the
be reasonably expected. By contrast, a more consistent accuracy of DFT is often not superior to that of well-​
picture comes from simulations, chiefly from classical parameterized classical potentials57. An interesting and
models, at just below the bulk transition temperature. powerful middle ground is provided by state-​of-the-​art
Currently, it seems clear that ~2 bilayers of water melt intermolecular potentials fitted to high-​quality first-​
slightly below the transition temperature (although principles data. The fitting can be done with traditional
probably the best converged estimate at Tm − 0.25 K is approaches or through machine learning. Such poten-
1.8 nm (~4 bilayers), albeit calculated for cubic ice)87. tials (for example, many-​body potential (MB-​pol)147,
The most conservative estimate from experiment comes the Gaussian approximation potential water model148,
from X-​ray absorption and puts the QLL at a minimum neural network water149, AMOEBA150 or the thole-​type
of ∼3 nm (>7 bilayers). The work of Elbaum et al.45,46 and model (TTM) potentials151) offer very high accuracy for
latterly of Sazaki and co-​workers33,34,109–111 suggest that pure water systems, come at a fraction of the cost of DFT
the vapour pressure, typically omitted in simulations, and offer promise for improved simulations of the QLL.
may play a strong role in conditioning the surface and An additional aspect that is generally not taken into
that the influence of saturation on measurements needs account is the quantum nature of the nuclei. However,
to be better understood at a molecular level. Another the light mass of the hydrogen atoms in water means
point that requires attention is the effect of stacking that quantum nuclear effects (such as quantum delo-
disorder on premelting. Even if experiments are con- calization, tunnelling and zero-​point energy) can have
ducted on a sample that is initially pure Ih, the high an impact on the properties of water and ice152. Indeed,
vapour pressure and vapour–solid phase equilibrium it has been suggested that quantum nuclear effects can
leave open the possibility that stacking disordered ice play a role in the onset of the QLL153, and this issue
(Isd) can nucleate and that the higher free energy of Isd deserves more investigation.
could lead to lower premelting temperatures. A related The second shortcoming of simulation approaches
issue to untangle in establishing the properties of the is related to the use of simplified structural models.
QLL is capillary condensation, in which cracks and/or Despite the enormous increases in computing power
step bunches can lead to an effective supercooling esti- and improvements in algorithms and software since the
mated to be 30–50 K (ref.143), leading to ice or condensate 1990s, the structural models generally used to examine
formation in equilibrium with the vapour144. Given that ice surfaces are highly idealized and simplified versions
bulk Isd can persist to 240 K (ref.145), its influence on the of reality. More specifically, the emerging picture of a
onset temperature of premelting cannot be ignored. In heterogenous QLL calls for yet larger simulation cells
this context, from the experimental side, a tantalizing to be considered in simulation studies. Care should
prospect is the recent advent of low-​kV transition elec- be paid to understand the size and length scale of the
tron microscopy (TEM)146, which may be able to image structural heterogeneities in the QLL and any asso-
ice if sufficiently thin samples can be prepared. ciated heterogeneities in the dynamical properties of
Simulation techniques have improved enormously the QLL154. Furthermore, Bjerrum defects, vacancies
since they were first applied to the study of the QLL in and ionic, line and extrinsic defects are usually omit-
the 1990s. However, there remain clear shortcomings ted. Bjerrum defects have been shown to migrate to the
with the simulation approaches. These can be grouped surface, where they can potentially trap ions — either
into two broad categories. The first one is related to autogenetic or extrinsic ions. Vacancies are believed to
the approximations of intermolecular interactions. The be the dominant defect in the bulk, but at the surface,
vast majority of simulation of the QLL has involved sim- their effects are expected to be amplified. A potentially
ple classical intermolecular potentials such as the atom- important point made by Watkins and co-​workers102
istic TIP4P model or the coarse-​grained mW model. is that because of the highly inhomogeneous binding
As we have seen, work with these models has been energies of water bound to the outer layer of ice, the
incredibly helpful in furthering the understanding of the perfect bilayer surface can reconstruct where weakly
QLL. However, they certainly do not provide an exact bound molecules can gain energy by binding to favour-
description of the true nature of the water structure able top-​surface locations above the outermost bilayer.

Nature Reviews | Chemistry


Reviews

These events have been seen in simu­lations44,155, but fur- The advent of computationally efficient and reasona-
ther insight into their importance is needed. Simulations bly accurate models such as mW, highly accurate poten-
suggest that ionic defects in the form of both hydronium tials (for example, those developed through machine
and hydroxide ions have the tendency to migrate to the learning164) and the ever-​increasing numbers of mole-
surface. DFT approaches taking into account the likely cules that can be tackled at the DFT level165,166 open up
proton arrangements at the surface suggest that these the prospect of gaining profound insights that are of
species are attracted to distinct parts of the surface — fundamental interest but are also potentially important
hydronium to proton-​depleted areas and hydroxide to for anti-​ice applications and for other materials. Efforts
proton-​rich areas. Line defects such as screw disloca- to achieve a closer integration of experimental studies
tions and step edges, known to occur with a high spatial with modelling also promise to be an effective means to
density and potentially a source of catalytically active progress the field167.
centres, have been repeatedly observed, but few struc- Finally, we reflect on the overarching questions: what
tural models have been reported to date. Incomplete is the onset temperature of premelting and what thick-
bilayers have been analysed in a recent DFT and STM ness is the QLL at a given temperature? For both ques-
study156. Junior and colleagues performed the first DFT tions, the point of reference is important. For a fictional,
calculations on model stacking faults and partial dislo- perfect, defect-free ice Ih basal terrace, simulations with
cations157 in the crystal bulk, but further work in this well-validated models are unanimous; the onset temper-
area is needed, perhaps capitalizing on novel 1D model- ature, as determined by order parameters to distinguish
ling approaches158,159 to explore what influence all of the between crystal and liquid, is ~240 K. The thickness of
point and line defects have on premelting. The advent the QLL gets larger than the equivalent of two bilayers
of highly accurate water models58,147,153 holds promise only at >270 K. Bjerrum defects, ionic defects and line
for simulating the long length scales needed for line defects, such as step and screw dislocations (that desta-
dislocations, in particular where electronic relaxation, bilize the lattice), will increase the QLL thickness but to
mimicked through changes of polarization, is likely to an unknown extent. Furthermore, the composition of the
be an essential feature of capturing the dislocation core QLL is dynamic, as shown beautifully by Pickering and
structure. Important developments have been made co-workers101, who observed patches of liquid ebbing
in the theoretical assessment of the effect of extrinsic and flowing within layers interposing and exchanging
defects, such as impurities, on premelting160. Beaglehole with crystalline regions. In reality, the experimental
and Wilson demonstrated that brine, silicate and HF work of Elbaum et al.45,46 and latterly of Sazaki and co-​
have a measurable influence on the onset of premelting workers33,34,109–111 has highlighted the complex relation-
and the QLL thickness. Understanding the influence of ship between vapour pressure and the QLL. The QLL is
brine on premelting is clearly particularly important to now understood to be inhomogeneous and dynamic and,
predicting the crystallization and melting of sea ice in subject to vapour pressure, to consist of droplets and wet-
water of varying salinity. ting layers, α-​QLLs and β-​QLLs, respectively. The influ-
ence of vapour pressure, equilibration and α-​QLLs and
Structure and properties of the α-​QLL. The obser- β-​QLLs on reported QLL thickness and the temperature
vation of two immiscible phases of water (α-​QLL and of premelting onset is unclear. Because different meas-
β-​QLLs) atop a third phase of water, crystalline ice, is very urement techniques have different sensitivities to the
curious. The non-​wetting nature of the α-​QLL is myste­ presence of α-​QLLs and β-​QLLs, resolving discrepancies
rious, and although progress in the theory has been between different experimental techniques is probably
made by Sazaki in this direction109, the observation of a not possible at a quantitative level. Moreover, character-
quadruple-point, solid–α-​QLL/β-​QLL–vapour is clearly ization experiments to examine the influence of defects,
of fundamental interest. α-​QLLs have been observed at for example, on QLL development, are certainly needed.
dislocation cores, while β-​QLLs are observed on pitted To move forward calibrating theory with experi-
surfaces; thus, untangling these structural influences mentation, X-​ray absorption20 and SFG16 appear to be
on the formation of QLLs is important. The α-QLL techniques that currently offer prospects of quantitative
presents an enormous challenge to atomistic model- comparison and interpretation of structural changes.
ling because of the length scale of the droplets, not to Beyond further elucidating the effects of vapour pres-
mention the technical challenges of thermostatting four sure, equilibration and defects on α-​QLLs and β-​QLLs,
phases. There may be benefit in revisiting premelting in using physical probes168 or chemical probes, such as
notionally simpler systems such as metals161 or Lennard-​ gases, as a means of charting structural and notable
Jones models162 to explore whether the phenomenon of chemical changes should be an illuminating direction
two distinct QLLs is particular to water ice (noting that to pursue. For example, SO2 and H2O2 are important
droplets and films of pentane have been observed to trace gases in atmospheric chemistry139,169,170, and uptake
coexist on water8,163 but that water ice appears to be a measurements for SO2 on ice at a range of temperatures
rare, if not unique, example of a material that exhib- exist169,171, providing a means to calibrate theory and
its three distinct coexisting phases). Moreover, at low experiment and potentially make useful predictions for
vapour saturation (when the β-​QLL is expected to be the atmospheric science community, provided exper-
exclusively present), as the bulk melting temperature is iments are reproducible and models can reproduce
approached and the thickness of the QLL increases, does those experiments.
partial or complete wetting occur? Is it possible to have
a partial β-QLL? Published online xx xx xxxx

www.nature.com/natrevchem
Reviews

1. Dash, J. G., Fu, H. & Wettlaufer, J. S. The premelting 33. Sazaki, G., Zepeda, S., Nakatsubo, S., Yokoyama, E. 63. Glebov, A., Graham, A. P., Menzel, A., Toennies, J. P.
of ice and its environmental consequences. Rep. Prog. & Furukawa, Y. Elementary steps at the surface of ice & Senet, P. A helium atom scattering study of the
Phys. 58, 115–167 (1995). crystals visualized by advanced optical microscopy. structure and phonon dynamics of the ice surface.
2. Faraday, M. On certain conditions of freezing water. Proc. Natl Acad. Sci. USA 107, 19702–19707 (2010). J. Chem. Phys. 112, 11011–11022 (2000).
Athenaeum 1181, 640–641 (1850). 34. Murata, K., Asakawa, H., Nagashima, K., Furukawa, Y. 64. Avidor, N. & Allison, W. Helium diffraction as a probe
3. Pruppacher, H. & Klett, J. Microphysics of Clouds and & Sazaki, G. Thermodynamic origin of surface melting of structure and proton order on model ice surfaces.
Precipitation Vol. 18 (Springer Netherlands, 2010). on ice crystals. Proc. Natl Acad. Sci. USA 113, J. Phys. Chem. Lett. 7, 4520–4523 (2016).
4. Bailey, M. & Hallett, J. Growth rates and habits of ice E6741–E6748 (2016). 65. Abascal, J. L. F., Sanz, E., Fernández, R. G. & Vega, C.
crystals between -20C and -70C. J. Atmos. Sci. 61, 35. Döppenschmidt, A. & Butt, H. J. Measuring the A potential model for the study of ices and amorphous
514–544 (2004). thickness of the liquid-​like layer on ice surfaces with water: TIP4P/ice. J. Chem. Phys. 122, 234511 (2005).
5. Nakaya, U. Snow crystals and aerosols. J. Fac. Sci. atomic force microscopy. Langmuir 16, 6709–6714 66. Groenzin, H., Li, I., Buch, V. & Shultz, M. J. The single-​
Hokkaido Univ. 2 Phys. 4, 341–354 (1955). (2000). crystal, basal face of ice Ihinvestigated with sum
6. Hammonds, K. et al. Correction for Brumberg et al., 36. Mazzega, E., Del Pennino, U., Loria, A. & Mantovani, S. frequency generation. J. Chem. Phys. 127, 214502
Single-​crystal Ih ice surfaces unveil connection Volta effect and liquid-​like layer at the ice surface. (2007).
between macroscopic and molecular structure. J. Chem. Phys. 64, 1028–1031 (1976). 67. Sun, Z., Pan, D., Xu, L. & Wang, E. Role of proton
Proc. Natl Acad. Sci. USA 114, E5276 (2017). 37. Weyl, W. Surface structure of water and some of its ordering in adsorption preference of polar molecule
7. Elbaum, M., Lipson, S. G. & Wettlaufer, J. S. physical and chemical manifestations. J. Colloid Sci. 6, on ice surface. Proc. Natl Acad. Sci. USA 109,
Evaporation preempts complete wetting. Europhys. 389–405 (1951). 13177–13181 (2012).
Lett. 29, 457–462 (1995). 38. Fletcher, N. H. Surface structure of water and ice. 68. Tribello, G. A., Slater, B. & Salzmann, C. G. A blind
8. Bonn, D. & Ross, D. Wetting transitions. Rep. Prog. Phil. Mag. 7, 255–269 (1962). structure prediction of ice XIV. J. Am. Chem. Soc. 128,
Phys. 64, 1085–1163 (2001). 39. Fletcher, N. H. Surface structure of water and ice. II. 12594–12595 (2006).
9. Bonn, D., Eggers, J., Indekeu, J. & Meunier, J. Wetting A revised model. Phil. Mag. 18, 1287–1300 (1968). 69. Tribello, G. A. & Slater, B. Proton ordering energetics
and spreading. Rev. Mod. Phys. 81, 739–805 (2009). 40. Fletcher, N. H. Reconstruction of ice crystal surfaces at in ice phases. Chem. Phys. Lett. 425, 246–250
10. Stranski, I. N. Uber den Schmelzvorgang bei low temperatures. Phil. Mag. 66, 109–115 (1992). (2006).
nichtpolaren Kristallen [German]. 41. Buch, V., Groenzin, H., Li, I., Shultz, M. J. & Tosatti, E. 70. Kuo, J.-L., Coe, J. V., Singer, S. J., Band, Y. B. &
Naturwissenschaften 28, 425–433 (1942). Proton order in the ice crystal surface. Proc. Natl Ojamäe, L. On the use of graph invariants for
11. Wettlaufer, J. S. Ice surfaces: macroscopic effects of Acad. Sci. USA 105, 5969–5974 (2008). efficiently generating hydrogen bond topologies and
microscopic structure. Philos. Trans. R. Soc. Lond. A 42. Pan, D. et al. Surface energy and surface proton order predicting physical properties of water clusters and
357, 3403–3425 (1999). of the ice Ih basal and prism surfaces. J. Phys. ice. J. Chem. Phys. 114, 2527 (2001).
12. Frenken, J. Kinetic Theory of Liquids (Oxford Univ. Condens. Matter 22, 074209 (2010). 71. Singer, S. J. et al. Hydrogen-​bond topology and the ice
Press, 1946). 43. Pan, D. et al. Surface energy and surface proton order VII/VIII and Ice Ih/XI proton-​ordering phase transitions.
13. Golecki, I. & Jaccard, C. Intrinsic surface disorder in of ice Ih. Phys. Rev. Lett. 101, 155703 (2008). Phys. Rev. Lett. 94, 135701 (2005).
ice near the melting point. J. Phys. C Solid State Phys. 44. Bishop, C. L. et al. On thin ice: surface order and 72. Knight, C. & Singer, S. J. Prediction of a phase
11, 4229–4237 (1978). disorder during pre-​melting. Faraday Discuss. 141, transition to a hydrogen bond ordered form of ice VI.
14. Mizuno, Y. & Hanafusa, N. Studies of surface 277–292 (2008). J. Phys. Chem. B 109, 21040–21046 (2005).
properties of ice using nuclear magnetic resonance. 45. Elbaum, M. & Schick, M. Application of the theory of 73. Fan, X., Bing, D., Zhang, J., Shen, Z. & Kuo, J.-L.
J. Phys. Colloques 48, C1-511–C1-517 (1987). dispersion forces to the surface melting of ice. Phys. Predicting the hydrogen bond ordered structures of
15. Dosch, H., Lied, A. & Bilgram, J. H. Glancing-​angle Rev. Lett. 66, 1713–1716 (1991). ice Ih, II, III, VI and ice VII: DFT methods with localized
X-​ray scattering studies of the premelting of ice 46. Elbaum, M. & Schick, M. On the failure of water to based set. Comput. Mater. Sci. 49, S170–S175
surfaces. Surf. Sci. 327, 145–164 (1995). freeze from its surface. J. Phys. I 1, 1665–1668 (2010).
16. Sánchez, M. A. et al. Experimental and theoretical (1991). 74. Del Ben, M., VandeVondele, J. & Slater, B. Periodic
evidence for bilayer-​by-bilayer surface melting of 47. Dzyaloshinskii, I. E., Lifshitz, E. M. & Pitaevskii, L. P. MP2, RPA, and boundary condition assessment of
crystalline ice. Proc. Natl Acad. Sci. USA 114, The general theory of van der Waals forces. Adv. Phys. hydrogen ordering in ice XV. J. Phys. Chem. Lett. 5,
227–232 (2017). 10, 165–209 (1961). 4122–4128 (2014).
17. Wei, X., Miranda, P. B. & Shen, Y. R. Surface 48. Löwen, H. & Lipowsky, R. Surface melting away from 75. Engel, E. A., Monserrat, B. & Needs, R. J. Anharmonic
vibrational spectroscopic study of surface melting of equilibrium. Phys. Rev. B 43, 3507–3513 (1991). nuclear motion and the relative stability of hexagonal
ice. Phys. Rev. Lett. 86, 1554–1557 (2001). 49. Somorjai, G. A. Introduction to Surface Chemistry and and cubic ice. Phys. Rev. X 5, 021033 (2015).
18. Conde, M. M., Vega, C. & Patrykiejew, A. The Catalysis (John Wiley & Sons, 1994). 76. Engel, E. A., Monserrat, B. & Needs, R. J. Vibrational
thickness of a liquid layer on the free surface of ice as 50. Materer, N. et al. Molecular surface structure of a low-​ effects on surface energies and band gaps in
obtained from computer simulation. J. Chem. Phys. temperature ice Ih(0001) crystal. J. Phys. Chem. 99, hexagonal and cubic ice. J. Chem. Phys. 145, 044703
129, 014702 (2008). 6267–6269 (1995). (2016).
19. Beaglehole, D. & Nason, D. Transition layer on the 51. Materer, N. et al. Molecular surface structure of 77. Abascal, J. L. F. & Vega, C. A general purpose model
surface on ice. Surf. Sci. 96, 357–363 (1980). ice(0001): dynamical low-​energy electron diffraction, for the condensed phases of water: TIP4P/2005.
20. Bluhm, H., Ogletree, D. F., Fadley, C. S., Hussain, Z. total-​energy calculations and molecular dynamics J. Chem. Phys. 123, 234505 (2005).
& Salmeron, M. The premelting of ice studied with simulations. Surf. Sci. 381, 190–210 (1997). 78. Suter, M. T., Andersson, P. U. & Pettersson, J. B. C.
photoelectron spectroscopy. J. Phys. Condens. Matter 52. Kroes, G. J. Surface melting of the (0001) face of Surface properties of water ice at 150–191K studied
14, L227–L233 (2002). TIP4P ice. Surf. Sci. 275, 365–382 (1992). by elastic helium scattering. J. Chem. Phys. 125,
21. Petrenko, V. F. The Surface of Ice: Special Report 53. Weber, T. A. & Stillinger, F. H. Molecular dynamics 174704 (2006).
94-22 (US Army Corps of Engineers, Cold Regions study of ice crystallite melting. J. Phys. Chem. 87, 79. Tammann, G. Zur Überhitzung von Kristallen
Research & Engineering Laboratory, 1994). 4277–4281 (1983). [German]. Z. Phys. Chem. 68, 257 (1909).
22. Petrenko, V. F. & Whitworth, R. W. Physics Of Ice 54. Jorgensen, W. L., Chandrasekhar, J., Madura, J. D., 80. Carignano, M. A., Shepson, P. B. & Szleifer, I.
(Oxford Univ. Press, 1999). Impey, R. W. & Klein, M. L. Comparison of simple Molecular dynamics simulations of ice growth from
23. Shultz, M. J. Ice surfaces. Annu. Rev. Phys. Chem. 68, potential functions for simulating liquid water. supercooled water. Mol. Phys. 103, 2957–2967
285–304 (2017). J. Chem. Phys. 79, 926–935 (1983). (2005).
24. Bartels-​Rausch, T. et al. A review of air-​ice chemical 55. Berendsen, H. J. C., Postma, J. P. M., van Gunsteren, W. F. 81. Nada, H. & Van Der Eerden, J. P. J. M.
and physical interactions (AICI): liquids, quasi-​liquids, & Hermans, J. in Intermolecular Forces (ed. Pullman, B.) An intermolecular potential model for the simulation
and solids in snow. Atmos. Chem. Phys. 14, 331–342 (D. Reidel Publishing Company, of ice and water near the melting point: a six-​site
1587–1633 (2014). 1981). model of H2O. J. Chem. Phys. 118, 7401–7413
25. Li, Y. & Somorjai, G. A. Surface premelting of ice. 56. Hayward, J. A. & Reimers, J. R. Unit cells for the (2003).
J. Phys. Chem. C 111, 9631–9637 (2007). simulation of hexagonal ice. J. Chem. Phys. 106, 82. Abascal, J. L. F., Fernández, R. G., Vega, C. &
26. Limmer, D. T. Closer look at the surface of ice. 1518–1529 (1997). Carignano, M. A. The melting temperature of the six
Proc. Natl Acad. Sci. USA 113, 12347–12349 57. Gillan, M. J., Alfè, D. & Michaelides, A. Perspective: site potential model of water. J. Chem. Phys. 125,
(2016). how good is DFT for water? J. Chem. Phys. 144, 166101 (2006).
27. Maeno, N. Z. U. & Nishimura, H. The electrical 130901 (2016). 83. Ambler, M., Vorselaars, B., Allen, M. P. & Quigley, D.
properties of ice surfaces. J. Glaciol. 21, 193–205 58. Cisneros, G. A. et al. Modeling molecular interactions Solid-​liquid interfacial free energy of ice Ih, ice Ic, and
(1978). in water: from pairwise to many-​body potential energy ice 0 within a mono-​atomic model of water via the
28. Kouchi, A., Furukawa, Y. & Kuroda, T. X-​ray diffraction functions. Chem. Rev. 116, 7501–7528 (2016). capillary wave method. J. Chem. Phys. 146, 074701
pattern of quasi-​liquid layer on ice crystal surface. 59. Nada, H. & Furukawa, Y. Anisotropic properties of (2017).
J. Phys. Colloques 48, C1-675–C1-677 (1987). ice / water interface: a molecular dynamics study. 84. Espinosa, J. R., Vega, C. & Sanz, E. Ice-​water
29. Dosch, H., Lied, A. & Bilgram, J. H. Disruption of the Jpn J. Appl. Phys. 34, 583–588 (1995). interfacial free energy for the TIP4P, TIP4P/2005,
hydrogen-​bonding network at the surface of Ihice near 60. Fernández, R. G., Abascal, J. L. F. & Vega, C. TIP4P/ice, and mW models as obtained from the mold
surface premelting. Surf. Sci. 366, 43–50 (1996). The melting point of ice Ih for common water models integration technique. J. Phys. Chem. C 120,
30. Furukawa, Y., Yamamoto, M. & Kuroda, T. calculated from direct coexistence of the solid-​liquid 8068–8075 (2016).
Ellipsometric study of the transition layer on the interface. J. Chem. Phys. 124, 144506 (2006). 85. McBride, C., Vega, C., Sanz, E., MacDowell, L. G.
surface of an ice crystal. J. Cryst. Growth 82, 61. Bolton, K. & Pettersson, J. B. C. A. Molecular & Abascal, J. L. F. The range of meta stability of
665–677 (1987). dynamics study of the long-​time ice ih surface ice-​water melting for two simple models of water.
31. Furukawa, Y. & Nada, H. Anisotropic surface melting dynamics. J. Phys. Chem. B 104, 1590–1595 Mol. Phys. 103, 1–5 (2005).
of an ice crystal and its relationship to growth forms. (2000). 86. Vega, C., Martin-​Conde, M. & Patrykiejew, A.
J. Phys. Chem. B 101, 6167–6170 (1997). 62. Smit, W. J., Versluis, J., Backus, E. H. G., Bonn, M. & Absence of superheating for ice Ih with a free surface:
32. Elbaum, M., Lipson, S. G. & Dash, J. G. Optical study Bakker, H. J. Reduced near-​resonant vibrational a new method of determining the melting point of
of surface melting on ice. J. Cryst. Growth 129, coupling at the surfaces of liquid water and ice. different water models. Mol. Phys. 104, 3583–3592
491–505 (1993). J. Phys. Chem. Lett. 9, 1290–1294 (2018). (2006).

Nature Reviews | Chemistry


Reviews

87. Limmer, D. T. & Chandler, D. Premelting, fluctuations, 114. Sazaki, G., Asakawa, H., Nagashima, K., Nakatsubo, S. 143. Fukuta, N. Activation of atmospheric particles as ice
and coarse-​graining of water-​ice interfaces. J. Chem. & Furukawa, Y. Double spiral steps on Ih ice crystal nuclei in cold and dry air. J. Atmos. Sci. 23, 741–750
Phys. 141, 18505C (2014). surfaces grown from water vapor just below the melting (1966).
88. Malkin, T. L. et al. Stacking disorder in ice I. point. Cryst. Growth Des. 14, 2133–2137 (2014). 144. Campbell, J. M., Meldrum, F. C. & Christenson, H. K.
Phys. Chem. Chem. Phys. 17, 60–76 (2015). 115. Thürmer, K. & Nie, S. Formation of hexagonal and Observing the formation of ice and organic crystals in
89. Moore, E. B. & Molinero, V. Structural transformation cubic ice during low-​temperature growth. Proc. Natl active sites. Proc. Natl Acad. Sci. USA 114, 810–815
in supercooled water controls the crystallization rate Acad. Sci. USA 110, 11757–11762 (2013). (2017).
of ice. Nature 479, 506–508 (2011). 116. Murata, K. I., Asakawa, H., Nagashima, K., Furukawa, Y. 145. Kuhs, W. F., Sippel, C., Falenty, A. & Hansen, T. C.
90. Moore, E. B. & Molinero, V. Ice crystallization in water’s & Sazaki, G. In situ determination of surface tension-​ Extent and relevance of stacking disorder in ‘ice Ic’.
‘no-​man’s land’. J. Chem. Phys. 132, 244504 (2010). to-shear viscosity ratio for quasiliquid layers on ice Proc. Natl Acad. Sci. USA 109, 21259–21264 (2012).
91. Hudait, A. & Molinero, V. What determines the ice crystal surfaces. Phys. Rev. Lett. 115, 256103 (2015). 146. Zhang, D. et al. Atomic-​resolution transmission
polymorph in clouds? J. Am. Chem. Soc. 138, 117. Bar-​Ziv, R. & Safran, S. A. Surface melting of ice electron microscopy of electron beam–sensitive
8958–8967 (2016). induced by hydrocarbon films. Langmuir 9, crystalline materials. Science 359, 675–679 (2018).
92. Moore, E. B. & Molinero, V. Is it cubic? Ice 2786–2788 (1993). 147. Reddy, S. K. et al. On the accuracy of the MB-​pol
crystallization from deeply supercooled water. 118. Liu, J. et al. Distinct ice patterns on solid surfaces with many-​body potential for water: interaction energies,
Phys. Chem. Chem. Phys. 13, 20008 (2011). various wettabilities. Proc. Natl Acad. Sci. USA 114, vibrational frequencies, and classical thermodynamic
93. Li, T., Donadio, D., Russo, G. & Galli, G. Homogeneous 11285-11290 (2017). and dynamical properties from clusters to liquid water
ice nucleation from supercooled water. Phys. Chem. 119. Molinero, V. & Moore, E. B. Water modeled as an and ice. J. Chem. Phys. 145, 194504 (2016).
Chem. Phys. 13, 19807 (2011). intermediate element between carbon and silicon. 148. Bartók, A. P., Gillan, M. J., Manby, F. R. & Csányi, G.
94. Pedevilla, P., Cox, S. J., Slater, B. & Michaelides, A. J. Phys. Chem. B 113, 4008–4016 (2009). Machine-​learning approach for one- and two-​body
Can ice-​like structures form on non-​ice-like substrates? 120. Benet, J., Llombart, P., Sanz, E. & MacDowell, L. G. corrections to density functional theory: applications
The example of the K-​feldspar microcline. J. Phys. Premelting-​induced smoothening of the ice-​vapor to molecular and condensed water. Phys. Rev. B 88,
Chem. C Nanomater. Interfaces 120, 6704–6713 interface. Phys. Rev. Lett. 117, 096101 (2016). 054104 (2013).
(2016). 121. Lohmann, U., Broekhuizen, K., Leaitch, R., Shantz, N. 149. Morawietz, T., Singraber, A., Dellago, C. & Behler, J.
95. Cox, S. J., Kathmann, S. M., Slater, B. & Michaelides, A. & Abbatt, J. How efficient is cloud droplet formation of How van der Waals interactions determine the unique
Molecular simulations of heterogeneous ice organic aerosols? Geophys. Res. Lett. 31, L05108 properties of water. Proc. Natl Acad. Sci. USA 113,
nucleation. I. Controlling ice nucleation through (2004). 8368–8373 (2016).
surface hydrophilicity. J. Chem. Phys. 142, 184704 122. Kang, H. Chemistry of ice surfaces. Elementary 150. Laury, M. L., Wang, L. P., Pande, V. S., Head-​Gordon, T.
(2015). reaction steps on ice studied by reactive ion & Ponder, J. W. Revised parameters for the AMOEBA
96. Cox, S. J., Kathmann, S. M., Slater, B. & Michaelides, A. scattering. Acc. Chem. Res. 38, 893–900 (2005). polarizable atomic multipole water model. J. Phys.
Molecular simulations of heterogeneous ice 123. Stubenrauch, C. J. et al. Assessment of global cloud Chem. B 119, 9423–9437 (2015).
nucleation. II. Peeling back the layers. J. Chem. Phys. datasets from satellites: project and database initiated 151. Xantheas, S. S., Burnham, C. J. & Harrison, R. J.
142, 184705 (2015). by the GEWEX radiation panel. Bull. Am. Meteorol. Development of transferable interaction models for
97. Lupi, L., Hudait, A. & Molinero, V. Heterogeneous Soc. 94, 1031–1049 (2013). water. II. Accurate energetics of the first few water
nucleation of ice on carbon surfaces. J. Am. Chem. Soc. 124. Shilling, J. E. et al. Measurements of the vapor pressure clusters from first principles. J. Chem. Phys. 116,
136, 3156–3164 (2014). of cubic ice and their implications for atmospheric ice 1493–1499 (2002).
98. Hudait, A., Allen, M. T. & Molinero, V. Sink or swim: clouds. Geophys. Res. Lett. 33, L17801 (2006). 152. Ceriotti, M. et al. Nuclear quantum effects in water
ions and organics at the ice–air interface. J. Am. 125. Nachbar, M., Duft, D. & Leisner, T. The vapor pressure and aqueous systems: experiment, theory, and current
Chem. Soc. 139, 10095–10103 (2017). over nano-​crystalline ice. Atmos. Chem. Phys. 18, challenges. Chem. Rev. 116, 7529–7550 (2016).
99. Qiu, Y. & Molinero, V. Why is it so difficult to identify 3419–3431 (2018). 153. Paesani, F. & Voth, G. A. Quantum effects strongly
the onset of ice premelting? J. Phys. Chem. Lett. 9, 126. Murray, B. J., Malkin, T. L. & Salzmann, C. G. The influence the surface premelting of ice. J. Phys. Chem.
5179–5182 (2018). crystal structure of ice under mesospheric conditions. C 112, 324–327 (2008).
100. Shepherd, T. D., Koc, M. A. & Molinero, V. The quasi-​ J. Atmos. Sol. Terr. Phys. 127, 78–82 (2015). 154. Fitzner, M., Sosso, G. C., Cox, S. J. & Michaelides, A.
liquid layer of ice under conditions of methane 127. Kanji, Z. A. et al. Overview of ice nucleating particles. Ice is born in low-​mobility regions of supercooled
clathrate formation. J. Phys. Chem. C 116, Meteorol. Monogr. 58, 1.1–1.33 (2017). liquid water. Proc. Natl Acad. Sci. USA 116
12172–12180 (2012). 128. Whale, T. F., Holden, M. A., Wilson, T. W., O’Sullivan, D. 2009–2014 (2019).
101. Pickering, I., Paleico, M., Sirkin, Y. A. P., Scherlis, D. A. & Murray, B. J. The enhancement and suppression of 155. Pedersen, A., Wikfeldt, K. T., Karssemeijer, L., Cuppen, H.
& Factorovich, M. H. Grand canonical investigation of immersion mode heterogeneous ice-​nucleation by & Jónsson, H. Molecular reordering processes on ice
the quasi liquid layer of ice: is it liquid? J. Phys. Chem. B solutes. Chem. Sci. 9, 4142–4151 (2018). (0001) surfaces from long timescale simulations.
122, 4880–4890 (2018). 129. Wren, S. N. & Donaldson, D. J. Laboratory study J. Chem. Phys. 141, 234706 (2014).
102. Watkins, M. et al. Large variation of vacancy formation of pH at the air–ice interface. J. Phys. Chem. C 116, 156. Bockstedte, M., Michl, A., Kolb, M., Mehlhorn, M. &
energies in the surface of crystalline ice. Nat. Mater. 10171–10180 (2012). Morgenstern, K. Incomplete bilayer termination of the
10, 794–798 (2011). 130. Buch, V., Milet, A., Vacha, R., Jungwirth, P. ice (0001) surface. J. Phys. Chem. C 120, 1097–1109
103. Mantovani, S., Valeri, S., Loria, A. & Del Pennino, U. & Devlin, J. P. Water surface is acidic. Proc. Natl (2016).
Viscosity of the ice surface layer. J. Chem. Phys. 72, Acad. Sci. USA 104, 7342–7347 (2007). 157. Silva Junior, D. L. & De Koning, M. Structure and
1077–1083 (1980). 131. Watkins, M., VandeVondele, J. & Slater, B. Point energetics of extended defects in ice Ih. Phys. Rev. B
104. Pinheiro Moreira, P. A. F. & De Koning, M. Trapping of defects at the ice (0001) surface. Proc. Natl Acad. Condens. Matter Mater. Phys. 85, 024119 (2012).
hydrochloric and hydrofluoric acid at vacancies on and Sci. USA 107, 12429–12434 (2010). 158. Walker, A. M., Gale, J. D., Slater, B. & Wright, K.
underneath the ice Ih basal-​plane surface. J. Phys. 132. Kim, S., Park, E. & Kang, H. Segregation of hydroxide Atomic scale modelling of the cores of dislocations in
Chem. A 117, 11066–11071 (2013). ions to an ice surface. J. Chem. Phys. 135, 074703 complex materials part 2: applications. Phys. Chem.
105. Smit, W. J. & Bakker, H. J. The surface of ice is like (2011). Chem. Phys. 7, 3235–3242 (2005).
supercooled liquid water. Angew. Chem. Int. Ed. Engl. 133. Orem, M. W. & Adamson, A. W. Physical adsorption of 159. Walker, A. M., Gale, J. D., Slater, B. & Wright, K.
56, 15540–15544 (2017). vapor on ice. J. Colloid Interface Sci. 31, 278–286 Atomic scale modelling of the cores of dislocations in
106. Michaelides, A. & Slater, B. Melting the ice one layer (1969). complex materials part 1: methodology. Phys. Chem.
at a time. Proc. Natl Acad. Sci. USA 114, 195–197 134. Kahan, T. F. & Donaldson, D. J. Photolysis of polycyclic Chem. Phys. 7, 3227–3234 (2005).
(2017). aromatic hydrocarbons on water and ice surfaces. 160. Wettlaufer, J. S. Impurity effects in the premelting of
107. Sadtchenko, V. & Ewing, G. E. A new approach to the J. Phys. Chem. A 111, 1277–1285 (2007). ice. Phys. Rev. Lett. 82, 2516–2519 (1999).
study of interfacial melting of ice: infrared 135. Ardura, D., Kahan, T. F. & Donaldson, D. J. Self-​ 161. Frenken, J. W. M. & Van Der Veen, J. F. Observation of
spectroscopy. Can. J. Phys. 81, 333–341 (2003). association of naphthalene at the air−ice interface. surface melting. Phys. Rev. Lett. 54, 134–137 (1985).
108. Sadtchenko, V. & Ewing, G. E. Interfacial melting of J. Phys. Chem. A 113, 7353–7359 (2009). 162. Köster, A., Mausbach, P. & Vrabec, J. Premelting,
thin ice films: an infrared study. J. Chem. Phys. 116, 136. Gunz, D. & Hoffmann, M. Field investigations on the solid-​fluid equilibria, and thermodynamic properties
4686–4697 (2002). snow chemistry in central and southern California 1. in the high density region based on the Lennard-​
109. Asakawa, H., Sazaki, G., Nagashima, K., Nakatsubo, S. Inorganic ions and hydrogen peroxide. Atmos. Environ. Jones potential. J. Chem. Phys. 147 144502
& Furukawa, Y. Two types of quasi-​liquid layers on ice 24A, 1661–1671 (1990). (2017).
crystals are formed kinetically. Proc. Natl Acad. Sci. 137. Conklin, M. H. & Bales, R. C. SO2 uptake on ice 163. Del Cerro, C. & Jameson, G. J. The behavior of
USA 113, 1749–1753 (2016). spheres: liquid nature of the ice-​air interface. pentane, hexane, and heptane on water. J. Colloid
110. Inomata, M. et al. Temperature dependence of the J. Geophys. Res. Atmos. 98, 16851–16855 (1993). Interface Sci. 78, 362–375 (1980).
growth kinetics of elementary spiral steps on ice basal 138. Conklin, M. H. & Bales, R. C. Correction to “SO2 164. Nguyen, T. T. et al. Comparison of permutationally
faces grown from water vapor. Cryst. Growth Des. 18, uptake on ice spheres: liquid nature of the ice-​air invariant polynomials, neural networks, and Gaussian
786–793 (2018). interface”. J. Geophys. Res. Atmos. 99, 8351 (1994). approximation potentials in representing water
111. Asakawa, H., Sazaki, G., Nagashima, K., Nakatsubo, S. 139. Ervens, B. Modeling the processing of aerosol and interactions through many-​body expansions. J. Chem.
& Furukawa, Y. Prism and other high-​index faces of ice trace gases in clouds and fogs. Chem. Rev. 115, Phys. 148, 241725 (2018).
crystals exhibit two types of quasi-​liquid layers. Cryst. 4157–4198 (2015). 165. VandeVondele, J., Borštnik, U. & Hutter, J. Linear
Growth Des. 15, 3339–3344 (2015). 140. Ocampo, J. & Klinger, J. Adsorption of N2 and CO2 on scaling self-​consistent field calculations with millions of
112. Sazaki, G., Zepeda, S., Nakatsubo, S., Yokomine, M. ice. J. Colloid Interface Sci. 86, 377–383 (1982). atoms in the condensed phase. J. Chem. Theory
& Furukawa, Y. Quasi-​liquid layers on ice crystal 141. Bolton, K. & Pettersson, J. B. C. Ice-​catalyzed Comput. 8, 3565–3573 (2012).
surfaces are made up of two different phases. ionization of hydrochloric acid. J. Am. Chem. Soc. 123, 166. Skylaris, C-​K., Haynes, P. D., Mostofi, A. A.
Proc. Natl Acad. Sci. USA 109, 1052–1055 (2012). 7360–7363 (2001). & Payne, M. C. Introducing ONETEP: linear-​scaling
113. Sazaki, G., Asakawa, H., Nagashima, K., Nakatsubo, S. 142. Fu, T. M. et al. Global budgets of atmospheric glyoxal density functional simulations on parallel computers.
& Furukawa, Y. How do quasi-​liquid layers emerge and methylglyoxal, and implications for formation of J. Chem. Phys. 122, 084119 (2005).
from ice crystal surfaces? Cryst. Growth Des. 13, secondary organic aerosols. J. Geophys. Res. Atmos. 167. Slater, B., Michaelides, A., Salzmann, C. G. &
1761–1766 (2013). 113, D15303 (2008). Lohmann, U. A. Blue-​sky approach to understanding

www.nature.com/natrevchem
Reviews

cloud formation. Bull. Am. Meteorol. Soc. 97, 179. Howe, R. & Whitworth, R. W. A determination of the 192. Thomson, D. On Growth and Form (Canto, 1961).
1797–1802 (2016). crystal structure of ice XI. J. Chem. Phys. 90, 193. Libbrecht, K. G. The formation of snow crystals. Am. Sci.
168. Weber, B. et al. Molecular insight into the slipperiness 4450–4453 (1989). 95, 52–59 (2007).
of ice. J. Phys. Chem. Lett. 9 2838–2842 (2018). 180. Leadbetter, A. J. et al. The equilibrium low-​ 194. Shultz, M. J. Crystal growth in ice and snow. Phys.
169. Clegg, M. & Abbatt, D. Uptake of gas-​phase SO2 and temperature structure of ice. J. Chem. Phys. 82, Today 71, 35–39 (2018).
H2O2 by ice surfaces: dependence on partial pressure, 424–428 (1985). 195. Fletcher, N. H. The Chemical Physics of Ice (Cambridge
temperature, and surface acidity. J. Phys. Chem. A 181. Bjerrum, N. Structure and properties of ice. Science Univ. Press, 1970).
105, 6630–6636 (2001). 115, 385–390 (1952).
170. Huthwelker, T., Ammann, M. & Peter, T. The uptake of 182. Itoh, H., Kawamura, K., Hondoh, T. & Mae, S. Acknowledgements
acidic gases on ice. Chem. Rev. 106, 1375–1444 Molecular dynamics studies of self-​interstitials in ice The authors thank J. Abbatt, T. Bartels-​Rausch and E. Wolff
(2006). Ih. J. Chem. Phys. 105, 2408–2413 (1996). for helpful information in compiling this Review. B.S. and A.M.
171. Langenberg, S. & Schurath, U. Gas chromatography 183. König, H. Eine kubische eismodifikation [German]. thank C. Vega, M. Fitzner, C. Salzmann, E. Wang and, in par-
using ice-​coated fused silica columns: study of Z. Kristallogr. Cryst. Mater. 105, 279–286 (1943). ticular, L. Macdowell for helpful comments on this Review. The
adsorption of sulfur dioxide on water ice. Atmos. 184. Shallcross, F. V. & Carpenter, G. B. X-​Ray diffraction constructive reviews from referees are also acknowledged.
Chem. Phys. 18, 7527–7537 (2018). study of the cubic phase of ice. J. Chem. Phys. 26, The work of A.M. is supported by the European Research
172. Owston, P. G. & Lonsdale, K. The crystalline structure 782–784 (1957). Council (ERC) under the European Union’s Seventh
of ice. J. Glaciol. 1, 118–123 (1948). 185. Whalley, E. Cubic ice in nature. J. Phys. Chem. 87, Framework Programme: Grant Agreement number 616121
173. Lonsdale, K. The structure of ice. Proc. R. Soc. Lond. A. 4174–4179 (1983). (HeteroIce).
Math. Phys. Sci. 247, 424–434 (1958). 186. Malkin, T. L. et al. Structure of ice crystallized from
174. Peterson, S. W. & Levy, H. A. A single-​crystal neutron supercooled water. Proc. Natl Acad. Sci. USA 109,
Author contributions
diffraction study of heavy ice. Acta Crystallogr. 10, 1041–1045 (2012).
The authors contributed equally to all aspects of the article.
70–76 (1957). 187. Handa, Y. P., Klug, D. D. & Whalley, E. Difference in
175. Wollan, E. O., Davidson, W. L. & Shull, C. G. Neutron energy between cubic and hexagonal ice. J. Chem.
diffraction study of the structure of ice. Phys. Rev. 75, Phys. 84, 7009–7010 (1986). Competing interests
1348–1352 (1949). 188. Lupi, L. et al. Role of stacking disorder in ice The authors declare no competing interests.
176. Bernal, J. D. & Fowler, R. H. A. Theory of water and nucleation. Nature 551, 218–222 (2017).
ionic solution, with particular reference to hydrogen 189. Pauling, L. The structure and entropy of ice and of other Publisher’s note
and hydroxyl ions. J. Chem. Phys. 1, 515–548 (1933). crystals with some randomness of atomic arrangement. Springer Nature remains neutral with regard to jurisdictional
177. Tajima, Y., Matsuo, T. & Suga, H. Phase transition in J. Am. Chem. Soc. 57, 2680–2684 (1935). claims in published maps and institutional affiliations.
KOH-doped hexagonal ice. Nature 299, 810–812 190. Nakamura, H. & Cartwright, J. H. E. De nive sexangula
(1982). - a history of ice and snow - part 1. Weather 71, Reviewers information
178. Kawada, S. Dielectric dispersion and phase transition 291–294 (2016). Nature Reviews Chemistry thanks V. Molinero and the other
of KOH doped ice. J. Physical Soc. Japan 32, 1442 191. Magnus, O. Historia de Gentibus Septentrionalibus anonymous reviewer(s) for their contribution to the peer
(1972). (Rome, 1555). review of this work.

Nature Reviews | Chemistry

You might also like