arxiv:1611.03394v1 [ ] 10 nov 2016 · pdf filetheory [29{32], consistently with ... 13 hollow...

7
Snowflake growth in three dimensions using phase field modelling G. Demange, 1 H. Zapolsky, 1 R. Patte, 1 and M. Brunel 2 1 GPM, UMR CNRS 6643, University of Rouen, 76575 Saint ´ Etienne du Rouvray, France 2 CORIA UMR 6614, University of Rouen, Avenue de l’universit´ e BP 12, 76801 Saint ´ Etienne du Rouvray, France (Dated: November 11, 2016) Snowflake growth provides us with a fascinating example of spontaneous pattern formation in nature. Attempts to understand this phenomenon have led to important insights in non-equilibrium dynamics observed in various active scientific fields, ranging from pattern formation in physical and chemical systems, to self-assembly problems in biology. Yet, very few models currently succeed in reproducing the diversity of snowflake forms in three dimensions, and the link between model parameters and thermodynamic quantities is not established. Here, we report a modified phase field model that describes the subtlety of the ice vapour phase transition, through anisotropic water molecules attachment and condensation, surface diffusion, and strong anisotropic surface tension, that guarantee the anisotropy, faceting and dendritic growth of snowflakes. We demonstrate that this model reproduces the growth dynamics of the most challenging morphologies of snowflakes from the Nakaya diagram. We find that the growth dynamics of snow crystals matches the selection theory, consistently with previous experimental observations. Snowflake growth in supersaturated atmosphere is one of the most familiar, and at the same time scientifically challenging physical phenomena [1]. Beyond aesthetic fascination for their symmetric shape, snow crystals pro- vide us with a unique example of self-patterning systems [2]. Early experiments on artificial snow crystals in cold chamber led by Nakaya [3], revealed this fundamental phase transition resulted in a wide manifold of patterns, exclusively determined by supersaturation and tempera- ture. This dependency was later formalized in the me- teorological classification of Magono [4], and the snow crystals morphology diagram of Nakaya [5]. Despite this clear experimental picture, the underlying physical rules remain thoroughly debatable [6], as evi- denced by the numerous models proposed to understand the variety of snowflake shapes, such as the surface dif- fusion model of Mason et al. [7], the quasi-liquid layer approach of Lacmann et al. [8], and the layer nucleation rates theory by Nelson [9]. Alternatively, many simulation methods were devel- oped to reproduce the growth dynamics of snowflakes. First step toward the comprehension of ice crystal growth was provided by molecular dynamics simulations [1012]. Unfortunately, such simulations are still confined to space and time scales by several orders inferior to snowflake characteristic scales [4]. Significant results were also achieved using the mesoscopic approach, such as the cellular automata model of Gravner and Griffeath [13]. Though this model remarkably describes the mor- phology of snowflakes, the numerous parameters of this model can hardly be related to physical quantities. From this perspective, a 3D sharp interface model was developed by Barrett et al. [14]. Different snowflake mor- phologies were simulated. Nevertheless, the side branch- ing [6, 15], surface markings [1618], and coalescence [19] of ice crystals could not be reproduced in this framework. Besides, only small supersaturations were prospected. Consequently, only the bottom of the Nakaya diagram was explored. This limitation comes from the Laplacian approximation framework, and the numerical cost of in- terface parametrization [20]. The phase field model has the decisive advantage to overcome explicit tracking of the sharp interface, by spreading it out over a small layer [20]. As a matter of fact, it has become standard to sim- ulate the dendritic growth in alloy solidification [2123]. However, despite a prototype attempt by Barret et al. in [24], the phase field approach was never used to simulate snow crystal growth in three dimensions. Indeed, until recently, it was assumed that the phase field approach was unable to reproduce facet formation and destabiliza- tion [25]. Yet, Debierre and Karma suggested in [26], that phase field could mimic faceting using highly anisotropic surface tension. In this paper, we report the simulations of snow crys- tals growth in three dimensions, using a modified phase field model. A new surface tension anisotropy func- tion accounting for the 6-fold horizontal and 2-fold ver- tical symmetry of snowflakes was derived. A supplemen- tary anisotropy function, and anisotropic diffusion terms were also included to simulate the vertical anisotropy of snowflakes [6]. To mimic faceting in snowflakes, the 2D regularization algorithm of Eggleston et al. [27] was ex- tended to three dimensions. As a result, the model repro- duces the growth of the main snowflake morphologies of the Nakaya diagram, varying only four phenomenological parameters. It is shown that these parameters can be re- lated to physical quantities. Simulated snowflakes show excellent agreement with experimental observations [1518, 28]. Their growth satisfies the microscopic solvability theory [2932], consistently with experiments [33]. Snowflake growth in supersaturated water vapour was simulated using a phase field approach in three dimen- sions, based on a methodology developed in [21]. In this model, two coupled variables φ and u are consid- ered. φ is an order parameter referring to the ice (+1) and vapour (-1) phases. The ice/vapour interface is de- scribed by a continuous variation of φ, connecting -1 and +1. u =(c - c I sat )/c I sat is the reduced supersatu- arXiv:1611.03394v1 [physics.comp-ph] 10 Nov 2016

Upload: tranthuy

Post on 06-Mar-2018

217 views

Category:

Documents


5 download

TRANSCRIPT

Page 1: arXiv:1611.03394v1 [ ] 10 Nov 2016 · PDF filetheory [29{32], consistently with ... 13 hollow prism 3.0 0.3 3.0 0.5 14-15 capped column 5.0 ... details the formation of a hollow prism13(p

Snowflake growth in three dimensions using phase field modelling

G. Demange,1 H. Zapolsky,1 R. Patte,1 and M. Brunel2

1GPM, UMR CNRS 6643, University of Rouen, 76575 Saint Etienne du Rouvray, France2CORIA UMR 6614, University of Rouen, Avenue de l’universite BP 12, 76801 Saint Etienne du Rouvray, France

(Dated: November 11, 2016)

Snowflake growth provides us with a fascinating example of spontaneous pattern formation innature. Attempts to understand this phenomenon have led to important insights in non-equilibriumdynamics observed in various active scientific fields, ranging from pattern formation in physical andchemical systems, to self-assembly problems in biology. Yet, very few models currently succeedin reproducing the diversity of snowflake forms in three dimensions, and the link between modelparameters and thermodynamic quantities is not established. Here, we report a modified phasefield model that describes the subtlety of the ice vapour phase transition, through anisotropic watermolecules attachment and condensation, surface diffusion, and strong anisotropic surface tension,that guarantee the anisotropy, faceting and dendritic growth of snowflakes. We demonstrate thatthis model reproduces the growth dynamics of the most challenging morphologies of snowflakesfrom the Nakaya diagram. We find that the growth dynamics of snow crystals matches the selectiontheory, consistently with previous experimental observations.

Snowflake growth in supersaturated atmosphere is oneof the most familiar, and at the same time scientificallychallenging physical phenomena [1]. Beyond aestheticfascination for their symmetric shape, snow crystals pro-vide us with a unique example of self-patterning systems[2]. Early experiments on artificial snow crystals in coldchamber led by Nakaya [3], revealed this fundamentalphase transition resulted in a wide manifold of patterns,exclusively determined by supersaturation and tempera-ture. This dependency was later formalized in the me-teorological classification of Magono [4], and the snowcrystals morphology diagram of Nakaya [5].

Despite this clear experimental picture, the underlyingphysical rules remain thoroughly debatable [6], as evi-denced by the numerous models proposed to understandthe variety of snowflake shapes, such as the surface dif-fusion model of Mason et al. [7], the quasi-liquid layerapproach of Lacmann et al. [8], and the layer nucleationrates theory by Nelson [9].

Alternatively, many simulation methods were devel-oped to reproduce the growth dynamics of snowflakes.First step toward the comprehension of ice crystal growthwas provided by molecular dynamics simulations [10–12]. Unfortunately, such simulations are still confinedto space and time scales by several orders inferior tosnowflake characteristic scales [4]. Significant resultswere also achieved using the mesoscopic approach, suchas the cellular automata model of Gravner and Griffeath[13]. Though this model remarkably describes the mor-phology of snowflakes, the numerous parameters of thismodel can hardly be related to physical quantities.

From this perspective, a 3D sharp interface model wasdeveloped by Barrett et al. [14]. Different snowflake mor-phologies were simulated. Nevertheless, the side branch-ing [6, 15], surface markings [16–18], and coalescence [19]of ice crystals could not be reproduced in this framework.Besides, only small supersaturations were prospected.Consequently, only the bottom of the Nakaya diagramwas explored. This limitation comes from the Laplacian

approximation framework, and the numerical cost of in-terface parametrization [20]. The phase field model hasthe decisive advantage to overcome explicit tracking ofthe sharp interface, by spreading it out over a small layer[20]. As a matter of fact, it has become standard to sim-ulate the dendritic growth in alloy solidification [21–23].However, despite a prototype attempt by Barret et al. in[24], the phase field approach was never used to simulatesnow crystal growth in three dimensions. Indeed, untilrecently, it was assumed that the phase field approachwas unable to reproduce facet formation and destabiliza-tion [25]. Yet, Debierre and Karma suggested in [26], thatphase field could mimic faceting using highly anisotropicsurface tension.

In this paper, we report the simulations of snow crys-tals growth in three dimensions, using a modified phasefield model. A new surface tension anisotropy func-tion accounting for the 6-fold horizontal and 2-fold ver-tical symmetry of snowflakes was derived. A supplemen-tary anisotropy function, and anisotropic diffusion termswere also included to simulate the vertical anisotropy ofsnowflakes [6]. To mimic faceting in snowflakes, the 2Dregularization algorithm of Eggleston et al. [27] was ex-tended to three dimensions. As a result, the model repro-duces the growth of the main snowflake morphologies ofthe Nakaya diagram, varying only four phenomenologicalparameters. It is shown that these parameters can be re-lated to physical quantities. Simulated snowflakes showexcellent agreement with experimental observations [15–18, 28]. Their growth satisfies the microscopic solvabilitytheory [29–32], consistently with experiments [33].

Snowflake growth in supersaturated water vapour wassimulated using a phase field approach in three dimen-sions, based on a methodology developed in [21]. Inthis model, two coupled variables φ and u are consid-ered. φ is an order parameter referring to the ice (+1)and vapour (−1) phases. The ice/vapour interface is de-scribed by a continuous variation of φ, connecting −1and +1. u = (c − cIsat)/c

Isat is the reduced supersatu-

arX

iv:1

611.

0339

4v1

[ph

ysic

s.co

mp-

ph]

10

Nov

201

6

Page 2: arXiv:1611.03394v1 [ ] 10 Nov 2016 · PDF filetheory [29{32], consistently with ... 13 hollow prism 3.0 0.3 3.0 0.5 14-15 capped column 5.0 ... details the formation of a hollow prism13(p

2

ration of water vapour, where cIsat(T ) is the saturationnumber density of vapour above ice, at temperature T .Then, the growth kinetics of snowflakes is governed bytwo non conservative phase field equations. Their adi-mensionalized form is given by:

A(n)2∂tφ = −f ′(φ) + λB(n)g′(φ)u (1)

+1

2∇Γ ·

(|∇φ|2

∂[A(n)2

]∂∇φ

+A(n)2∇Γφ

)

∂tu = D∇Γ · (q(φ)∇Γ)u− Lsat

2B(n)∂tφ, (2)

where space and time are scaled by the interface widthW0, and the relaxation time τ0 respectively. In equation(1), the double well potential f(φ) = −φ2/2 + φ4/4 isthe free energy density of the ice/vapour system, at tem-perature T , and saturation concentration c = cIsat. Thesecond term in equation (1) accounts for the couplingbetween u and φ, and promotes ice phase growth in su-persaturated atmosphere, where c > cIsat. Formally, itcorresponds to the first order term in the Taylor expan-sion of the bulk potential, for c in the neighbourhood ofcIsat. The coupling constant λ can thus be computed byλ = ∂cf |csat csat/(30H), where H is the free energy bar-

rier between vapor and solid phases. g′(φ) = (1 − φ2)2

is an interpolation function introduced in [21]. Its formallows to keep the bulk potential minima at φ = ±1 forany u. To describe the strong anisotropy of snowflakegrowth along the vertical axis [6], we propose to intro-

duce the anisotropy function B(n) =√n2x + n2

y + Γ2n2z.

Here, n = −∇φ/|∇φ| is the unit normal vector of φ. Theparameter Γ > 0 governs the preference between hori-zontal and vertical growth, called the primary habit ofsnowflakes. It can be empirically related to the temper-ature [9]. The addition of the first two terms in equation(1) corresponds to the anisotropic thermodynamic driv-ing force.

The last two terms in equation (1) describe theice/vapour interface formation and propagation [21],where A(n) is the surface tension anisotropy function.In this work, a new expression of the anisotropy functionwas derived: A(n) = 1 + εxy cos(6θ) + εz cos(2ψ), where

θ = arctan(ny/nx) and ψ = arctan(√n2x + n2

y/nz) are

the polar and azimuthal angles respectively. It accountsfor both the horizontal 6-fold symmetry, and the verticalplanar symmetry of snowflakes. εxy and εz are anisotropyconstants.

The equation (2) describes the one-sided diffusion ofwater molecules in vapour, and vapour condensationon ice. Diffusion is controlled by two quantities: thefunction q(φ) = 1 − φ, which prohibits diffusion within

ice, and the reduced diffusion coefficient D = Dτ0/W20 ,

where D is the diffusion coefficient. The second termin equation (2) accounts for conversion of vapour intoice. Lsat can thus be interpreted as the rate of water

depletion in vapour, via molecule attachment at the in-terface. In this study, Lsat was treated as a numericalparameter, and its setting was conditioned by the choiceof λ. Here, Lsat ∼ 1, to allow a sufficiently fast crystalgrowth. Obviously, attachment kinetics limited growthenabling snowflake faceting is lost. Faceting is yet re-covered through the highly anisotropic interface, as sug-gested by Debierre and Karma in [26]. To account for theanisotropy of the water molecule attachment kinetics onsnow crystals [8, 9], the anisotropy function B(n) is in-troduced in the second term of equation (2) as well. Thevertical/horizontal growth preference is also consideredthrough anisotropic diffusion [11], ∇Γ = (∂x, ∂y,Γ∂z).

Constants λ, W0, τ0 and D are entangled by theasymptotic analysis mapping the phase field model toStefan sharp interface model, as shown in [21]. This

sets D = 0.6267λ, and W0 = d0λ/0.8839, where d0 isthe isotropic capillarity length. This analysis also linksthe anisotropic interface width W0A(n) to the charac-teristic time of interface propagation τ0A(n)2, whereτ0 = 0.6267λW 2

0 .The coupling constant was set to λ = 3.0 as in [22].

The horizontal anisotropy constant was set to εxy = 0.1for vertical growth (Γ ≥ 1), and εxy = 0.2 for horizon-tal growth (Γ < 1). To reproduce different snowflakemorphologies, parameters Γ, u0, Lsat and εz were varied.All study case parameters are gathered in Table I. Thesimulation method is provided in appendix B and C.

Γ u0 Lsat εz1 stellar dendrite 0.5 0.7 1.0 0.052 fern dendrite 0.5 0.8 1.6 0.053 dendritic arms plate 0.4 0.6 1.0 0.14 stellar plate 0.5 0.5 1.8 0.35 & 6 stars 0.5 0.5 1.0 0.37 double plate 0.4 0.5 1.6 0.18 sectored plate 0.4 0.5(0.8) 1.0 0.25(0.1)9 solid plate 0.25 0.4 2.0 0.210 scrolls on plate 0.4(0.2) 0.5(0.8) 1.0 0.25(0.4)11 & 12 needles 3.0 0.8 1.0 0.513 hollow prism 3.0 0.3 3.0 0.514-15 capped column 5.0(0.2) 0.8 1.0(2.0/1.0) 0.5(0.4)

TABLE I. Parameters used in the simulation of different mor-phologies of snowflakes. Parenthesis correspond to the casewhere parameter were changed during simulations.

Two limit cases of snowflake growth are displayed inFig. 1. Fig. 1(a) shows the different stages of the growthof a fernlike dendrite 2 (p. 59 of [15]), and Fig. 1(b)details the formation of a hollow prism 13 (p 64-66 of[15]).

For both cases, the growth stages are very similar tothe experimental growth kinetics obtained by Libbrecht[34]: the snowflake first aligns on the equilibrium Wulffshape, until the Mullins-Sekerka instability [35] occursfor a critical crystal radius. This instability is related tothe Berg effect [36], stating that the supersaturation fieldaround a faceted snow crystal is largest at facet edges.

Page 3: arXiv:1611.03394v1 [ ] 10 Nov 2016 · PDF filetheory [29{32], consistently with ... 13 hollow prism 3.0 0.3 3.0 0.5 14-15 capped column 5.0 ... details the formation of a hollow prism13(p

3

t = 0 50 100 250 500

(a) Fern dendrite 2

t = 0 10 100

(b) Hollow prism 13

FIG. 1. Isosurface representation (φ = 0) of snowflake growth,using the software Blender for visual rendering. Top: ferndendrite 2. The growth mechanism is essentially two dimen-sional [6]. First, the initial disk grows into a transient flatfaceted hexagon. Then, the branching instability occurs inthe horizontal plane (t = 50), for a critical crystal radius.The resulting faceted dendrites are equipped with developedfaceted side branches. Bottom: prism 13. The seed growsvertically into a prism aligned on the Wulff shape. Then thebasal facet breaks, and a hollow is formed at t = 10. Time isindicated in τ0 unit.

Instability thus occurs when the snowflake reaches a crit-ical radius [37], and kinetic effects at corners overcomethe highly anisotropic surface tension [6].

In Fig. 1(a), the simulated fernlike dendrite morphol-ogy differs from the non axisymmetric shape predictedby the theory of Brener [32], usually used to describedendrites with cubic symmetry [23, 38]. We suggest thisis due to the strong vertical anisotropy flattening, andto faceting, which limits the formation of a horizontallacuna at the tail of dendrites. The obtained featheryshape with a transversally sharp tip rather resembles Fu-rukawa’s experimental ice crystals [16]. Besides, experi-mental snowflakes display characteristic surface patterns,such as ridges [13], and flat basal planes forming steps[16]. Such patterns are also reproduced by our modelin Fig. 1(a). At low temperatures, surface nucleationand spiral growth are the leading growth processes onthe basal faces for horizontal growth [9]. Though suchmechanisms are not explicitly included in our model, thepresence of surface steps in our simulations evokes theterrace growth resulting from nucleation, experimentallyobserved on ice crystals [18]. It can be underlined thatthe succession of flat basal planes on both real and sim-ulated snowflakes, reflects the main stages of snowflakehistory. For instance, the hexagonal Wullf shape before

branching instability is clearly memorized in Fig. 1(a).This is consistent with real snowflakes observation, pro-vided in appendix A.

Fig. 2 displays the time evolution of snowflake size L,defined as the arm length for horizontal growth (blue),and the column length for vertical growth (red).

0 50 100 150 200 2500

20

40

60

Time t (τ0 unit)

SizeL

(W0

unit)

0.2 0.3 0.4 0.6

0.02

0.05

0.1

0.2

u0

V(W

0/τ0

unit)

FIG. 2. Snowflake size L (W0 units) vs. time t (τ0 units)(marks), for u0 = 0.6. Two growth regimes appear, with atransition at t = 50 for vertical growth, and t = 80 for hori-zontal growth (dots). First regime: L(t) ∝ tα (ρ = 0.99), withα = 0.62 for horizontal growth (line), and α = 0.63 for ver-tical growth (dashes). Second regime: L(t) ∝ At (ρ = 0.99),with A = 0.12 for horizontal growth (line), and A = 0.41 forvertical growth (dashes). Insert: tip velocity V (W0/τ0 units)vs. reduced supersaturation u0 (marks). V = A′u0 (ρ = 0.99)with A′ = 0.62 for horizontal growth (line), and A′ = 0.62 forvertical growth (dashes). Blue: horizontal growth (parame-ters 2 except u0 varied from 0.2 to 0.6). Red: vertical growth(parameters 11 except u0 varied from 0.2 to 0.6).

It can be seen that two growth regimes appear. After afirst growth regime, when both dynamics are similar, theneedle growth accelerates, while plate-like growth slowsdown. In the case of horizontal growth, Libbrecht ex-perimentally noted in [39], that this change of regimecoincides with full faceting occurrence on the prismaticface, causing growth to decelerate. In the case of verti-cal growth, it was also observed in [39], that the growthacceleration was related to the formation of a vicinal sur-face at the needle tip, which enhances water attachmentat the tip. During the first regime, the crystal growth isslower than diffusion, and the system satisfies the Lapla-cian approximation. Within this framework, Algrem etal. found in [40], that arm growth should display theself similar scaling behaviour L(t) ∝ t3/5. This is con-firmed by our simulations. It is interesting to note thatthe disparity of our simulations with the t1/2 diffusivelaw of Zener et al. for spherical precipitates [41], is dueto surface tension.

In the second regime, the growth kinetics is linear (lineand dashes in Fig. 2), consistently with the selectiontheory [17, 31]. The associated tip velocity V is thus theslope of this linear function. Using the same procedure

Page 4: arXiv:1611.03394v1 [ ] 10 Nov 2016 · PDF filetheory [29{32], consistently with ... 13 hollow prism 3.0 0.3 3.0 0.5 14-15 capped column 5.0 ... details the formation of a hollow prism13(p

4

for different u0, V versus u0 could be plotted in the insert.It appears that V is proportional to u0. Therefore, thegrowth velocity of both faceted and non faceted dendritessatisfies the universal law [29].

Γ < 1 Γ ≥ 1

1 - stellar dendrite 2 - fern dendrite 11 - needle

3 - dendrite plate 4 - stellar plate 12 needles cluster

5 - simple star 6 - 12-star 13 - hollow prism

7 - double plate 8 - sectored plate 14 - capped col. a

9 - solid plate 10 - scrolls on plate 15 - capped col. b

FIG. 3. Simulated snowflake classification, using the softwareBlender for visual rendering. Γ < 1 gives the flat snow crystals1 to 9. Γ ≥ 1 gives columns 11 & 13. Primary habit switchis achieved by shifting Γ during simulations. Γ from 3.0 to0.2 at t = 70, gives capped columns 14 and 15. Γ from 0.4to 3.0 gives scrolls on plate 10. For Γ < 1, no facet breakinghappens when u0 ≤ 0.4 (solid plate 9). Branching instabilityappears for u0 = 0.5 (snowflakes 3 to 8), and side branchingrequires u0 ≥ 0.7 (dendrites 1 and 2). For Γ > 1, u0 = 0.3gives a prism 13, and u0 = 0.8 induces vertical branching withneedles 11 and 12. Lsat = 1.0 gives a thick side branching forthe stellar dendrite 1. Rising Lsat to Lsat = 1.6 refines sidebranches with the fern dendrite 1. Higher values of Lsat de-stroy side branching, and lead to branches maximum facetingwith the stellar plate 4 (Lsat = 1.8).

It is consistent with previous experimental [33], andnumerical observations [14, 26]. Langer et al. argued in[29], that snowflake dendrites frontmost tip is molecularlyrough for both non-faceted and faceted dendrites. Theattachment kinetics can thus be considered isotropic withcircular symmetry.

Fig. 3 compiles equilibrium shapes of simulatedsnowflakes, using the parameters in table I. An achieve-ment of our model is to recover the principal snowflakemorphologies of the Nakaya diagram [5], depending ex-clusively on four parameters: Γ, u0, Lsat and εz.

The first parameter Γ fully determines the primaryhabit in our model. We suggest this parameter maybe fitted on the yield between prismatic and basal at-tachment coefficients [42]. This yield accounts for thealternation of dominating growth mechanisms with thetemperature. Two theories are proposed to explain thisalternation of growth preference: surface nucleation forlow temperatures [9], and quasi-liquid layer growth caus-ing surface roughening [8] near the melting point [6].Then, Hertz-Knudsen relation for facet normal velocity[17] makes supersaturation u0 the priming parameter infacet instability and growth velocity [9]. In the phasefield model, the impact of u0 is reinterpreted as the ther-modynamic driving force, and higher values also favourbranching instability and faster growth [20]. The param-eter Lsat controls the compactness of branching for nonfaceted growth [25]. Here, low values foster branchinginstability, while raising Lsat drives the system towardquasistatic diffusion, and fosters a stronger faceting. Asfor the parameter εz, it mainly influences the formationof surface patterns for horizontal growth.

Especially interesting are two model examples. First,the double plate 7 (p. 75 in [15]), results from the sand-wich instability of a flat snowflake, and the twelve armsstar 6 was simulated by the aggregation of two simplestars, with a 30◦ tilt [19].

Finally, simulated snowflake morphology depends ontemperature and density excess over vapour/water equi-librium ρWsuper = ρ − ρWsat (g.m−3). It is quantitativelyconsistent with the Nakaya diagram (see [43]). Us-ing u0 = (ρWsuper + ρWsat − ρIsat)/ρ

Isat, where ρIsat is the

vapour/ice equilibrium density, and fitting Γ on [42], oursimulations predict plate formation near (T, ρWsuper) =(−19, 0.15) and (−12, 0.4), versus (−19, 0.15), (−12, 0.4),but also (−10, 0.18) in the Nakaya diagram. Simulateddendrites occur near (−16, 0.4), versus (−15, 0.4) and(−15, 0.25) for experiments. Finally, hollow columns areformed at (−8, 0.6) in our simulations, versus (−8, 0.6)and (−8, 0.3) in the Nakaya diagram. However, smallsupersaturations ρIsat < ρ < ρWsat are beyond reach forour model. This shortcoming is complementary to themodel of Barrett et al. [14]. Indeed, the latter is boundto ρ ≤ ρWsat or T → 0◦C, due to the Laplacian approx-imation for small supersaturations [6]. Our model onthe contrary requires larger values for Lsat, and greatercomputational resources are needed to reach such super-saturations.

Page 5: arXiv:1611.03394v1 [ ] 10 Nov 2016 · PDF filetheory [29{32], consistently with ... 13 hollow prism 3.0 0.3 3.0 0.5 14-15 capped column 5.0 ... details the formation of a hollow prism13(p

5

In this paper, we have shown that the proposed mod-ified phase field model is able to reproduce the com-plex dynamics of snowflake growth. Benefiting fromthe modularity and versatility of the phase field model,this approach can be extended to describe the growthof snowflakes in the real atmospheric conditions. It cannotably be equipped with stochastic effects, for the asym-metric growth of more exotic morphologies of snowflakesin the Nakaya diagram. Continuum fluid dynamics canalso be easily introduced in the kinetics equation equa-tion (1) and equation (2), to simulate the air flow aroundthe snowflake during its fall from clouds [44]. Besides,this approach is not restricted to snowflake growth. Thedeveloped approach opens a new way to answer numer-ous outstanding questions concerning dendritic growthin wide range of materials. We can cite, for instance,the faceted dendritic growth observed in pure isotac-tic polystyrene films, or the snowflake-like growth of

graphene monocrystal [45].

ACKNOWLEDGEMENTS

This work was supported in part by the grant fromLABEX ”Cistic”. The simulations were performed at theCentre de Ressources Informatiques de Haute-Normandie(CRIHAN) and at the IDRIS of CNRS.

AUTHOR CONTRIBUTIONS

G.D. and H.Z. developed the model and wrote themanuscript. R.P. participated in the code development.All authors analyzed the data, discussed the results andcommented on the manuscript.

[1] B. Pamuk, J. M. Soler, R. Ramırez, C. P. Herrero, P. W.Stephens, P. B. Allen, and M.-V. Fernandez-Serra, Phys.Rev. Lett. 108, 193003 (2012).

[2] P. Ball, Nature Materials 15, 1060 (2016).[3] U. Nakaya, I. Sato, and Y. Sekido, Journal of the Fac-

ulty of Science, Hokkaido Imperial University. Ser. 2 2, 1(1938).

[4] C. Magono, W. Chung, et al., Journal of the Faculty ofScience, Hokkaido University. Series 7, Geophysics 2, 321(1966).

[5] U. Nakaya, Compendium Meteor , 207 (1951).[6] K. G. Libbrecht, Reports on progress in physics 68, 855

(2005).[7] B. J. Mason, Contemporary physics 33, 227 (1992).[8] T. Kuroda and R. Lacmann, Journal of Crystal Growth

56, 189 (1982).[9] J. Nelson, Philosophical Magazine A 81, 2337 (2001).

[10] H. Nada and Y. Furukawa, Journal of crystal growth 169,587 (1996).

[11] Y. Furukawa and H. Nada, The Journal of PhysicalChemistry B 101, 6167 (1997).

[12] J. Benet, P. Llombart, E. Sanz, and L. G. MacDowell,Physical Review Letters 117, 096101 (2016).

[13] J. Gravner and D. Griffeath, Physical Review E 79,011601 (2009).

[14] J. W. Barrett, H. Garcke, and R. Nurnberg, PhysicalReview E 86, 011604 (2012).

[15] K. G. Libbrecht, Ken Libbrecht’s field guide to snowflakes(Voyageur Press, 2006).

[16] Y. Furukawa and W. Shimada, Journal of crystal growth128, 234 (1993).

[17] Y. Saito, Statistical physics of crystal growth, Vol. 2(World Scientific, 1996).

[18] G. Sazaki, S. Zepeda, S. Nakatsubo, E. Yokoyama, andY. Furukawa, Proceedings of the National Academy ofSciences 107, 19702 (2010).

[19] K. Kikuchi and H. Uyeda, Atmospheric Research 4748,169 (1998).

[20] I. Singer-Loginova and H. Singer, Reports on progress inphysics 71, 106501 (2008).

[21] A. Karma and W.-J. Rappel, Physical Review E 53,R3017 (1996).

[22] J. Ramirez, C. Beckermann, A. Karma, and H.-J.Diepers, Physical Review E 69, 051607 (2004).

[23] A. Cartalade, A. Younsi, and M. Plapp, Computers &Mathematics with Applications 71, 1784 (2016).

[24] J. W. Barrett, H. Garcke, and R. Nurnberg,ZAMM-Journal of Applied Mathematics and Me-chanics/Zeitschrift fur Angewandte Mathematik undMechanik 93, 719 (2013).

[25] R. Kobayashi, Physica D: Nonlinear Phenomena 63, 410(1993).

[26] J.-M. Debierre, A. Karma, F. Celestini, and R. Guerin,Physical Review E 68, 041604 (2003).

[27] J. J. Eggleston, G. B. McFadden, and P. W. Voorhees,Physica D: Nonlinear Phenomena 150, 91 (2001).

[28] K. G. Libbrecht, arXiv preprint arXiv:1211.5555 (2012).[29] J. Langer, R. Sekerka, and T. Fujioka, Journal of Crystal

Growth 44, 414 (1978).[30] J. Langer and H. Muller-Krumbhaar, Acta Metallurgica

26, 1681 (1978).[31] M. B. Amar and E. Brener, Physical review letters 71,

589 (1993).[32] E. Brener, Journal of crystal growth 166, 339 (1996).[33] K. G. Libbrecht, T. Crosby, and M. Swanson, Journal

of crystal growth 240, 241 (2002).[34] K. G. Libbrecht, “Snowcrystals.com,” http://www.

snowcrystals.com.[35] W. W. Mullins and R. Sekerka, Journal of applied physics

35, 444 (1964).[36] W. Berg, in Proceedings of the Royal Society of Lon-

don A: Mathematical, Physical and Engineering Sciences,Vol. 164 (The Royal Society, 1938) pp. 79–95.

[37] T. Fujioka and R. Sekerka, Journal of Crystal Growth24, 84 (1974).

[38] H. Singer and J. Bilgram, EPL (Europhysics Letters) 68,240 (2004).

[39] K. G. Libbrecht, arXiv preprint arXiv:1602.08528(2016).

Page 6: arXiv:1611.03394v1 [ ] 10 Nov 2016 · PDF filetheory [29{32], consistently with ... 13 hollow prism 3.0 0.3 3.0 0.5 14-15 capped column 5.0 ... details the formation of a hollow prism13(p

6

[40] R. Almgren, W.-S. Dai, and V. Hakim, Physical reviewletters 71, 3461 (1993).

[41] C. Zener, Journal of Applied Physics 20, 950 (1949).[42] K. G. Libbrecht, arXiv preprint arXiv:0810.0689 (2008).[43] J. S. Marshall and M. P. Langleben, Journal of Meteo-

rology 11, 104 (1954).[44] J.-H. Jeong, N. Goldenfeld, and J. A. Dantzig, Phys.

Rev. E 64, 041602 (2001).[45] E. Whiteway, W. Yang, V. Yu, and M. Hilke, arXiv

preprint arXiv:1509.01579 (2015).

Appendix A: comparison of simulations with realsnowflakes

(a) Simple star from [34] (b) Simple star

(c) Stellar dendrite from [34] (d) Stellar dendrite

(e) Sectored plate from [34] (f) Sectored plate

FIG. 4. Comparison between real snowflakes photographs(left) taken from [34], and our phase field simulations(right). Visual rendering for our simulations uses the soft-ware Blender.

Appendix B: simulation method

Phase field simulations were performed using theFourier-spectral semi-implicit scheme with periodicboundary conditions. For horizontal growth, simulations

were performed with grid spacing ∆x = ∆y = ∆z = 0.8,and time step ∆t = 0.05 [22], on a 400 × 400 × 64 simu-lation box. Vertical growth simulations required a moreprecise discretization. Grid spacing was thus reduced to∆x = ∆y = ∆z = 0.4, and ∆t = 0.01 as in [22], on a 128× 128 × 256 simulation box. Snow crystal growth wasinitiated by a circular-disk shape germ (φ = 1) of radiusR = 8∆x, within water vapour (φ = −1) of homogeneousreduced supersaturation u0 > 0. A new regularizationmethod for high interfacial energy leading to crystal miss-ing orientations was also adapted from two dimensions[27], to three dimensions. This allowed to overcome therestriction for the critical values of the anisotropy con-stants and choose εxy > 1/35 and εz > 1/3, as requiredto achieve faceting [20].

Appendix C: 3D faceting algorithm

Faceting requires the anisotropy constants εxy and εzvalues in A(n), to exceed 1/35 for 6-fold horizontal sym-metry, and 1/3 for 2-fold vertical symmetry [20]. Abovesuch values, metastable and unstable crystal orientationsn where the stiffness becomes negative, emerge. A(n)must hence be modified, so that the stiffness remainspositive. A preliminary step is to determine missing ori-entations. This can be done by detecting sign change inthe stiffness. This causes a local convexity inversion inthe 2D polar plot of 1/A(θ, φ0), for any fixed azimuthalangle φ0, or in the polar plot of 1/A(θ0, φ) for any fixedpolar angle θ0. This reads A+∂2

θθA = 0 and A+∂2φφA = 0

respectively, providing us with the maximum anisotropyconstants εmxy and εmz before missing orientation:

εmxy =

1 + εz cos(2φ)

35

εmz =1 + εxy cos(6θ)

3.

(C1)

Restricting the study to θ ∈ [−π/6, π/6] and φ ∈ [0, π]due to rotation periodicity of A(θ, φ), missing θ orienta-tions lie within a φ-dependant angular sector [−θm(φ) +kπ/3, θm(φ) + kπ/3] (k ∈ Z) for εxy > εmxy. Same goesfor φ orientations lying within [−φm(θ) + kπ, φm(θ) +kπ] for εz > εmz . Following the procedure developedby Eggleston in [27], θm and φm respectively satisfy∂θ[cos(θ)/A(θ, φ)] = 0 and ∂φ[cos(φ)/A(θ, φ)] = 0, lead-ing to:

6εxy sin(6θm) cos(θm)

sin(θm)= (1 + εxy cos(6θm) + εzcos(2φ))

2εz sin(2φm) cos(φm)

sin(φm)= (1 + εxy cos(6θ) + εzcos(2φ

m)) .

(C2)The regularization step finally consists in replacingA(θ, φ) by a simple trigonometric function for missingorientation, as suggested by Debierre et al. [26]:

Page 7: arXiv:1611.03394v1 [ ] 10 Nov 2016 · PDF filetheory [29{32], consistently with ... 13 hollow prism 3.0 0.3 3.0 0.5 14-15 capped column 5.0 ... details the formation of a hollow prism13(p

7

A(n) =

Aθ(θ, φ) = A1(φ) +B1 cos(θ), |θ| < θm, |φ| ≥ φm

Aφ(θ, φ) = A2(θ) +B2 cos(φ), |θ| ≥ θm, |φ| < φm

α(θ, φ)Aθ + (1− α(θ, φ))Aφ, |θ| < θm, |φ| < φm

1 + εxy cos(6θ) + εz cos(2ψ) otherwise.(C3)

Coefficients A1(φ), B1, A2(θ), B2 and α(θ, φ) are set toensure continuity of A(n) and its derivatives:

A1(φ) = 1 + εxy cos(6θm) + εz cos(2φ)− 6εxy sin(6θm)

sin(θm)cos(θm)

B1 =6εxy sin(6θm)

sin(θm)

A2(θ) = 1 + εxy cos(6θ) + εz cos(2φm)− 2εz sin(2φm)

sin(φm)cos(φm)

B2 =2εz sin(2φm)

sin(φm)

α(θ, φ) =|θ − θm|√

(θ − θm)2 + (φ− φm)2.

(C4)