liquid snowflake formation in superheated ice

Upload: bryan-graczyk

Post on 03-Apr-2018

223 views

Category:

Documents


0 download

TRANSCRIPT

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    1/115

    Liquid Snowflake Formation in

    Superheated Ice

    Matthew G. Hennessy

    St Hughs College

    University of Oxford

    A thesis submitted for the degree of

    Master of Science in Mathematical Modelling and Scientific

    Computing

    Trinity 2010

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    2/115

    This one is for my parents, Peter and Debra Hennessy.

    Without their love and support,

    I would not have made it to where I am today.

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    3/115

    Acknowledgements

    First and foremost, I would like to extend my sincerest thanks to Dr

    Stephen Peppin and Dr Richard Katz, who supervised this project. Their

    genuine interest in all aspects of the project, in combination with their

    endless enthusiasm, made carrying out this research the most enjoyable

    experience, even when this involved a week of consecutive experimental

    failures.

    I am very thankful to Prof Grae Worster, as he not only hosted me during

    my visit to the University of Cambridge, but he also provided invaluable

    ideas with regards to the linear stability analysis. Furthermore, I would

    like to acknowledge Colin Macdonald for his discussions and his advice

    about the level set method. I am also grateful for the various discus-

    sions Ive had about liquid snowflakes with Dr Rob Style and Prof John

    Wettlaufer over the course of the summer.

    Finally, I would like to thank my friend Iain Moyles, who was not only

    brave enough to read the first draft of this thesis, but whose success on

    the west coast is a constant source of motivation to always stay on top of

    my academic game.

    This publication was based on work supported in part by Award No KUK-

    C1-013-04, made by King Abdullah University of Science and Technology

    (KAUST), and with funding from the Natural Sciences and EngineeringResearch Council of Canada (NSERC).

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    4/115

    Abstract

    Liquid snowflakes, also called Tyndall figures, are small volumes of water

    that resemble the classical shape of a snowflake that is composed of ice.

    They initially form as cylindrical discs of liquid in superheated crystals

    of ice. Several experimental studies have shown that liquid snowflakes

    acquire their shape through an instability that occurs in the circular in-terface.

    This thesis aims to begin a novel mathematical investigation of the evolu-

    tion of two dimensional, planar, liquid snowflakes. In particular, several

    linear stability analyses are carried out in order to gain insight into the

    physical mechanisms that drive the instability of the circular interface.

    The nonlinear evolution of liquid snowflakes is studied via direct numeri-

    cal simulation. The numerical methods are based on the level set method.

    The results from the mathematical analysis show that the interfacial insta-

    bility is driven by superheating in the solid and is inhibited by diffusion

    of thermal energy in the liquid. Furthermore, the existence of a large

    wavenumber mode that grows faster than all of the other modes is shown.

    The wavenumber of this mode corresponds to the wavenumber of the inter-

    face when it becomes unstable, and its theoretical value is shown to agree

    with experimental measurements. Moreover, the functional relationship

    between this wavenumber and the experimental parameters is found using

    asymptotic techniques.

    Linear stability theory predicts that the circular interface of a liquid

    snowflake eventually restabilizes due to large temperature gradients in

    the water. As this phenomenon has yet to be observed experimentally, a

    hypothesis is formulated which suggests that growth in the vertical direc-

    tion is required for keeping the temperature of the water small enough to

    prevent thermal diffusion from stabilizing the interface. This hypothesis

    is also the first of its kind to explain why liquid snowflakes remain as

    circular cylinders if their vertical growth is inhibited.

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    5/115

    Contents

    1 Introduction 1

    1.1 Experimental Observations . . . . . . . . . . . . . . . . . . . . . . . . 2

    1.2 A Similar Physical System . . . . . . . . . . . . . . . . . . . . . . . . 5

    1.3 Aims of This Thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

    1.3.1 Outline of Thesis . . . . . . . . . . . . . . . . . . . . . . . . . 8

    2 Mathematical Model 9

    2.1 Evolution Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

    2.1.1 Volumetric Heating due to Absorption of Radiation . . . . . . 11

    2.2 Conditions at the Ice-Water Interface . . . . . . . . . . . . . . . . . . 12

    2.2.1 The Gibbs-Thompson Condition . . . . . . . . . . . . . . . . . 132.2.1.1 Surface Energy Anisotropy . . . . . . . . . . . . . . 15

    2.2.2 The Stefan Condition . . . . . . . . . . . . . . . . . . . . . . . 16

    2.3 Non-dimensionalization . . . . . . . . . . . . . . . . . . . . . . . . . . 18

    2.3.1 Summary of Non-dimensional Equations . . . . . . . . . . . . 21

    2.4 The Quasi-steady Approximation . . . . . . . . . . . . . . . . . . . . 21

    3 Linear Stability Analysis 23

    3.1 An Illustrative Example . . . . . . . . . . . . . . . . . . . . . . . . . 23

    3.1.1 Stability of the Basic State . . . . . . . . . . . . . . . . . . . . 253.2 The Effects of Superheating due to Radiation Absorption . . . . . . . 29

    3.2.1 The Basic State . . . . . . . . . . . . . . . . . . . . . . . . . . 29

    3.2.2 Stability of the Basic State . . . . . . . . . . . . . . . . . . . . 31

    3.2.3 Physical Interpretation of Results . . . . . . . . . . . . . . . . 38

    3.3 Linear Stability of a Liquid Disc . . . . . . . . . . . . . . . . . . . . . 43

    3.3.1 Stability of the Basic State . . . . . . . . . . . . . . . . . . . . 48

    3.3.1.1 Asymptotic Analysis . . . . . . . . . . . . . . . . . . 51

    3.3.1.2 Physical Interpretation of Results . . . . . . . . . . . 54

    i

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    6/115

    3.3.2 Extensions to Three Spatial Dimensions . . . . . . . . . . . . 58

    4 Numerical Simulations 594.1 A Note on Alternative Numerical Approaches . . . . . . . . . . . . . 60

    4.2 The Level Set Method . . . . . . . . . . . . . . . . . . . . . . . . . . 61

    4.3 Implementation of the Numerical Scheme . . . . . . . . . . . . . . . . 63

    4.3.1 Updating the Temperature Profiles . . . . . . . . . . . . . . . 64

    4.3.1.1 Boundary Conditions on the Computational Domain 67

    4.3.2 Extrapolating the Interface Velocity . . . . . . . . . . . . . . . 68

    4.3.2.1 Additional Numerical Details . . . . . . . . . . . . . 70

    4.3.3 Advancing the Level Set Function in Time . . . . . . . . . . . 71

    4.3.4 Reinitializing the Level Set Function . . . . . . . . . . . . . . 72

    4.3.5 Further Implementation Details . . . . . . . . . . . . . . . . . 74

    4.4 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

    4.4.1 Validation of Code . . . . . . . . . . . . . . . . . . . . . . . . 74

    4.4.2 Long-term Stabilization of the Interface . . . . . . . . . . . . . 75

    4.4.3 Effects of Radiation Absorption in the Liquid Phase . . . . . . 76

    4.4.4 Breaking the Symmetry . . . . . . . . . . . . . . . . . . . . . 78

    4.4.5 Surface Energy Anisotropy . . . . . . . . . . . . . . . . . . . . 78

    5 Conclusion 81

    5.1 Summary of Key Results . . . . . . . . . . . . . . . . . . . . . . . . . 81

    5.2 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82

    5.3 Final Thoughts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83

    A Experimental Details 85

    A.1 Growing Large Crystals of Ice . . . . . . . . . . . . . . . . . . . . . . 85

    A.2 The General Experimental Setup . . . . . . . . . . . . . . . . . . . . 88

    A.2.1 Estimation of The Radiation Intensity . . . . . . . . . . . . . 89

    B High Order WENO Discretizations 91

    C Constructing the Initial Level Set Function 95

    Bibliography 100

    ii

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    7/115

    List of Figures

    1.1 A six-fold symmetric liquid snowflake that forms in superheated ice.

    The dark spot is a vapour bubble. This figure was produced using the

    experimental setup described in Appendix A. . . . . . . . . . . . . . . 2

    1.2 The evolution of three liquid snowflakes, one of which shows a re-

    markable hexagonal symmetry. The overlap occurs because each liquid

    snowflake is at a different depth below the surface of the ice. These

    images were produced using the experimental setup described in Ap-

    pendix A. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

    1.3 The hexagonal structure of an ice crystal. The planes that are formed

    by the hexagons are called basal planes and the axis that is normal

    to the basal plane is called the c-axis. Circles denote the approximate

    location of oxygen atoms. . . . . . . . . . . . . . . . . . . . . . . . . 5

    2.1 A beam of light with area A passing through a slab of thickness z. . 12

    2.2 The Gibbs-Thomson effect leads to the transfer of heat from liquid

    fingers to solid fingers, which drives the interface to a planar state. . . 13

    2.3 The surface energies associated with a liquid drop on a substrate that

    is surrounded by a solid. . . . . . . . . . . . . . . . . . . . . . . . . . 14

    2.4 An interface moving through a solid with velocity vn. Also shown are

    the diffusive fluxes, Jl

    and Js, of the liquid and the solid, respectively. 17

    3.1 A planar interface (dashed) growing into a solid with velocity v along

    the z-axis. Linear stability theory will predict the growth of perturba-

    tions to the interface, shown as a solid line. . . . . . . . . . . . . . . . 25

    3.2 A sinusoidally perturbed interface requires superheating in the solid

    and supercooling in the liquid in order to sustain the growth of both

    solid fingers and liquid fingers. . . . . . . . . . . . . . . . . . . . . . . 28

    iii

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    8/115

    3.3 When the interface of a liquid snowflake goes unstable, liquid fingers

    grow into the superheated ice. The initial circular interface can still

    be seen as a faint line from which the fingers grow out of. The image

    was obtained using the experimental setup described in Appendix A . 28

    3.4 Schematic diagram of a solid-liquid system being heated by a light

    source which travels with the interface. . . . . . . . . . . . . . . . . . 30

    3.5 Temperature profiles of a solid being superheated by incident radiation

    as viewed from a frame that moves with the planar interface. Param-

    eter values for the non-dimensional numbers can be found in Table

    2.2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

    3.6 The growth rate of perturbations to a planar interface as a functionof the non-dimensional wavenumber (top) and the dimensional wave-

    length (bottom). Also shown is the asymptotic growth rate for large

    Stefan numbers (stars). Parameter values can be found in Table 2.1

    and Table 2.2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

    3.7 The critical wavenumber (top) and the critical absorption coefficient

    (bottom) as functions of the critical beam radius. Also shown are the

    regions of stability and instability. Parameter values can be found in

    Table 2.1 and Table 2.2. . . . . . . . . . . . . . . . . . . . . . . . . . 383.8 Isotherms of the solid phase in front of two interfaces with different

    curvature. The interface on the left has wavenumber a = 2, and the

    interface on the right has wavenumber a = 6. The isotherms bunch

    together near the liquid fingers in the right figure, indicating that dif-

    fusion is enhanced near these regions. This figure was created using

    solution in (3.12) with the parameters given in Table 2.2. . . . . . . . 39

    3.9 The wavelength of the most unstable mode as a function of the beam

    intensity at the surface of the ice. Parameter values are given in Table

    2.1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

    3.10 The critical radiation intensity as a function of the absorption coeffi-

    cient for various depths under the ice surface. Also depicted are the

    regions where a planar water-ice interface is expected to become unsta-

    ble. Absorption coefficients that are less than 1 m1 correspond to the

    visible spectrum, whereas the larger values correspond to the infrared

    spectrum. Parameter values are given in Table 2.1. . . . . . . . . . . 42

    iv

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    9/115

    3.11 Evolution of the temperature profile associated with a growing liquid

    disc. In each figure, the liquid region is given by r < R. Parameter

    values can be found in Table 2.2. . . . . . . . . . . . . . . . . . . . . 46

    3.12 The growth rate of a liquid disc surrounded by superheated solid. The

    overdot denotes differentiation with respect to (dimensional) time. The

    parameters used in this figure are those of Table 2.1 and they corre-

    spond to an ice-water system. . . . . . . . . . . . . . . . . . . . . . . 47

    3.13 The relative growth rates of perturbations with wavenumbers from

    n = 2 (circles) to n = 35 (stars) as a function of the disc radius. The

    non-dimensional parameters that were used to generate this figure are

    given in Table 2.2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 503.14 Critical basic state radii as a function of the perturbation wavenumber.

    The top figure shows the radii where perturbations first become un-

    stable. The bottom figure shows the cut-off radii, where perturbations

    begin to decay. Parameters can be found in Table 2.1. . . . . . . . . . 53

    3.15 The top figure shows, for each wavenumber n, the basic state radius

    where this mode achieves its maximum relative growth rate. The bot-

    tom figure shows the maximum relative growth rate of each mode.

    Parameters are from Table 2.1. . . . . . . . . . . . . . . . . . . . . . 543.16 Isotherms of the liquid region, which bunch together near solid fingers

    that extend into the liquid. This leads to additional transfer of heat

    which causes these fingers to melt and retreat. Figure was produced

    using (3.33) and by setting R = 1, R = 0.1, and n = 6. Parameter

    values can be found in Table 2.2. . . . . . . . . . . . . . . . . . . . . 55

    3.17 The ratio of the perturbation amplitude to the basic state radius

    as a function of time. Breakdown of the linearized equations is ex-

    pected when this ratio becomes greater than one. The perturbation

    has wavenumber n = 7 and the initial conditions were R(0) = 0.2 and

    R(0) = 0.001. Parameter values can be found in Table 2.2. . . . . . . 56

    4.1 The solid-liquid interface can be viewed as the zero level set of some

    function . The signs of this function can also be associated with a

    particular phase. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

    v

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    10/115

    4.2 The two-stage extrapolation process that is used to define a velocity

    for the level set equation. The first stage (left) involves computing

    the temperature gradient of the liquid near the interface and then

    propagating these values into the solid region. The second phase (right)

    computes the temperature gradient of the solid near the interface and

    then sends these values into the liquid region. . . . . . . . . . . . . . 69

    4.3 Growth of a circular disc as predicted by the level set method. As the

    number of grid points N increases, the curves converge to the exact

    curve at a rate that is at least first order in h = 1/(N 1). Theabsolute error is the difference in the final radii. . . . . . . . . . . . . 75

    4.4 The restabilization of the interface, which was initially assumed tobe a superposition of a circle and a wavenumber seven mode. The

    simulation ran until t = 1, and the curves are 25 time steps apart.

    Moreover, 250 grid points were used in each direction. . . . . . . . . . 76

    4.5 The effect of increasing the relative absorption coefficient, which is

    equivalent to decreasing the amount of radiation that is absorbed by

    the liquid. The curves in the top three figures are 20 time steps apart.

    The bottom figures show the corresponding temperature fields at the

    final time-step. Each computation used 200 grid points in each direction. 774.6 The growth of a liquid snowflake that does not start from the centre

    of the radiation beam. The figures on the left show the evolution of

    the interface, whereas the figures on the right show the evolution of

    the temperature profile. The simulation was carried out using 250 grid

    points in each direction and the absorption coefficient in the liquid was

    taken to be zero. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

    4.7 The effects of six-fold anisotropy in the surface energy. The interfaces

    are shown every 20 time steps and 400 grid points were used in each

    direction. Furthermore, to reduce stabilization, the absorption of ra-

    diation in the liquid was neglected. . . . . . . . . . . . . . . . . . . . 80

    A.1 Liquid dendrites growing out of a grain boundary. . . . . . . . . . . . 87

    A.2 Using polarizers, the individual crystals of ice can be seen as dark

    spots. In this case, the ice block is composed of one large crystal, with

    several smaller crystals scattered throughout. . . . . . . . . . . . . . . 88

    A.3 Schematic diagram of the general experimental setup. . . . . . . . . . 90

    vi

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    11/115

    C.1 The parameter s is chosen to minimize the distance between the point

    (x0, z0) that lies on the interface and the point (x, z). . . . . . . . . . 99

    C.2 An example of the log of the residual error 0 1 when the initialinterface is assumed to be hexagonal. The exact parametrization of

    the interface is given by x0 = cos(s)[1+(1/35) cos(6s)], z0 = sin(s)[1+

    (1/35) cos(6s)], where 0 s < 2. A 100 100 grid was used in thecomputation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99

    vii

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    12/115

    List of Tables

    2.1 Typical parameter values for an ice-water system. A reference of [E]

    denotes a value that was measured experimentally. . . . . . . . . . . . 18

    2.2 Summary of the non-dimensional numbers and their typical values,

    which are based on a temperature scale of T 2 K and a lengthscale of l 1 mm. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

    viii

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    13/115

    Chapter 1

    Introduction

    In 1858, John Tyndall made a remarkable discovery when studying the effects of

    sunlight on ice slabs from Norway and Wenham Lake [41]. Using a convex lens to

    focus a concentrated beam of sunlight into the interior of the ice slab, Tyndall noticed

    the appearance of a vast number of lustrous spots that resembled bubbles of air. Upon

    further investigation, however, Tyndall found that these were not bubbles of air, but

    were, in fact, volumes of water. Moreover, each volume of water had a remarkable

    flower-like shape that exhibited a high degree of symmetry (see Figure 1.1). Tyndall

    subsequently called these bright spots liquid flowers, but today they are commonly

    referred to as Tyndall figures, internal melt figures, and liquid snowflakes [19, 29, 38].Despite being discovered over 150 years ago, the precise physical mechanisms that

    govern the formation of liquid snowflakes are still unknown. However, it is known

    that a small portion of the electromagnetic radiation that passes through the ice will

    be absorbed, which, in turn, will raise the temperature of the ice. Although the exact

    temperature rise will depend on the intensity and the wavelength of the incident light

    [19], in many cases it is possible to bring the interior of ice to its equilibrium melting

    temperature of approximately 273 K. If the block of ice is actually a single crystal

    that is nearly free of defects and impurities, then mass heterogeneous nucleation ofwater will not occur when the ice reaches 273 K and the interior of the ice will not

    melt. Instead, the ice will become superheated, that is, its temperature will be raised

    above its equilibrium melting temperature. However, it is extremely rare for ice to be

    completely free of defects, and therefore, some nucleation of water will occur. In fact,

    it is this water which gives rise to liquid snowflakes. The liquid that forms absorbs

    nearly 100 times more radiation than the surrounding ice [19, 44], and therefore, its

    temperature is expected to rise rapidly. To moderate the temperature rise of the

    liquid, heat exchange with the surrounding ice is expected to occur. This can cause

    1

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    14/115

    Figure 1.1: A six-fold symmetric liquid snowflake that forms in superheated ice. The dark

    spot is a vapour bubble. This figure was produced using the experimental setup describedin Appendix A.

    the ice to melt, which effectively leads to a growth of the liquid snowflake. This is

    merely a hypothesis, however, and it has yet to be validated.

    1.1 Experimental Observations

    There are several characteristics of liquid snowflakes that have been consistently ob-

    served in experiments. While some of these can be explained using the physical and

    molecular properties of water and ice, the cause of many of them remain unknown.

    The purpose of this section is to give an overview of the common experimental obser-

    vations, and to provide explanations for the ones that are well understood. Many of

    these observations were made by the author using the experimental setup described

    in Appendix A, and these observations agree with those which have been documented

    in [29, 38].

    One of the most prominent features of a liquid snowflake is the appearance of

    a dark circle near its centre (see Figure 1.1). This circle is, in fact, a bubble ofwater vapour that forms because of the density difference between water and ice. In

    particular, when a mass of ice melts, the resulting water has less volume than the

    ice, and hence the bubble forms to fill the void. Originally, it was suspected that

    this bubble was composed of air that had been trapped in the ice. However, Tyndall

    disproved this hypothesis by placing a piece of ice that contained a large number of

    liquid snowflakes in beaker of warm water [41]. Any air bubbles that were contained

    in the liquid snowflakes would be liberated when the ice melted and they would then

    rise to the surface of the water. No such bubbles were observed, however.

    2

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    15/115

    A particularly interesting aspect of liquid snowflakes is the morphology of the

    interface that forms between the ice and the water. Generally, liquid snowflakes begin

    as circular discs that grow outwards in a radially symmetry manner. If the intensity

    of the incident radiation is sufficiently strong, however, the circular interface becomes

    unstable and small amplitude perturbations that resemble sine waves appear. These

    perturbations can subsequently grow into vast dendritic structures, as Figure 1.2

    shows. An interesting question arises about wavenumber selection at the interface

    when it first becomes unstable. From Figure 1.2 it can be seen that corrugations

    with a large wavenumber grow from the circular interface, which is surprising because

    surface energy effects at the interface tend to inhibit the growth of such modes. Thus,

    it seems likely that there is an alternative physical mechanism that is promoting thegrowth of high wavenumber modes, and that this mechanism is competing with the

    dampening effects of surface energy.

    Circular liquid snowflakes have been observed to grow into hexagons, and the

    hexagonal interface can also become unstable. Hexagonal liquid snowflakes are par-

    ticularly interesting because there appear to be two modes of instability that can

    occur. The first mode is analogous to the circular liquid snowflake case where the

    entire interface develops small, outward growing corrugations. This case can be seen

    in Figure 1.2. The second mode of instability involves the vertices of the hexagongrowing into large lobes, so that the overall shape resembles a flower with six petals.

    Figure 1.1 shows this case. In both of these cases, the six-fold symmetry is usually

    preserved.

    There are a number of open questions about the physical mechanisms that govern

    the evolution of the interface. For example, the precise relationship between the

    intensity of the radiation and the shape of the interface has yet to be determined.

    Obtaining quantitative data about this relationship is not trivial, however, as it is

    quite common to see liquid snowflakes that have highly different morphologies, but

    which are in close proximity to each other.

    When liquid snowflakes form in a single crystal of ice, they are always oriented in

    the same direction [38, 41]. More specifically, if the liquid snowflakes are imagined

    as planar objects with zero thickness, then their planes are parallel to one another.

    This characteristic is intimately related to the underlying microscopic structure of

    the ice crystal in which the liquid snowflakes are contained. In the absence of defects

    and impurities, the crystalline lattice of ice at standard temperature and pressure is

    a hexagonal prism [19], as Figure 1.3 shows. The planes that the hexagons lie in are

    called basal planes, and the direction normal to basal plane is called the c-axis. The

    3

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    16/115

    Figure 1.2: The evolution of three liquid snowflakes, one of which shows a remarkablehexagonal symmetry. The overlap occurs because each liquid snowflake is at a different

    depth below the surface of the ice. These images were produced using the experimentalsetup described in Appendix A.

    4

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    17/115

    c

    Figure 1.3: The hexagonal structure of an ice crystal. The planes that are formed by thehexagons are called basal planes and the axis that is normal to the basal plane is called the

    c-axis. Circles denote the approximate location of oxygen atoms.

    anisotropy of an ice crystal causes melting to occur preferentially in the basal plane,

    and therefore, liquid snowflakes grow into these planes. Since the basal planes in a

    single crystal of ice are parallel, it follows that the liquid snowflakes will then have

    the same orientation. Because of this fact, liquid snowflakes are often used to deduce

    the orientation of the basal plane and the c-axis of an ice crystal [38].

    Although most of the growth of a liquid snowflake occurs in the basal plane, there

    is a small component of growth along the c-axis. Experiments have shown that growth

    in this direction is, in fact, necessary for the ice-water interface to become unstable.

    In particular, [25] found that if liquid snowflakes did not grow beyond a critical height

    of 10 m, then no perturbations formed on the interface. Similar data is shown in

    [38], however, no attempt is made to measure the critical height in this case. Neither

    of these works were able to provide any explanation for why this critical height exists.

    1.2 A Similar Physical System

    The formation of liquid snowflakes in superheated ice is remarkably similar to ice

    formation in supercooled water. In this latter system, highly purified water is lowered

    to a temperature below 273 K and an initial mass of ice is nucleated by some form

    of perturbation. Experiments typically show that the solid nucleus initially grows

    as a circular disc. If the amount of supercooling is not too large, then most of the

    growth will be in the macroscopic basal planes of the nucleus that are formed by

    its circular cross-sections [35]. Eventually, the circular disc of ice becomes unstable,

    resulting in a slightly corrugated interface [19]. If the water is sufficiently supercooled,

    the corrugations develop into dendrites, similar to in the case of liquid snowflakes.

    5

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    18/115

    Furthermore, [36] has shown that ice growth along the c-axis is particularly crucial for

    instabilities to form. In fact, they deduce that morphological instability is controlled

    by the thickness of the ice disc and not by its radius.

    A particularly appealing feature of the ice-growth system is that it is relatively

    simple and it doesnt require a constant source of external energy to supercool the

    water. This is in contrast to liquid snowflake experiments, which require a source of

    radiation to superheat the ice. The simplicity of the supercooled water system has

    been exploited by numerous authors which have formulated mathematical models of

    the freezing process. Many of these models aim to investigate the morphological in-

    stabilities that appear. The first of these studies was carried out in 1963 by Mullins

    and Sekerka [28], who investigated the stability of a solid sphere growing into super-cooled liquid. They found that the sphere must grow beyond a certain radius for any

    instabilities to occur, and that this radius is related to the critical radius of nucleation

    (see [10] for more information). Subsequent studies built upon the ideas of Mullins

    and Sekerka using linear stability analysis and by investigating solids with more re-

    alistic geometries. For example, [12] investigated the two dimensional evolution of a

    cylindrical disc in a three dimensional temperature field. More specifically, growth

    along the c-axis was neglected and the analysis assumed that the solid and the liquid

    had the same thermal properties. This latter assumption, in particular, fails to holdfor an ice-water system. However, this work did produce growth rates for the disc

    radius and showed how the growth rate of a perturbation depends on its wavenumber.

    Unfortunately, a lack of experimental data prevented an in-depth discussion of these

    theoretical results. The full three dimensional growth of a cylindrical disc of ice was

    studied in [45], where it was found that symmetry breaking occurs when the cylinder

    grows beyond a critical thickness and the circular faces acquire different radial growth

    rates. This, in turn, leads to the formation of a tapered cylinder. The critical thick-

    ness was also shown to be related to the morphological stability of the liquid-solid

    interface. Moreover, the critical thickness was found to be inversely proportional to

    the amount of supercooling, which compares favourably with experimental data. This

    not overly surprising, however, because the growth along the c-axis was assumed to be

    governed by slow interfacial kinetics and the analysis relied on an empirical formula

    to relate the vertical growth rate to the supercooling.

    6

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    19/115

    1.3 Aims of This Thesis

    Liquid snowflakes have been used for decades to examine the mechanical and ther-modynamic properties of ice crystals [29, 38], yet no mathematical theory has been

    developed for their formation and evolution. Perhaps this is due to a belief that the

    growth of liquid snowflakes in superheated ice is sufficiently similar to the growth of

    ice in supercooled water that the theoretical results for the latter system carry over

    to the former system. However, [38] has shown that this is not the case, and that the

    two systems can vary significantly. In particular, the dendritic growth rates of liquid

    snowflakes were measured experimentally and then, using the theory developed in

    [23], were compared to theoretical growth rates of ice dendrites in supercooled water.

    The theoretical growth rates poorly matched those which were measured experimen-

    tally. Moreover, experimental data from both systems was used to compare how the

    tip radii of liquid and solid dendrites depended on the amount of superheating in

    the ice and supercooling in the liquid, respectively. It was found that an order of

    magnitude more supercooling was required to produce an ice dendrite with the same

    tip radius as a liquid dendrite. Therefore, the two systems exhibit markedly different

    quantitative behaviour.

    The general goal of this thesis is to begin developing a theory for the formation

    of liquid snowflakes. This will be done by first building a suitable mathematical

    model that is based on well established physical principles and not on empirically

    determined relationships. Using this model, a combination of analytical techniques

    and numerical methods will be used to study the growth of liquid snowflakes and

    their morphological instabilities. In particular, linear stability analysis will be used

    to investigate the mechanisms that govern the transition from a circular interface to a

    corrugated interface. Unfortunately, linear stability theory is expected to break down

    once the corrugations become so large that the nonlinear terms in the equations can no

    longer be neglected. To study the nonlinear evolution of the interface, which includesthe formation of dendrites, numerical simulations that are based on the level set

    method will be used. Furthermore, this thesis will have an experimental component

    to it, and this will allow data to be collected that can be used for comparison and

    validation purposes. By combining analytical work with numerical simulations and

    experimental observations, it is hoped that a deeper understanding of the physics

    which govern the evolution of liquid snowflakes will be obtained and that this, in

    turn, will provide new insight into the many unanswered equations of liquid snowflake

    evolution.

    7

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    20/115

    1.3.1 Outline of Thesis

    The second chapter of this thesis will provide a detailed description of the mathemat-ical equations that are used to model the growth of liquid snowflakes. In particular,

    the governing equations and their boundary conditions will be derived and put into

    non-dimensional form. Chapter 3 will be dedicated to linear stability theory, and

    several systems of increasing complexity and realism will be analyzed. Chapter 4 will

    discuss the level set method and show the results of several numerical simulations.

    Each of these three chapters will have a large focus on the physical interpretation of

    the equations and the results that are obtained. Finally, the thesis will conclude in

    Chapter 5, where the main results will be summarized and possible areas of future

    work will be mentioned.

    8

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    21/115

    Chapter 2

    Mathematical Model

    Obtaining a complete mathematical description of the formation and the evolution of

    a liquid snowflake is a highly complex problem which would have to account for the

    nucleation of the liquid disc and the vapour bubble, and then model their subsequent

    growth. The nucleation dynamics are particularly complicated because they depend

    on the exact nature of the microscopic defects in the ice crystal. However, for the

    purposes of this study it will suffice to assume that nucleation has already occurred

    and that the initial state of the system contains a liquid disc that is surrounded by

    superheated ice. Further simplifications to a mathematical model can be obtained by

    neglecting the appearance of the vapour bubble, which can be justified using the factthat the amount of thermal interaction between the water and the vapour is expected

    to be small. In particular, by assuming the thermal properties of the vapour are

    similar to air, owing to the fact that the vapour is expected to have a low density, it

    can be shown that the thermal conductivity of water is approximately twenty times

    larger than that of air [7]. This implies there will be little heat exchange between

    these two regions. Moreover, the latent heat which is released upon vapourization of

    liquid into gas can also be neglected because the temperature of the liquid is expected

    to be small [5].By neglecting nucleation and the vapour bubble, a relatively simple mathematical

    model that describes the evolution of liquid snowflakes can be developed. Such a

    model will be based on two main parts; evolution equations for the liquid and the

    solid, and governing equations for the dynamics at the water-ice interface. This

    chapter will derive and discuss these key equations and put them into their most

    simple, yet physically enlightening form, by performing a non-dimensionalization.

    9

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    22/115

    2.1 Evolution Equations

    The primary physical principle which governs the evolution of the solid phase andthe liquid phase is that of conservation of thermal energy. The mathematical mani-

    festations of such conservation laws are well developed, and an equation representing

    the conservation of thermal energy in a phase i can be written as [16]

    t(icpiTi) + Ji = Qi, (2.1)

    where i, cpi, and Ti denote the density, the specific heat capacity, and the temperature

    of that phase, respectively, Ji is the flux of thermal energy, and Qi represents source

    and sink terms. The quantity icpiTi is the thermal energy per unit volume.The main contribution to the thermal flux in both the solid phase and the liquid

    phase is from diffusion, which arises from the tendency of heat to flow from regions of

    high temperature to regions of low temperature. This behaviour is quantified using

    Fouriers law of heat conduction [16],

    Jd = kiTi, (2.2)

    where Jd is the diffusive flux and ki is the thermal conductivity of phase i. A sec-

    ondary contribution to the flux in the liquid phase arises from advection, which is

    the transport of heat by the motion of the water within the liquid snowflake. The

    velocity of the fluid is expected to be small, however, and therefore the advective flux

    can be neglected from the model.

    The temperature of each phase is expected to vary by at most a couple of degrees,

    which allows the material properties of the system (heat capacity, density, thermal

    conductivity) to be modelled as temperature independent. Moreover, the water and

    the ice are assumed to be sufficiently homogeneous that these properties can, in fact,

    be taken as constants. As the vapour bubble is also being neglected, the conservation

    of mass requires the densities of water and ice to be equal. Using these assumptions

    and by combining (2.1) and (2.2), the governing equations for the temperature profiles

    of the two phases can be written as

    cplTlt

    = kl2Tl + Ql, (2.3a)

    cpsTst

    = ks2Ts + Qs, (2.3b)

    where the subscripts l and s denote liquid and solid, respectively. The exact form

    of the source terms Ql and Qs which, in this context, represent the absorption of

    electromagnetic radiation, will be discussed below.

    10

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    23/115

    The boundary conditions that are imposed away from the interface depend on the

    particular problem that is being studied, but in general they will be of the standard

    type (Dirichlet, boundedness, etc.).

    2.1.1 Volumetric Heating due to Absorption of Radiation

    When electromagnetic radiation is incident on water or ice, a portion of this radiation

    is absorbed and converted into thermal energy, which then leads to an increase in the

    temperature. The source of this radiation is usually from a focused beam of light, and

    therefore, at the surface of the ice, the intensity profile I is approximately Gaussian

    [26],

    I = I0e(r/rb)2 ,

    where r denotes the distance from the centre of the beam to where I is measured,

    rb is a measure of the beam radius, and I0 denotes the peak intensity of the beam

    (which occurs directly in the centre).

    At the microscopic level, the absorption of radiation occurs when the electrons

    that are bound to the constituent atoms of the medium enter an excited state by

    absorbing a photon. A reduction of photons corresponds to a reduction in the intensity

    of radiation. While absorption is a quantum mechanical phenomenon, it can be

    formulated in a classical sense using wave theory, and it can be shown that the decay

    rate of a beam as it travels through a medium is exponential [15]. This leads to an

    intensity of the form

    I(r, z) = I0 e(r/rb)

    2z, (2.4)

    where z is the distance travelled through the medium, and is the absorption coeffi-

    cient of the medium. The inverse of the absorption coefficient represents the distance

    that a beam of light can travel before its intensity is reduced by a factor of 1

    e1.

    Furthermore, the absorption coefficient is strongly dependent on the wavelength of

    light and it may vary by several orders of magnitude over a small range of the elec-

    tromagnetic spectrum.

    To derive an expression for the volumetric heating that occurs as a medium absorbs

    radiation, consider a beam of light that passes through a thin slab of material with

    a thickness z, as shown in Figure 2.1. Recalling that intensity is a measure of the

    power per unit area, the amount of power that is lost as the beam passes through

    this material is

    P = [I(r, z) I(r, z+ z)]A,

    11

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    24/115

    z

    A

    I(r, z+ z)

    I(r, z)

    Figure 2.1: A beam of light with area A passing through a slab of thickness z.

    where A is the area of the beam. The volumetric loss of power, which is equivalent to

    the volumetric heating in the material (by conservation of energy), can be found by

    expanding this expression about small z, keeping only the leading order term, and

    then dividing by Az. The result is

    Q = Iz

    = I(r, z),

    which is readily obtained using (2.4).

    All of the problems that are investigated in this thesis are posed in a particular

    two dimensional basal plane that is assumed to be a distance d below the surface

    of the ice. Therefore, the radiation that is absorbed by the water and by the ice in

    this plane has decayed by the same amount, in particular, by a factor of esd. By

    defining a new incident intensity as I1 = I0esd, the volumetric heating terms that

    appear in the evolution equations for the temperature (2.3) can be written as

    Ql = l I1 e(r/rb)

    2

    , (2.5a)

    Qs = s I1 e(r/rb)

    2

    . (2.5b)

    2.2 Conditions at the Ice-Water Interface

    Closing the mathematical model for liquid snowflake evolution requires additionalboundary conditions at the ice-water interface. The first condition is to impose con-

    tinuity of the temperature field across the interface,

    Ts = Tl = TI,

    where TI denotes the interfacial temperature. This temperature can be calculated

    from the Gibbs-Thomson condition, which uses thermodynamic arguments to estab-

    lish the equilibrium melting temperature at a curved interface. The second and final

    condition, called the Stefan condition, describes the velocity of the interface by con-

    sidering local conservation of energy.

    12

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    25/115

    Heat flow

    TI > Tm

    TI < Tm

    Solid

    Liquid

    Figure 2.2: The Gibbs-Thomson effect leads to the transfer of heat from liquid fingers tosolid fingers, which drives the interface to a planar state.

    2.2.1 The Gibbs-Thompson Condition

    One of the most distinguishing features of a liquid snowflake is its shape, which is

    the result of a highly curved interface that forms between the solid phase and theliquid phase. The intermolecular forces that act between the water and the ice at this

    interface lead to a net surface energy, which nature tends to minimize. The physical

    mechanism that drives this minimization is a shift in the phase equilibria at curved

    interfaces that causes thermal energy to diffuse from liquid fingers to solid fingers [43].

    This, in turn, causes the liquid fingers to freeze and the solid fingers to melt, thus

    reducing the surface area of the interface and hence minimizing its surface energy.

    Figure 2.2 shows a schematic of this effect.

    This process can be formalized by first considering a two-dimensional liquid drop

    that is placed on a substrate of length L. Furthermore, assume that the drop is

    surrounded by a solid (see Figure 2.3) and let the surface energy per unit length

    of the liquid-substrate interface, the solid-substrate interface, and the solid-liquid

    interface be denoted by ls, ss, and , respectively. The net surface energy is then

    given by

    E = ls(b a) + ss(L + a b) +b

    a

    1 + h2x dx,

    where a and b denote the contact points of the liquid on the substrate, and h denotes

    the height of the drop as a function of distance along the substrate. Furthermore, thesubscript on h denotes differentiation with respect to x. The height of the drop will

    be that which minimizes the surface energy E subject to a constraint which conserves

    the initial area of the drop,

    A =

    ba

    h(x) dx.

    This minimization problem can be solved using the Euler-Lagrange equations [32],

    and it can be shown that h will satisfy

    = , (2.6)

    13

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    26/115

    a b

    Solid

    Liquid

    ss

    ls

    h(x)

    x

    Substrate

    Figure 2.3: The surface energies associated with a liquid drop on a substrate that issurrounded by a solid.

    where is the curvature of the interface given by

    = ddx

    hx

    1 + h2x

    and is a Lagrange multiplier that corresponds to the jump in pressure across the

    interface [18]. This expression can be generalized to arbitrary geometries by writing

    the curvature as the divergence of the normal vector n at the interface, = n.Moreover can be replaced by the surface energy per unit area if a two dimensional

    interface is being considered.

    The jump in the pressure at the interface is a result of a shift in the thermodynamic

    equilibrium, which can be quantified using the Gibbs free energy . If Tm and pm

    denote the equilibrium temperature and pressure of a planar solid-liquid interface1,

    respectively, then at a curved interface, the Gibbs free energy looks approximately

    like

    (TI, pI) (Tm, pm) + p

    (p pm) + T

    (TI Tm), (2.7)

    where pI is the pressure at the interface. In particular, pI = ps at the solid side of

    the interface, and pI = pl at the liquid side, where pl and ps denote the interfacial

    pressures of the solid and the liquid, respectively. The partial derivatives of the Gibbsfree energy are well known from classical thermodynamics (see [8], for example) and

    are given by

    p=

    1

    ,

    T= s,

    where s is the entropy per unit mass. By demanding that the Gibbs free energy is

    continuous across the interface,

    (TI, pl) = (TI, ps),

    1

    For a water-ice interface, these values are well known as Tm = 273.15 K and pm = 101.3 kPa[3].

    14

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    27/115

    and using the approximation in (2.7), the difference in pressure across the interface

    is given by

    ps pl = (sl ss)(TI Tm),where sl and ss denote the entropy per unit mass of the liquid phase and the solid

    phase, respectively, and the assumption that ice and water have equal densities has

    been used. The change in entropy across the interface is given by sl ss = L/Tm,where L is the latent heat of fusion. Therefore, the jump in interfacial pressure can

    be written in terms of physical variables as

    ps pl = LTIT

    m

    1 .Recalling that the Lagrange multiplier in (2.6) is exactly the jump in the pressure

    across the interface, the interfacial temperature can be written as

    TI = Tm

    1

    L

    , (2.8)

    which is known as the Gibbs-Thomson condition. For this equation to accurately

    model the above physical process, the curvature of a liquid finger that extends into

    the solid must be taken as negative.

    2.2.1.1 Surface Energy Anisotropy

    Surface energy effects arise because of the microscopic interactions between molecules

    near the interface. The molecules in a solid, however, are typically arranged in a well

    defined crystal lattice whose geometrical properties can give rise to strong anisotropies

    in the surface energy. The crystalline structure of ice, for example, has six fold

    symmetry [19], which is inherited in the shape of both conventional snowflakes and

    liquid snowflakes (see Figure 1.1). This implies that the effects of surface energy

    anisotropy, which are due to microscopic properties of the solid, are propagated to

    the macroscale and should be incorporated into the mathematical model.

    An extended Gibbs-Thomson condition can be derived by assuming the surface

    energy per unit area is a function of the normal vector at the solid-liquid interface,

    = (n), which models the fact that the surface energy will depend on the orientation

    of the interface relative to the structure of the crystal. The resulting surface energy

    minimization problem can solved with the Euler-Lagrange equations to obtain

    +

    d2

    d2

    = , (2.9)

    15

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    28/115

    where denotes the angle that the normal vector makes with the horizontal axis, is

    the curvature of the interface, and denotes the pressure jump across the interface.

    Using this expression, the Gibbs-Thomson condition can be written as

    TI = Tm

    1 +

    L

    , (2.10)

    where the subscripts with denote differentiation with respect to that variable.

    The surface energy of a crystal with n fold symmetry is typically written as [10]

    = 0

    1 +

    nn2 1 cos(n)

    ,

    where n is a measure of the anisotropy. The case when |n| < 1 is considered weakanisotropy, as the expression + will have the same sign for all angles . On

    the other hand, strong anisotropy occurs when |n| > 1. When the system is inthermodynamic equilibrium, (2.9) implies that ( + ) is constant. Thus, when

    the anisotropy is weak, the interface will look approximately circular, as expected.

    When the anisotropy is strong, however, the interface will be composed of straight

    line segments and it will resemble a polygon, which can be deduced using the Wulff

    construction (see [10, 33]).

    Surface energy anisotropy is a particularly complicated phenomenon and it willonly be considered in the numerical simulations. Therefore, in the subsequent equa-

    tions and analysis it can be assumed that the surface energy is isotropic, unless

    otherwise stated.

    2.2.2 The Stefan Condition

    The Gibbs-Thomson condition provides an expression for the temperature at the

    interface. However, the location of the interface is not known a priori, and it must be

    solved as part of the problem. In this case, the velocity of the interface can be deducedfrom the conservation of energy. In order for a drop of water that is surrounded by

    ice to grow, the neighbouring solid must absorb enough thermal energy to stretch

    the interface and to overcome the latent heat of fusion. The thermal energy that is

    required to do this is supplied by the net heat flux into that region.

    To formalize this argument, consider an interface that is growing through the ice

    with a normal velocity vn, as shown in Figure 2.4. The total mass of ice that melts

    in a short interval of time t is given by m = A(t)vn t, where A is the area of

    the interface at time t. The latent heat that is required to melt this mass of ice is

    wl = LAvn t. Furthermore, if the area of the interface at time t + t is A(t + t),

    16

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    29/115

    t

    t+ t

    Liquid

    vn

    A(t+ t)

    A(t)

    Jl

    Js

    n

    Solid

    Figure 2.4: An interface moving through a solid with velocity vn. Also shown are thediffusive fluxes, Jl and Js, of the liquid and the solid, respectively.

    then the work that is required to stretch the interface is ws = [A(t + t) A(t)].Thus, the total energy that is required to advance the interface is

    w = LA(t)vn t + [A(t + t) A(t)],

    and this is supplied by thermal energy that flows into the ice via diffusion,

    wd = (klTl + ksTs) nA(t) t,

    where n denotes a unit vector that points into the solid. By conservation of energy,

    these two expressions must be equal,

    LA(t)vn t + [A(t + t) A(t)] = (klTl + ksTs) nA(t) t.

    Dividing this equation by A(t)t and then taking the limit as t 0 yields

    Lvn +

    A

    A

    t= (klTl + ksTs) n.

    From differential geometry it is known that (see [2], for example)

    1

    A

    A

    t= vn,

    which allows the energy balance equation to be written as

    (L + ) vn = klTl n + ksTs n. (2.11)

    This expression for the velocity of the interface is known as the Stefan condition, and

    it provides closure to the boundary value problem.

    17

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    30/115

    Table 2.1: Typical parameter values for an ice-water system. A reference of [E] denotes avalue that was measured experimentally.

    Parameter Description Value Unit Ref.

    Density of water/ice 1000 kgm3 [3]cpl Heat capacity of water 4181 Jkg1K1 [7]cps Heat capacity of ice 2050 Jkg1K1 [7]kl Thermal conductivity of water 0.6 Wm1K1 [7]ks Thermal conductivity of ice 2.0 Wm1K1 [7]L Latent heat of fusion 3.33 105 Jkg1 [3]

    TmMelting temperature of a

    273 K [3]planar ice-water interface

    Surface energy per unit area

    0.033 Jm2

    [19]of ice-water interfacel Absorption coefficient of water 1 104 m1 [44]s Absorption coefficient of ice 100 m

    1 [19]d Distance below ice surface 0.01 m [E]rb Radius of light beam 0.01 m [E]I0 Intensity of incident radiation 300 Wm2 [E]

    2.3 Non-dimensionalization

    The governing differential equations and their associated boundary conditions dependon a large number of physical parameters. Moreover, the magnitudes of these param-

    eters can vary significantly (see Table 2.1), and therefore, it is difficult to establish

    the primary physics that drive the evolution of liquid snowflakes. To overcome this

    difficulty, the equations can be written in terms of non-dimensional variables and

    parameters that characterize the relative importance of each term in the equations.

    To begin, a non-dimensional temperature u is defined for the solid and liquid

    regions

    ui =

    Ti

    Tm

    T ,where i is either l or s, and T sets the temperature scale of the problem. A possible

    choice for this scale is the amount of superheating that is measured in the ice, which is

    usually in the range of 0.01 K to 0.1 K [19, 38]. However, this parameter will actually

    be chosen to balance terms in the governing equations. The lengths are scaled by l so

    that x = lx, where x is non-dimensional vector of the coordinate length variables

    and l will be found later. Similarly, time is rescaled so that t = t, where is the time

    scale and t is a unitless time. With these scalings, the equations for the temperature

    18

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    31/115

    can be written in terms of characteristic time scales,

    ult =

    dl 2

    ul +

    hl e(r/)2

    ,

    ust

    =krcpr

    dl2us + r

    cpr

    hle(r/)

    2

    ,

    where the primes have been dropped, = rb/l is a rescaled beam radius, and cpr =

    cps/cpl, kr = ks/kl, and r = s/l are relative measures of the heat capacities,

    thermal conductivities, and absorption coefficients, respectively. The time scale dl

    measures the amount of time it takes heat to diffuse across a distance l in the liquid,

    and hl is the time required for the radiation to raise the temperature of the liquid

    by an amount T. Explicit expressions for these time scales are given by

    dl =cpll

    2

    kl,

    hl =cplT

    lI1,

    and the analogous time scales in the solid are ds = (cpr/kr)dl and hs = (cpr/r)hl.

    Inserting typical parameter values shows that diffusion is approximately seven times

    faster in ice compared to in water, and that water heats up fifty times quicker than

    the surrounding ice. The time scales of diffusion can also be used to define the lengthscales of diffusion in the liquid and the solid, ldl and lds, respectively. Similarly, one

    of the heating time scales can be used to choose a temperature scale.

    The Gibbs-Thomson condition (2.8) can be written as

    uI = ,

    where uI is the non-dimensional interface temperature, is the non-dimensional cur-

    vature of the interface, and is a non-dimensional number that characterizes the

    relative effects of the surface energy,

    =

    Ll

    TmT

    .

    Using this parameter, a capillary length scale lcap can be defined that represents the

    smallest radius of curvature that an interface can obtain before surface energy begins

    to flatten it out,

    lcap =

    L

    TmT

    .

    Inserting parameter values shows that lcap

    (2.7

    105/T) mm

    K, which implies

    that for most temperature scales, the capillary length will be small. This most likely

    19

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    32/115

    explains the appearance of a highly corrugated interface when a liquid snowflake first

    becomes unstable.

    The Stefan condition for the motion of the interface (2.11) can be put into non-

    dimensional form by defining a non-dimensional velocity vn = vn/l. Upon dropping

    the prime on the velocity, the Stefan condition becomes

    m

    (1 + ) vn = ul n + krus n, (2.12)

    where is a non-dimensional number that represents the relative energy needed to

    stretch the interface and m denotes the time scale of melting additional ice in front

    of the interface. More specifically, = /(Ll), which is very small provided that

    the length scale of the problem, l, is chosen to be larger than the capillary length.

    The time scale of melting is

    m =Ll2

    T kl,

    which is usually very large because of the high latent heat and the small temperature

    scales.

    The equations can be put into their final non-dimensional form by choosing scales

    for the time, the length, and the temperature. These scales are often chosen to

    balance the magnitudes of various terms in the governing equations, which ensures

    that the key physics occur at leading order in the model. However, the physics that

    drive the interfacial instability in liquid snowflakes are unknown, and when combined

    with the fact that the physical processes occur on different time scales, ensuring that

    the proper terms are in balance is a particularly nontrivial task to accomplish.

    As previously mentioned, the time scale associated with melting is typically large,

    and therefore it will not be balanced with the other time scales. Instead, the balancing

    will occur between the diffusion and the absorption time scales of the liquid. This

    leads to a relationship between the temperature scale and the length scale, T =

    lI1l2/kl. The length scale is determined from the average size of a liquid snowflake,

    which is on the order of millimeters. Hence, l 1 mm, which leads to a temperaturescale of T 2 K. The associated time scale of heating due to radiation absorption isroughly eight seconds, which is reasonable, as this implies that it takes eight seconds

    for the water to heat up by two Kelvin. To simplify the equations, the time scale

    is taken to be the time scale of thermal diffusion in the liquid.

    With a time scale for the problem, the Stefan condition can be written as

    S(1 + ) vn = ul n

    + krus n

    ,

    20

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    33/115

    where S = L/(cplT) is often referred to as the Stefan number [11, 43]. From a

    physical standpoint, this number measures the ratio of energy required to overcome

    the latent heat of fusion to the energy required to raise the temperature of the liquid

    by an amount T. For the ice-water system considered here, S 40, which isconsidered to be large. The interface stretching parameter is small, 107, andtherefore only a tiny percentage of the energy goes into stretching the interface. In

    fact, this number is so small that it will be taken as zero, for simplicity.

    2.3.1 Summary of Non-dimensional Equations

    The full system of non-dimensional equations and their boundary conditions will now

    be summarized. The evolution equations for the temperatures are given by

    ult

    = 2ul + e(r/)2, (2.13a)

    cprust

    = kr2us + re(r/)2. (2.13b)

    The Gibbs-Thomson equation for the interface temperature reads

    uI = (2.14)

    and the Stefan condition for the velocity of the interface is

    Svn = ul n + krus n. (2.15)

    To close the problem, suitable boundary conditions away from the interface must

    be imposed, and initial conditions for the temperature profiles and the interface are

    required. These will be discussed in the individual problems that are considered. A

    summary of the non-dimensional parameters and their typical values can be found in

    Table 2.2.

    2.4 The Quasi-steady Approximation

    While the Stefan number represents a ratio of energies, it was obtained by dividing

    the time scale of melting by the time scale of diffusion in the liquid, and hence it

    also represents a ratio of these time scales. The Stefan number associated with liquid

    snowflake experiments is large, which implies that diffusion of heat occurs much faster

    than melting. Therefore, by the time the ice-water interface can advance, thermal

    diffusion has nearly brought the temperature profiles to their steady state, which

    21

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    34/115

    Table 2.2: Summary of the non-dimensional numbers and their typical values, which arebased on a temperature scale of T 2 K and a length scale of l 1 mm.

    Parameter Definition Value

    cpr cps/cpl 0.5kr ks/kl 3.3r s/l 0.01 rb/l 10

    Ll

    TmT

    1.4 105

    Ll1.0 107

    SL

    cplT

    41

    suggests that the transient dynamics of the temperatures are not significant and can

    be neglected.

    To make this statement rigorous, a new time t that is associated with the time

    scale of melting can be defined as t = t/S, which leads to the evolution equations for

    the temperatures becoming

    1

    S

    ul

    t

    =

    2ul + e

    (r/)2,

    cprS

    ust

    = kr2us + re(r/)2.

    Thus, when the Stefan number is large, the terms involving the time derivatives can

    be neglected and temperature profiles can be assumed to be in their steady state.

    Note, however, that the temperature profiles will still be a function of time because

    of the order one motions of interface which follow from the Stefan condition

    vn = ul n + krus n, (2.16)

    where vn = vn/S a rescaled velocity. Neglecting the transient dynamics of the tem-

    peratures is commonly referred to as the quasi-steady approximation [11, 43], owing

    to the fact that the temperature profiles continuously evolve despite being in their

    steady state at any given moment in time.

    22

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    35/115

    Chapter 3

    Linear Stability Analysis

    Experimental observations show that liquid snowflakes often begin as circular discs

    which then grow outwards in a radially symmetric manner. However, after a certain

    amount of time, the interface becomes unstable and small, sinusoidal perturbations

    with a well defined wavenumber appear. Understanding this phenomenon from a

    mathematical perspective is a problem that is well-suited for a linear stability analysis,

    which investigates the growth of small amplitude perturbations to a known basic

    state. In the context of liquid snowflakes, the basic state corresponds to the outward

    growing liquid disc. The key assumption in linear stability theory is to assume that the

    perturbations to the basic state are sufficiently small that their growth is dominatedby the linear terms in the governing equations. This allows the nonlinear terms to be

    neglected which simplifies the analysis of the perturbations by a considerable amount.

    In general, linear stability theory aims to predict the conditions and the physical

    parameters that cause perturbations to grow in time. This corresponds to the basic

    state becoming unstable, which is an experimentally observable phenomenon. Of

    particular interest in this study is to determine how the wavenumber of the interface

    depends on the intensity of the incident light, as this relationship could be determined

    experimentally.This chapter will investigate the stability of several systems of increasing complex-

    ity. By starting with the simplest models and then building upwards, the key physics

    of the problem can be discovered and understood, which will aid the subsequent

    interpretation of more realistic models.

    3.1 An Illustrative Example

    There are a number of physical mechanisms that could drive the interfacial instabil-

    ity, for example, large temperature gradients in the water and superheating in the

    23

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    36/115

    surrounding ice, both of which result from the absorption of radiation. Moreover, the

    exchange of heat between these two regions is maximized by a highly curved interface,

    which surface energy effects tend to minimize. Thus, there is competitive behaviour

    which will play an important role in selecting the wavenumber of the interface. To

    gain some insight about how these various physical mechanisms interact with each

    other, the dynamics of a planar interface that has prescribed temperature gradients

    in the solid and the liquid regions will first be studied.

    As a basic state, consider a planar interface that is growing into the ice by travel-

    ling along the z-axis with a velocity v, as shown in Figure 3.1. For convenience, the

    system will be written in terms of a single coordinate that moves with the interface,

    = z vt. Negative values of this coordinate will correspond to the liquid region,and positive values will correspond to the solid. Finally, it will be assumed that the

    temperature profiles of the basic state are linear near the interface, so that

    ul() = Gl, < 0,us() = Gs, > 0,

    (3.1)

    where the bars are used to denote variables that are associated with the basic state,

    and Gl and Gs represent the temperature gradients in their respective regions. In

    reality, these temperature gradients will depend on various physical parameters of

    the system. The motivation for using such temperature profiles comes from the fact

    that in a small enough region around the interface, any temperature profile will be

    approximately linear. It should be noted that these profiles are consistent with the

    Gibbs-Thomson interface condition (2.14), in particular, us = ul = 0 at = 0,

    since the curvature of a planar interface is zero. Furthermore, a positive temperature

    gradient in the solid implies that it is, in fact, superheated.

    The interface velocity can be found by applying the Stefan condition (2.16), which,

    in these moving coordinates, is given by

    v = duld

    + krdusd

    , = 0. (3.2)

    Inserting the temperature profiles given in (3.1), the velocity of the planar interface

    is given by

    v = Gl + krGs. (3.3)

    From this expression it can be seen that there are two mechanisms that lead to an

    advance of the interface; superheating in the solid and a liquid with a temperature

    24

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    37/115

    x

    h(t) h(t) + h(x, t)

    v

    Liquid

    Solidz

    Figure 3.1: A planar interface (dashed) growing into a solid with velocity v along thez-axis. Linear stability theory will predict the growth of perturbations to the interface,shown as a solid line.

    profile that decreases near the interface1. Both of these mechanisms make physical

    sense, because they both lead to a transfer of thermal energy to the interface, which

    can then be used to overcome the energy barriers that are associated with melting

    additional ice. A particularly interesting feature of this result is that the effects of a

    nonzero temperature profile in the solid are either amplified or quenched, depending

    on the magnitude of kr, which denotes the ratio of thermal conductivities. For a

    system composed of ice and water, this ratio is approximately three. Therefore, theeffects of superheating in the ice are magnified, which may have an important role in

    the growth rate of a liquid disc.

    3.1.1 Stability of the Basic State

    The idea behind linear stability analysis is to determine the growth rate of pertur-

    bations that are added to the basic state. Such perturbations will be denoted with

    tildes, to make them distinct from the basic state variables. Like the temperature

    profiles of the basic state, the form of the perturbations will also be prescribed,

    ul(x,,t) = ule

    t+a+iax, (3.4a)

    us(x,,t) = use

    ta+iax, (3.4b)

    h(x, t) = het+iax, (3.4c)

    where h denotes the perturbation to the planar interface, a represents the wavenumber

    of the perturbations, and denotes the growth rate of the perturbations. The variable

    1

    A liquid with this temperature profile is really the typical case, since a liquid that has anincreasing temperature profile towards the interface is actually supercooled.

    25

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    38/115

    t denotes a time that is independent of the time t that appears in the expression for

    the moving coordinate . Furthermore, in these moving coordinates, the interface is

    at = h(x, t). Therefore, the interfacial boundary conditions can be written as

    ul(x, h, t) = us(x, h, t

    ) = ,which is the Gibbs-Thomson condition, and

    v +h

    t= ul n + krus n, = h(x, t)

    is the Stefan condition. The normal vector and the curvature are given by

    n =hxex + ez

    1 + h2x, =

    hxx

    (1 + h2x)3/2, (3.5)

    where the subscripts on h denote differentiation with respect to x, and ex and ez

    represent unit vectors along the positive x and z axes, respectively. As it currently

    stands, the above set of boundary conditions pose a nonlinear problem that must be

    solved in order to obtain the growth rate of the perturbations. However, since it is

    the growth of small amplitude perturbations that is of interest, these conditions can

    be expanded about h = 0, which significantly simplifies the problem. Under such

    assumptions, the expressions for the normal vector and the curvature simplify to

    n ez, hxx.By writing the temperature fields as u = u + u, the interfacial boundary conditions

    can be linearized to read

    ul +duld

    h + ul = hxx,

    us +dusd

    h + us = hxx,

    v +h

    t=

    dul

    d+

    d2ul

    d2h +

    ul

    + kr dus

    d+

    d2us

    d2h +

    us

    ,

    which hold at the linearized interface position = 0. Using the fact that the tem-

    peratures of the planar basic state are zero at the interface and that the basic state

    satisfies (3.2), these boundary conditions simplify to become

    duld

    h + ul = hxx, (3.6a)dusd

    h + us = hxx, (3.6b)h

    t = d2ul

    d2 h +ul

    + kr

    d2usd2 h +

    us

    , (3.6c)

    26

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    39/115

    which are still valid at = 0. By inserting the expressions for the basic state (3.1) and

    the perturbations (3.4) into these boundary conditions, a linear system of algebraic

    equations for h, ul, and us emerges,

    Glh + ul = a2h,Gsh + us = a

    2h,

    h = aul akrus.

    These equations can be combined to yield an analytical expression for the growth of

    the perturbations,

    = a

    krGs Gl (1 + kr)a2 . (3.7)This expression clearly shows the stabilizing effects of surface tension, which tend to

    inhibit the growth of large wavenumbers. Moreover, this stabilization mechanism is

    enhanced in a water-ice system because the relative thermal conductivity is greater

    than one. Furthermore, it can be seen that superheating in the solid can drive the

    growth of perturbations, whereas the same temperature gradient in the liquid that

    acts to advance the planar interface, actually stabilizes the system. This compet-

    itive behaviour can be understood by considering what is happening physically ata sinusoidally perturbed planar interface. Such an interface will have liquid fingers

    that extend into the solid, and solid fingers that extend into the liquid, as Figure

    3.2 shows. An unstable growth of the interface requires that both the solid fingers

    and the liquid fingers will increase deeper into their respective regions. For a liquid

    finger to continue its growth, the neighbouring solid must be superheated. Similarly,

    supercooled liquid is required for the solid fingers to grow (see Figure 3.2). If, how-

    ever, the temperature of the liquid decreases towards the interface (corresponding

    to a positive Gl), the liquid is not supercooled and thus growth of the solid fingers

    is inhibited. Therefore the system is, in some sense, stabilized, and this is what is

    reflected in the linear stability analysis. It should be emphasized, however, that the

    interface can still be unstable because the growth of liquid fingers will be sustained

    by local superheating in the solid. The fact that linear stability theory tends to find

    the average growth rate of the fingers is a severe shortcoming, and it remains unclear

    how this will influence the predicted behaviour of more complex models.

    As the water in a liquid snowflake is constantly absorbing incoming radiation, it is

    not expected to be supercooled. The above arguments therefore predict that when the

    interface becomes unstable, the only fingers that will grow are those extending from

    27

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    40/115

    to growRequires superheating

    Solid

    Liquid

    Requires supercoolingto grow

    Figure 3.2: A sinusoidally perturbed interface requires superheating in the solid andsupercooling in the liquid in order to sustain the growth of both solid fingers and liquidfingers.

    Figure 3.3: When the interface of a liquid snowflake goes unstable, liquid fingers grow intothe superheated ice. The initial circular interface can still be seen as a faint line from whichthe fingers grow out of. The image was obtained using the experimental setup described inAppendix A

    the liquid region into the solid. This behaviour is, indeed, observed experimentally,

    as Figure 3.3 illustrates. In particular, experiments consistently show the growthof liquid fingers into the superheated ice, a fact which can be deduced because the

    circular shape of the interface can often still be seen even after it has become unstable.

    More specifically, all of the fingers grow outwards from the circular interface, which

    implies the fingers are composed of liquid. Furthermore, solid fingers would appear

    as protrusions that extend into the centre of the liquid snowflake, and these have yet

    to be observed experimentally.

    The expression for the growth rate of the perturbations (3.7) can written in a

    28

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    41/115

    simplified form as

    = a(1 + kr) G

    a2 ,

    where G denotes a weighted average of the interfacial temperature gradients

    G = ksGs klGlks + kl

    .

    For small wavenumbers, the surface energy effects are small and these perturbations

    will grow provided that the average temperature gradient is positive. However, surface

    energy tends to inhibit large wavenumber perturbations. In particular, perturbations

    with wavenumbers greater than acut = (G/)1/2 are stabilized by surface energyeffects and hence they do not grow. The most unstable perturbation is that which hasa wavenumber that maximizes the growth rate, and this is readily computed as ad =

    acut/

    3. Knowing the wavenumber of the most unstable perturbation is particularly

    useful because it is often this perturbation that is observed experimentally.

    3.2 The Effects of Superheating due to Radiation

    Absorption

    The previous example showed that a local region of superheated ice near the interfacecan lead to the growth of perturbations. In the context of liquid snowflake formation,

    this superheating is caused by the absorption of radiation, and therefore, it can be

    expected that the physical characteristics of this radiation will largely influence the

    stability of the system. To study the basic relationship between the properties of

    the radiation and the evolution of the interface, another model involving a planar

    interface will be used. In this case, however, the temperature of the ice will be solved

    as part of the problem, and for simplicity, the temperature of the water will be taken

    as constant.

    3.2.1 The Basic State

    Proceeding in similar manner to the previous example, the basic state is taken to be

    a planar interface that is moving along the positive z axis with velocity v. Moreover,

    it is assumed that there is a source of radiation that follows this interface, as shown

    in Figure 3.4. The temperature of the liquid is assumed to be constant, and the

    temperature of the solid will satisfy

    cpr ust

    = kr 2

    usz2

    + re(zvt)2/2,

    29

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    42/115

    = 0

    SolidLiquid

    v

    x

    v

    Figure 3.4: Schematic diagram of a solid-liquid system being heated by a light sourcewhich travels with the interface.

    together with the boundary conditions us = 0 at z = vt, which is the Gibbs-Thomson

    condition for a planar interface, and the far field condition us 0 as z . Thisfar field condition is chosen to ensure that any superheating in the solid is solely due

    to the absorption of radiation. The Stefan condition for the interface velocity is given

    by

    Sv = krusz

    , z = vt.

    To reduce the number of parameters in the problem, time can be scaled according to

    t = (kr/cpr)t, which leads to the equation for the temperature becoming

    ust

    =2usz2

    + Ae(zvt)2/2,

    where A = r/kr is a measure of the effective absorption coefficient of ice. The Stefan

    condition becomes

    Sv =usz

    , z = vt,

    where S = S/cpr is a modified Stefan number (which is still large), and v = (kr/cpr)v

    is a rescaled velocity. The primes on these variables will be dropped in subsequent

    computations for notational convenience.It is again useful to write these equations in terms of a variable = z vt that

    travels with the interface. Under this coordinate transformation, the equation for the

    temperature in the solid is given by

    v dusd

    =d2usd2

    + Ae(/)2

    , (3.8)

    which has boundary conditions us = 0 at = 0 and us 0 as . The Stefancondition reads

    Sv =us , = 0. (3.9)

    30

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    43/115

    The solution to this problem can be found analytically, and it can be written most

    simply in terms of integrals involving the error function

    us() =

    2A e(v/2)

    2

    ev erf

    v

    2

    d

    ev

    0

    ev erf

    v

    2

    d

    , (3.10)

    where is a dummy variable and the interface velocity is given by

    v =

    2

    A

    S.

    To compute this velocity, the integral identity

    ev erf

    v2

    d = v1ev erf

    v2

    + v1e(v/2)2

    1 erf

    ,

    has been used, which is readily obtained via integration by parts.

    The dependence of the temperature on the modified Stefan number is shown in

    Figure 3.5. While the decay rate of these profiles varies significantly, their behaviour

    near the interface is almost identical, which suggests that the Stefan number will

    not play a significant role in the dynamics that are predicted by the linear stability

    analysis. The relationship between the characteristics of the incident radiation and

    the interface velocity is best seen by writing the velocity in dimensional form,

    v =

    2

    sI1rbL

    . (3.11)

    This expression shows, rather unsurprisingly, that the interface velocity is determined

    by the rate at which the radiation can supply the ice with enough energy to overcome

    the latent heat. Since ice is a poor absorber of radiation, the velocity predicted by this

    expression is orders of magnitude smaller than what is observed experimentally. This

    may suggest that strong temperature gradients in the water play an important role in

    driving the initial growth of a circular liquid snowflake. However, it is important to

    keep in mind that (3.11) was the result of studying the evolution of a planar interface,

    and this could be a poor representation of the circular interface of a liquid snowflake.

    3.2.2 Stability of the Basic State

    To investigate the growth of perturbations, the multidimensional analogues of the

    basic state equations must be considered. In particular, the temperature of the solid

    will satisfy

    ust

    = 2usx2

    + 2usz2

    + Ae(zvt)2/2 ,

    31

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    44/115

    0 1 2 3 4 5 6 7 8 9 100

    0.05

    0.1

    0.15

    0.2

    0.25

    0.3

    0.35

    /

    Tus

    [K]

    S= 1

    S= 10

    S= 40

    Figure 3.5: Temperature profiles of a solid being superheated by incident radiation asviewed from a frame that moves with the planar interface. Parameter values for the non-

    dimensional numbers can be found in Table 2.2.

    32

  • 7/28/2019 Liquid Snowflake Formation in Superheated Ice

    45/115

    with boundary conditions us = at z = h(x, t), where h denotes the location ofthe interface, and us

    0 as z

    . Moreover, the temperature is assumed to be

    periodic in the x direction. The Stefan condition is given by

    Sh

    t= us n, z = h(x, t),

    where, as mentioned above, S is a modified Stefan number. Expressions for the

    curvature and the normal vector n are given in (3.5).

    Before solving these equations, a change of variable is made so that the equations

    can be written in terms of the usual coordinate = z vt that follows the planarinterface. Furthermore, a new time t = t/S is introduced that is independent from

    the time in the moving coordinate . The derivatives of the old variables t and z are

    related to the derivatives of the new variables and t via

    t=

    1

    S

    t v

    ,

    z=

    .

    Therefore, the evolution equation for the non-dimensional temperature in these new

    coordinates is given by

    1

    S

    ust

    v us

    =2usx2

    +2us2

    + Ae(/)2

    ,

    and the interface is now at = h(x, t) Svt. The Gibbs-Thompson condition forthe interfacial temperature remains the same, and