enthalpy method how-to

Upload: john-brimlow

Post on 03-Jun-2018

215 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/12/2019 Enthalpy Method How-To

    1/14

  • 8/12/2019 Enthalpy Method How-To

    2/14

    4.2 OVERVIEW OF NUMERICAL METHODS 211

    All such approaches work well, more or less, for simple Stefan problems (that

    would arise e.g., in laboratory settings) in which we know what to expect, namely

    a single sharp front separating the two phases. It is not difficult to realize however,

    that such problems are not the rule in practice, particularly when time dependence

    of heat input/output or thermal cycling occur. For example, in Latent Heat Ther-

    mal Energy Storage, one must deal with cases of extreme thermal cycling, multiplefronts, disappearing phases and non-predictable behavior (1.3, 5.3). Internal

    heating is another source of difficulties, first documented by [ATTHEY], in which

    extended mushy zones may appear instead of sharp fronts. Constitutional super-

    cooling of binary alloys results in similar effects which may not be ignored ( see

    [ALEXIADES-WILSON-SOLOMON, 1985] ). Simultaneous mass transfer by

    diffusion and/or convection complicate the phase change process to the point that

    we cannot guess a priori the qualitative picture in enough detail to even be able to

    formulate the problem in the classical fashion of a Stefan type problem with sharp

    front, etc. If so many complications can arise in 1-dimensional processes, what

    about 2- and 3-dimensional processes? Despite the difficulties, successful methods

    for 2-D hydrodynamic instability problems have been developed by Glimm and

    coworkers, [GLIMM]. Surveys of front tracking methods appear in [MEYER,

    1978], [CRANK, 1981, 1984], [ALBRECHT-COLLATZ-HOFFMANN], etc.

    Such reasons make front-tracking schemes unviable as general simulation tools

    for modeling realistic phase-change processees. The only viable general approach

    is the so-called enthalpy method, precisely because it bypasses the explicit track-

    ing of the interface. In this approach the jump condition (Stefan condition) is not

    forcedon the solution, but it is obeyed automatically by it as a natural boundary

    condition (in the sense of the Calculus of Variations). Its theoretical basis is a

    formulation of the Stefan problem different than the classical one, the so-called

    weak or enthalpy formulation, described in 4.4. It is similar to the weak formu-

    lations commonly used in gas dynamics for shocks (see [HYMAN] for a brief

    overview). Another fixed-domain method (as opposed to front-tracking), based on

    variational inequalities [DUVAUT], [ODEN-KIKUCHI] reformulation of the

    Stefan problem and finite elements, lacks the direct physical interpretation of the

    enthalpy method and has not lived up to its initial promise for Stefan-type prob-

    lems.

    It can be safely concluded today that the enthalpy method, to which we turn in

    the next section, discretized by (integrated) finite differences, is the most versatile,

    convenient, adaptable, and easily programmable numerical method available for

    phase change problems in 1, 2 or 3 space dimensions.

    We hasten to add however that it does not solve all the problems. Excluded

    are problems which we do not know how to formulate weakly due to their special

    interface conditions. Such is the case with supercooling problems, where the

    instability of the interface must be studied. A very successful computational

    approach for such problems is another fixed-domain type formulation, the so-

    called phase-field approach, under intense development lately, [CAGINALP,

    1989, 1991], [KOBAYASHI].

  • 8/12/2019 Enthalpy Method How-To

    3/14

    212 CHAPTER 4

    4.3. THE ENTHALPY METHOD

    IN ONE SPACE DIMENSION

    4.3.A Introduction

    The, so called, enthalpy or weak solution approach is based on the fact that the

    energy conservation law, expressed in terms of energy (enthalpy) and temperature,

    together with the equation of state contain all the physical information needed to

    determine the evolution of the phases. It turns out that, for the purpose of

    obtaining numerical schemes, the most appropriate and convenient way to state

    energy conservation is the primitive integral heat balance over arbitrary volumes

    and time-intervals, from which all other formulations can be obtained, namely

    t+t

    t

    t

    V E dV

    dt =t+t

    t

    V q

    . n

    dS dt (1)

    whereE= e is the energy density (per unit volume), and q . n

    is the heat flux

    into the volumeV across its boundaryV,n being the outgoing unit normal toV.The distinct advantage of this primitive form is that it is valid irrespectively of

    phase, and even ifE andq

    experience jumps, so it is actually more general than

    the localized differential form

    (2)Et+ divq

    = 0 .

    The two forms are equivalent for smoothE,q

    , thanks to the Divergence Theorem

    ( 1.2 ). In the presence of a phase-change, the partial differential equation (2) can

    only be interpreted in the classical pointwise sense inside each phase separately,

    and then conservation across the interface must be imposed explicitly as an

    additional interface (Stefan) condition, making front-tracking necessary.

    Alternatively, the PDE (2) may be interpreted in a generalized (weak) sense

    globally, as described in detail in 4.4. It turns out that the numerical solutionsobtained via the enthalpy method approximate this weak solution, as we shall

    show in 4.5.

    In this section, we describe the enthalpy method for Stefan problems in one

    space dimension, and its numerical implementation via time-explicit or time-

    implicit schemes.

    4.3.B The enthalpy method

    The idea of the enthalpy approach is very simple, direct, and physical. We

    partition the volume occupied by the phase-change material into a finite number of

    control volumesVj and apply energy conservation, (1), to each control volume to

    obtain a discrete heat balance. Note that this is the same discrete heat balance asfor plain heat conduction ( 4.1.B ), and we use it to update the enthalpy, Ej , of

  • 8/12/2019 Enthalpy Method How-To

    4/14

    4.3 THE ENTHALPY METHOD IN ONE SPACE DIMENSION 213

    each control volume. From the equation of state we know that Ej 0 ==>Vj issolid,Ej L ==>Vj is liquid, and 0 Vj is partially liquid andpartially solid, so we call it "mushy". A mushy cell contains an interface and the

    fraction of the cell occupied by liquid is naturally given by the value of the

    liquid fraction : j = EjL .

    Note that in this scheme the phases are determined by the enthalpy alone, with no

    mentioning of interface location(s). It is a volume-tracking scheme, as opposed

    to front-tracking. Since the front location may be recovereda posteriori from

    the values of the enthalpy, it may be characterized as a front-capturing scheme,

    similar in spirit to shock-capturing schemes of gas dynamics (see [HYMAN]

    for an overview of various types of schemes).

    Let us see how the method works in detail, by considering the heat conduction

    problem of4.1, except now we assume that our slab 0x l is occupied by amaterial that changes phase at a melt temperatureTm . We assume that initially the

    material is solid with

    (3)T(x, 0) = Tinit(x) Tm , 0x l ,the facex = 0 is heated convectively byT(t)Tm :

    (4)q(0, t) = h [ T(t)T(0, t) ] , t> 0 ,

    and the facex =l is insulated :

    (5)q(l, t) = 0 , t> 0 .

    The energy conservation law in its integrated form (1) applied to the present one-

    dimensional control volumes Vj= [xj12 ,xj+12 ] A (see 4.1.B ) becomestn+1

    tn

    t

    A

    xj+12

    xj12

    E(x, t)dx

    dt = tn+1

    tn

    Axj+12

    xj12

    qx(x, t)dx dt. (6)

    We seek numerical approximations to the temperature, energy and flux

    obeying (3)-(6) withq= k Tx of course.This problem is formally identical to the heat conduction problem (1b,c),(8a)

    of4.1. The two differ in that now the enthalpyE is the sum of sensible and latent

    heat in the liquid, so that, instead of (6) of4.1.B, we have

    E(x, t) =

    T(x,t)

    Tm

    cS()d , T(x, t) Tm (liquid)(7)

    The phases are described by

  • 8/12/2019 Enthalpy Method How-To

    5/14

    214 CHAPTER 4

    (8a)E(x, t) 0 => solid at (x, t)

    (8b)0 interface at (x, t)

    (8c)E(x, t) L => liquid at (x, t) .

    Thus, only the relation between E andT is now different than in 4.1, while the

    discretization of (6) is still (12) of 4.1.B. The flexibility and generality of this

    approach will be further illustrated in 4.3.H where even a wall layer will be

    incorporated into the global scheme by adjusting the energy.

    Consider the case in which

    (9)cS , cL = constants .Then (7) becomes

    E =

    cS[ T Tm ] , TTm

    (10)

    or, solving forT,

    T =

    Tm+ E

    cS

    , E 0 ( solid )

    Tm , 0

  • 8/12/2019 Enthalpy Method How-To

    6/14

    4.3 THE ENTHALPY METHOD IN ONE SPACE DIMENSION 215

    The updating algorithm from any time tn to the nexttn+ tn proceeds as fol-lows: Knowing the enthalpy, temperature and phase (see below) of each control

    volume, we compute the resistances and fluxes, which are then used to update the

    enthalpies, which in turn yield new temperatures and phase states.

    The most convenient phase-indicator is the liquid fraction of a control volume

    Vj , defined as

    nj =

    0 , if Enj 0 ( solid )Enj

    L, if 0 < Enj < L ( mushy )

    1 , if LEnj (liquid)

    . (13)

    If 0 < nj < 1 the control volume is said to be mushy with liquid volume nj xj

    and solid volume (1 nj )xj (per unit cross sectional area).The definitions of resistances and fluxes between control volumes are identical

    to their definitions in 4.1, with the resistance atxj12 expressed as

    Rj12 = xj1

    2kj1

    +xj

    2kj

    .

    The effective conductivitykj of a mushy control volume depends on the structure

    of the phase-change front, and it is not always clear how to choose it, especially in

    2 or 3 dimensional situations. Some alternative choices are:

    Sharp front(s): A control volume containing a sharp front consists of layers of

    solid and liquid in a "serial" arrangement, for which the effective resistivity is

    the sum of the resistivities of the layers. With the layer thicknesses determined

    from the solid and liquid fractions, we have

    (14a)1

    knj=

    nj

    kL(Tm)+

    1njkS(Tm)

    , j= 1, 2, . . . ,M.

    Columnar front: A front consisting of columns of solid and liquid constitutes a"parallel" arrangement, so the effective conductivity is the sum of the conduc-

    tivities of the phases :

    (14b)knj = njkL(Tm)+ (1

    nj )kS(Tm) .

    Amorphous mixture of solid and liquid: The inter-phase region may be a random

    mixture of solid and liquid. In this case one may use the following formula

    which interpolates the previous two cases [CHEMICAL ENGINEERING

    GUIDE, p.242]:

    knj = kS(Tm)1+ 2/3( 1)

    1+ (2/3 )( 1), = nj , =

    kL(Tm)

    kS(Tm). (14c)

    In most situations we do not know which of the above cases is relevant. In 2

    or 3 space dimensions even a sharp front will generally not be moving in the

  • 8/12/2019 Enthalpy Method How-To

    7/14

    216 CHAPTER 4

    direction of one of the axes, so the choice is not clear. A simple expedient is to

    take the average of the solid and liquid conductivities,

    (14d)knj = 12 (kS+ kL)

    However, the best alternative, when applicable, is to employ the "Kirchoff trans-

    formation" (see (7)4.4.D), to replace the temperatureT by the "Kirchoff tempera-ture"u. This can be used when the conductivity is a function of temperature only.

    In particular, for constantkS,kL the "Kirchoff temperature" is

    u =

    kS[ T Tm ]0

    kL [ T Tm ]

    if T < Tm

    if T = Tmif T > Tm

    . (15)

    Thenq k Tx= ux , so the discrete flux is simplyqj12= ( uj1uj ) /x . Tocompare this with (12e), resubstitute u in terms ofT: uj= kj [ Tj Tm ], andwrite it as

    qj12 = Tj1Tm

    Rj1

    +TmTj

    Rj

    , Rj : = x/kj . (16)

    Thus the flux neatly splits to a sum of two terms, one for each of the two adjacent

    nodes. Note that if one of the nodes, say node j, is mushy thenTj= Tm and thenode is not contributing to conduction. We see that in this prescription each node

    has its own resistivity Rj= x/kj , which may be found conveniently from

    (17)Rj= x { j/kL+ (1j) /kS} ,

    the value for mushy nodes being irrelevant since they are not contributing to the

    flux. No averaging of values is used, so this prescription results in the highest

    effective conductivity. It is the best choice for the enthalpy scheme since it is con-

    sistent with the mushy nodes being treated as isothermal.

    Note that such issues arise only whenkL(Tm) andkS(Tm) are substantially dif-

    ferent, in which case the sensitivity of the solution to the choice of effective con-ductivity should be examined by comparing the results from the various choices.

    (14a) yields the lowest effective conductivity and (16)-(17) yields the highest, so

    these two pretty much bracket the system behavior.

    We emphasize again that the interface location is not involved in the computa-

    tion at all, this being an essential advantage of the enthalpy method. If the prob-

    lem being modeled admits a sharp interface, then the enthalpy scheme ought to

    produce asingle mushy node at each time step. If at timetn the mushy node is the

    m-th node, then a good approximation to the interface locationX(tn) is given by

    (18)Xn : = xm12+ nmxm .

    REMARK 1. Missing the phase-change

    The latent heat effect is only felt in mushy nodes, so each control volume

    should pass through the mushy state before changing phase. The algorithm

  • 8/12/2019 Enthalpy Method How-To

    8/14

    4.3 THE ENTHALPY METHOD IN ONE SPACE DIMENSION 217

    described here is "robust" in this respect, so skipping this transition indicates a

    bug in the code, or too large a time-step. Some other implementations may not

    be so robust and one must be careful not to miss the transition. Especially

    prone are algorithms that track the temperature and account for the latent heat

    via a source term.

    REMARK 2. Temperature dependent heat capacities

    Ifc S=cS(T),cL=cL(T), then the equation of state is the nonlinear rela-tion (7), and findingT fromE is not quite as simple as in the constant cS,cL

    case. If E 0 we need to find T from the equation E=T

    Tm

    cS()d, forwhich a Newton-Raphson method may be employed. Alternatively, we may

    rewrite the equation asdE

    dT= cS(T), or

    dT

    dE=

    1

    cS(T), which is an ODE with

    initial condition T= Tm for E= 0. Similarly, if E L, then we may

    solve the equationE=T

    Tm cL()d+ L via a Newton-Raphson method, or

    solve the ODEdT

    dE=

    1

    cL(T)with initial condition T= Tm for E= L.

    Any ODE solver can be used for this purpose, e.g. forward Euler, backward

    Euler, Runge-Kutta, etc.

    Actually, the temperature dependence of heat capacities is commonly

    expressed in the form

    ci(T) = Ai+ B iT+Ci

    T2, i= S,L , (19)

    withT in degrees Kelvin and Ai, Bi, Ci given constants. Then the integrals

    expressing the sensible heat can be computed analytically, and the resulting

    algebraic equations may be solved very effectively via a Newton-Raphson

    method. Note that this needs to be done for each node at each time step,adding considerably to the expense of the computation. A reasonable starting

    value is the temperature at the previous time step,Tnj .

    4.3.C A time-explicit scheme

    Choosing = 0 in (12), the fluxes are evaluated at the old time tn and weassume that up to time tn+1 the process is driven by these fluxes. The explicit

    scheme proceeds as follows.

    Initially, the phase and temperature of each control volume are known with

    T0j = Tinit(xj) , j= 1, 2, . . . ,M, (20)

    which in turn determine the enthalpies E

    0

    j , j= 1, 2, . . . ,M, via (10). Assume thatwe have found enthalpies, temperatures and phase-states ( j ) through then-th

  • 8/12/2019 Enthalpy Method How-To

    9/14

    218 CHAPTER 4

    time step. From (13) we find the liquid fractions,nj , hence the phase of each node

    and from (12f) the (mean) temperatures. If only one node is mushy the interface

    location at timetn is given by (18).

    Now we compute the conductivities from (14), the resistances and fluxes from

    (12b,c,e), with= 0, and thenEn+1j is found from (12d), j= 1, 2, . . . ,M. Note thatfor the boundary control volumes ( j= 2,M 1 ) we may use the analog to theimplicit relation (43) of 4.1, in order to guarantee the Maximum Principle and

    still have the CFL condition guarantee no growth of errors. Thus, the updating of

    enthalpies to timetn+1 is complete.

    The stability condition is the same as in the pure conduction case, namely

    tn 1

    2

    (min x)2

    (maxn), (21)

    where min x=j=1,2,...,M

    min xj and

    maxn = max

    kL(Tnj )

    cj,

    kS(Tnj )

    cj, j= 1, 2, . . . ,M

    .

    It is good practice to take a number slightly smaller than 12 in order to avoid stabil-

    ity problems arising from roundoff. The value of maxn can be computed at each

    time step tn and then we can use tn=1

    2

    (min x)2

    maxnas the next time step. As

    this quantity may become impractically small, it is good programming practice to

    halt the computation iftn becomes smaller than a prescribed minimum t, andcarefully examine what caused it to become so small.

    The great advantage of the time-explicit scheme lies in its simplicity and the

    ease with which it can be programmed. In situations where the time-step must be

    small for physical reasons (to capture rapidly moving fronts or resolve rapid

    changes in data, for example), the stability requirement may not impose undue

    restrictions, and the explicit scheme may turn out to be as efficient as implicit

    schemes. The extreme case of laser annealing with picosecond or nanosecondpulses is such a situation [ALEXIADES et al, 1985a].

    4.3.D Performance of the explicit scheme on a one-phase problem

    To see how the scheme of4.3.C performs, we test it on the simplest problem

    with known exact solution, namely the one-phase Stefan Problem with constant

    imposed temperature at x= 0. The Neumann (similarity) solution in dimension-less variables appears in (14)-(16) of2.1.

    To retain direct physical meaning, we implement the enthalpy scheme on the

    original formulation (1)-(4) of2.1, which can be made identical to the dimension-

    less formulation ((19)-(23), 2.1) by choosing

    Tm= 0 . = cL= kL= 1 , TL= 1 , L= 1St

    . (22)

  • 8/12/2019 Enthalpy Method How-To

    10/14

    4.3 THE ENTHALPY METHOD IN ONE SPACE DIMENSION 219

    We simulate melting in the slab 0x 1 for two extreme values of the Stefannumber: St= . 1 andSt= 5; the corresponding transcendental roots are found tobe= . 22 and= 1. 06.

    To exhibit the convergence of the algorithm, we discretize the slab 0x 1withM= 10, 20 and 40 uniform subintervals. Since=kL/cL= 1 (from (22)),

    the corresponding time steps will be (see (21))t 12

    ( 110

    )2, 12

    ( 120

    )2 and 12

    ( 140

    )2,

    showing the severe limitation on the time step imposed by the stability criterion.

    Table 4.3.1 shows temperatures at several locations x, at time t= 3 for theproblem withSt= 0. 1 and at timet= . 12 whenSt= 5. At these times the meltfronts have not reached x= . 8 yet, so there has been no backface influence. Thesecond column shows the exact (Neumann) temperatures found as in 2.1. The

    numerically computed temperatures withM= 10, 20 and 40 nodes are listed in theother columns (linearly interpolated from nodal values for M= 20 and M= 40).Observe the progressive convergence to the exact solution as M increases. It is

    also interesting to compare the code execution run times for the three mesh sizes:

    on an IBM-PC/XT, they were 12.5, 41.5 and 260 seconds forSt= 0. 1 and 7, 9 and24 seconds forSt= 5, respectively forM= 10, 20 and 40 illustrating the dramaticslowdown the finer mesh causes.

    In Table 4.3.2 we compare the exact and numerical interface locations at a few

    sample times from the same runs as above.

    In Figures 4.3.1 and 4.3.4, the computed temperature history (melting curve) at

    a fixed location is compared with the exact (Neumann) solution whenStSt= 0. 1 andStSt= 5 respectively. The staircase shape is characteristic of enthalpy methodsand it is much more pronounced forM= 10 nodes (Fig. 4.3.1(a), 4.3.4(a)) than for

    Table 4.3.1: Exact and Computed Temperature Profiles

    For St= 0. 1 at timet= 3. 0 For St= 5 at timet= . 12

    x Texact M= 10 M= 20 M= 40 x Texact M= 10 M= 20 M= 40

    0 1.0000 1.0000 1.0000 1.0000 0 1.000 1.000 1.000 1.000

    .1 .8667 .8667 .8677 .8672 .1 .8132 .8144 .8135 .8133

    .2 .7336 .7333 .7359 .7346 .2 .6341 .6360 .6346 .6342

    .3 .6010 .6000 .6051 .6026 .3 .4692 .4711 .4698 .4693

    .4 .4691 .4666 .4755 .4712 .4 .3236 .3254 .3243 .3238

    .5 .3380 .3333 .3473 .3405 .5 .2003 .2035 .2015 .2004

    .6 .2080 .2000 .2203 .2105 .6 .1001 .1076 .1017 .1002

    .7 .0793 .0667 .0942 .0809 .7 .0220 .0331 .0193 .0241

    .8 0.0 0.0 0.0 0.0 .8 0.0 0.0 0.0 0.0

    .9 0.0 0.0 0.0 0.0 .9 0.0 0.0 0.0 0.0

    1.0 0.0 0.0 0.0 0.0 1.0 0.0 0.0 0.0 0.0

  • 8/12/2019 Enthalpy Method How-To

    11/14

    220 CHAPTER 4

    Table 4.3.2: Exact and Computed Interface Locations

    For St= 0. 1 For St= 5

    time Xexact M= 10 M= 20 M= 40 time Xexact M= 10 M= 20 M= 40

    0 0. 0. 0. 0. 0 0. 0. 0. 0.

    .5 .3111 .3089 .3101 .3108 .02 .2998 .2995 .3008 .3004

    .0 .4400 .4375 .4401 .4399 .04 .4240 .4237 .4243 .4251

    1.5 .5389 .5369 .5390 .5388 .06 .5193 .5227 .5197 .5206

    2.0 .6223 .6207 .6218 .6225 .08 .5996 .6082 .6023 .6006

    .10 .6704 .6958 .6724 .6722

    Neumann solution

    numerical with M = 10 nodes

    Figure 4.3.1(a). Temperature history atx = . 3, withM= 10 nodes,compared with the exact solution, StSt= 0. 1 .

    M= 40 nodes (Fig. 4.3.1(b), 4.3.4(b)). This is due to the fact that while the inter-face lies anywhere inside a particular mesh interval, the temperature of that inter-

    val is held atTm, so the temperature in the rest of the slab relaxes to a steady state

    corresponding to a fixed isotherm through that node. When the interface moves tothe next mesh interval, the temperature adjusts rapidly and then relaxes to a new

    steady state. It follows that the duration of each step is strictly a function of the

    time the interface remains in each mesh interval, and therefore, the finer the mesh

    the shorter the steps. Indeed, usingM= 80 nodes the computed and exact solu-tions would be indistinguishable graphically. Figures 4.3.2 and 4.3.5 show that the

    interface location computed with onlyM= 10 andM= 20 nodes, forStSt= 0. 1 andStSt= 5 respectively, agrees well with the exact interface. Finally, temperature pro-files with M= 10 and M= 40 nodes are shown in Figures 4.3.3(a),(b) (at timet= 2.,StSt= 0. 1) and Figures 4.3.6(a),(b) (at timet= . 1,StSt= 5).

    These figures bare out the fact that while numerical methods for phase-change

    problems can easily capture interface locations and even temperature profiles, the

    errors show up vividly in temperature history plots.

  • 8/12/2019 Enthalpy Method How-To

    12/14

    4.3 THE ENTHALPY METHOD IN ONE SPACE DIMENSION 221

    numerical with M = 40 nodes

    Neumann solution

    Figure 4.3.1(b). Temperature history atx = . 3, withM= 40 nodescompared with the exact solution, StSt= 0. 1 .

    Neumann solution

    numerical with M = 10 nodes

    Figure 4.3.2. Melt front location withM= 10 nodes,compared with the exact solution, StSt= 0. 1 .

    numerical with M = 10 nodes

    Neumann solution

    Figure 4.3.3(a). Temperature profile at timet= 2., withM= 10 nodes,compared with the exact solution, StSt= 0. 1 .

  • 8/12/2019 Enthalpy Method How-To

    13/14

    222 CHAPTER 4

    Neumann solution

    numerical with M = 40 nodes

    Figure 4.3.3(b). Temperature profile at timet= 2., withM= 40 nodes,compared with the exact solution, StSt= 0. 1 .

    numerical with M = 10 nodes

    Neumann solution

    Figure 4.3.4(a). Temperature history atx = . 3, withM= 10 nodes,compared with the exact solution, StSt= 5 .

    numerical with M = 40 nodes

    Neumann solution

    Figure 4.3.4(b). Temperature history atx = . 3, withM= 40 nodes,compared with the exact solution, StSt= 5 .

  • 8/12/2019 Enthalpy Method How-To

    14/14

    4.3 THE ENTHALPY METHOD IN ONE SPACE DIMENSION 223

    Neumann solution

    numerical with M = 20 nodes

    Figure 4.3.5. Melt front location withM= 20 nodes,compared with the exact solution, StSt= 5 .

    Neumann solution

    numerical with M = 10 nodes

    Figure 4.3.6(a). Temperature profile at timet= . 1, withM= 10 nodes,compared with the exact solution, StSt= 5 .

    Neumann solution

    numerical with M = 40 nodes

    Figure 4.3.6(b). Temperature profile at timet= . 1, withM= 40 nodes,compared with the exact solution, StSt= 5 .