You are on page 1of 39

`

PHASE FIELD MODELING


A REPORT SUBMITTED ON THE COMPLETION OF SUMMER INTERNSHIP AT BHABHA ATOMIC RESEARCH CENTRE .

UNDER THE GUIDANCE OF: DR. ASHOK ARYA MATERIALS SCIENCE DIVISION BHABHA ATOMIC RESEARCH CENTRE MUMBAI

SUBMITTD BY: NITIN SINGH 09MT3010 IIT KHARAGPUR

[Pick the date]

INTRODUCTION: The phase-field method has recently


emerged as a powerful computational approach to modeling and predicting mesoscale morphological and microstructure evolution in materials. It describes a microstructure using a set of conserved and non conserved field variables (or order parameters) that are continuous across the interfacial regions. The temporal and spatial evolution of the field variables is governed by the CahnHilliard nonlinear diffusion equation and the Allen-Cahn relaxation equation. With the fundamental thermodynamic and kinetic information as the input, the phase-field method is able to predict the evolution of arbitrary morphologies and complex microstructures without explicitly tracking the positions of interfaces.

1 - WHY PHASE FIELD MODELLING IS IMPORTANT IN MATERIAL SCIENCE?


The properties of most engineered materials have a connection with their underlying microstructure. For example, the crystal structure and impurity content of silicon will determine its band structure and its subsequent quality of performance in modern electronics. Most large-scale civil engineering applications demand high-strength steels containing a mix of refined crystal grains and a dispersion of hard and soft phases throughout their microstructure. For aerospace and automotive applications, where weight to strength ratios are a paramount issue, lighter alloys are strengthened by precipitating second-phase particles within the original grain structure. The combination of grain boundaries, precipitated particles, and the combination of soft and hard regions allow metals to be very hard and still have room for ductile deformation. It is notable that the lengthening of span bridges in the world can be directly linked to the development of pearlitic steels. In general, the technological advance of societies has often been linked to their ability to exploit and engineer new materials and their properties. In most of the above examples, as well as a plethora of untold others, microstructures are developed during the process of solidification, solid-state precipitation, and thermomechanical processing. All these processes are governed by the fundamental
1

[Pick the date]

physics of free boundary dynamics and nonequilibrium phase transformation kinetics. For example, in solidification and recrystallization both of which serve as a paradigm of a firstorder transformation nucleation of crystal grains is followed by a competitive growth of these grains under the drive to reduce the overall free energy bulk and surface of the system, limited, however, in their kinetics by the diffusion of heat and mass. Thermodynamic driving forces can vary. For example, solidification is driven by bulk free energy minimization, surface energy and anisotropy. On the other hand, strain-induced transformation must also incorporate elastic effects. These can have profound effects on the morphologies and distribution of, for example, second-phase precipitates during heat treatment of an alloy. The above raised arguments are quite sufficient to support the cause of understanding and simulating the formation of microstructure. Phase Field Modeling has emerged as a powerful tool to simulate the evolution of microstructure which is much easier than its predecessor (sharp interface approach) for such work in terms of mathematics and its application. The following section will make it much clearer. 1.2 - SHARP INTERFACE APPROACH: In conventional modeling technique of for phase transformations and microstructural evolution i.e the sharp interface approach, the interfaces between different domains are considered to be infinitely sharp, and a multi-domain structure is described by the position of the interfacial boundaries. The kinetics of microstructure formation is then modeled by a set of partial differential equations that describe the release and diffusion of heat, the transport of impurities, and the complex boundary conditions that govern the thermodynamics at the interface for each domain . As a concrete example, in the solidification of a pure material the advance of the solidification front is limited by the diffusion of latent heat away from the solidliquid interface, and the ability of the interface to maintain two specific boundary conditions; flux of heat toward one side of the interface is balanced
2

[Pick the date]

by an equivalent flux away from the other side, and the temperature at the interface undergoes a curvature correction known as the GibbsThomson condition. These conditions are mathematically expressed in the following sharp interface model, commonly known as the Stefan problem: T/t = .(k.T/.cp) = .( T) Lf Vn = ksT. n|sint - kLT. n|Lint Tint = Tm-(TM/Lf) (Vn/)

(1)

where T =T(x, t) denotes temperature, k thermal conductivity (which assumes values ks and kL in the solid and liquid, respectively), the density of the solid and liquid, cp the specific heat at constant pressure, the thermal diffusion coefficient, Lf the latent heat of fusion for solidification, the solidliquid surface energy, TM the melting temperature, the local solid liquid interface curvature, Vn the local normal velocity of the interface, and the local atomic interface mobility. Finally, the subscript int refers to interface and the superscripts S and L refer to evaluation at the interface on the solid and liquid side, respectively. Like solidification, there are other diffusion-limited phase transformations whose interface properties can, on large enough length scales, be described by specific sharp interface kinetics. Most of them can be described by sharp interface equations analogous to those in Equation 1. Such models often referred to as sharp interface models operate on scales much larger than the solidliquid interface width, itself of atomic dimensions. As a result, they incorporate all information from the atomic scale through effective constants such as the capillary length, which depend on surface energy, the kinetic attachment coefficient, and thermal impurity diffusion coefficient. 1.3 - SHARP INTERFACE MODELS VS DIFFUSE INTERFACE MODELS (MORE GENERALLY REFERRED TO AS PHASE FIELD MODELS): A limitation encountered in modeling free boundary problems is that the appropriate sharp interface model is often not known for
3

[Pick the date]

many classes of phenomena. For example, the sharp interface model for phase separation or particle coarsening, while easy to formulate nominally , is unknown for the case when mobile dislocations and their effect of domain coarsening are included. A similar situation is encountered in the description of rapid solidification when solute trapping and drag are relevant. There are several sharp interface descriptions of this phenomenon, each differing in the way they treat the phenomenological drag parameters and trapping coefficients and lateral diffusion along the interface. Another drawback associated with sharp interface models is that their numerical simulation also turns out to be extremely difficult. The most challenging aspect is the complex interactions between topologically complex interfaces that undergo merging and pinch-off during the course of a phase transformation. Such situations are often addressed by applying somewhat arbitrary criteria for describing when interface merging or pinch-off occurs and by manually adjusting the interface topology. It is worth noting that numerical codes for sharp interface models are very lengthy and complex, particularly in 3D. Along with these two drawbacks of sharp interface models, one would not be able to completely appreciate the diffuse interface approach if the most important advantage of the later over former is not mentioned here. Main advantage gained by using phase-field method to model phase transitions, compared to the sharp-interface method, is that the explicit tracking of the moving surface, the liquid and solid interface, is completely avoided. Instead, the phase of each point in the simulated volume is computed at each time step. In classical formulation the basic equations have to be written for each medium and the interface boundary conditions must be explicitly tracked. In diffuse-interface theory the basic equations, with supplementary phase field terms, are deduced from a free energy functional for the whole system and interface conditions do not occur. In fact, they are replaced by a partial differential equation for the phase field.

[Pick the date]

2 DETAILS OF PHASE FIELD MODELLING:


As already mentioned, the phase field method has proved to be extremely powerful in the visualization of the development of microstructure without having to track the evolution of individual interfaces, as is the case with sharp interface models. The method, within the framework of irreversible thermodynamics, also allows many physical phenomena to be treated simultaneously. The primary purpose of this section is to present the general concepts underlying the phase field modeling. Imagine the growth of a precipitate which is isolated from the matrix by an interface. There are three distinct entities to consider: the precipitate, matrix and interface. The interface can be described as an evolving surface whose motion is controlled according to the boundary conditions consistent with the mechanism of transformation. The interface in this mathematical description is simply a two dimensional surface; it is said to be a sharp interface which is associated with an interfacial energy per unit area. In the phase field method, the state of the entire microstructure is represented continuously by a single variable known as the order parameter . For example =1, =0 and 0<<1 represent the precipitate, matrix and interface respectively. The latter is therefore located by the region over which changes from its precipitate value to its matrix value, as shown in figure below.

[Pick the date]

The range over which it changes is the width of the interface. The set of values of the order parameter over the whole volume is the phase field. The total free energy G of the volume is then described in terms of the order parameter and its gradients, and the rate at which the structure evolves with time is set in the context of irreversible thermodynamics, and depends on how G varies with . It is the gradients in thermodynamic variables that drive the evolution of structure. Consider a more complex example, the growth of a grain within a binary liquid (Fig.2). In the absence of fluid flow, in the sharp interface method, this requires the solution of seven equations involving heat and solute diffusion in the solid, the corresponding processes in the liquid, energy conservation at the interface and the GibbsThomson capillarity equation to allow for the effect of interface curvature on local equilibrium. The number of equations to be solved increases with the number of domains separated by interfaces and the location of each interface must be tracked during transformation. This may make the computational task prohibitive. The phase field method clearly has an advantage in this respect, with a single functional to describe the evolution of the phase field, coupled with equations for mass and heat conduction, i.e. three equations in total, irrespective of the number of particles in the system. The interface illustrated in Fig. 2b simply becomes a region over which the order parameter varies between the values specified for the phases on either side. The locations of the interfaces no longer need to be tracked but can be inferred from the field parameters during the calculation.

Fig. 2 (a) sharp interface

(b) diffuse interface


6

[Pick the date]

Notice that the interface in Fig. 2b is drawn as a region with finite width, because it is defined by a smooth variation in between = 0(solid) and =1(liquid). The order parameter does not change discontinuously during the traverse from the solid to the liquid. The position of the interface is fixed by the surface where =0.5. 2.1 - ORDER PARAMETER: The order parameter in phase field modeling is a function of space and time which may or may not have macroscopic physical interpretations. For two-phase materials, is typically set to 0 and 1 for the individual phases, and the interface is the domain where 0<<1. For the general case of N phases present in a matrix, there will be a corresponding number of phase field order parameters i with i=1 to N. i=1 then represents the domain where phase i exists, i=0 where it is absent and 0<i<1 its bounding interfaces. Suppose that the matrix is represented by o then it is necessary that at any location: N

=1
i

i=0 It follows that the interface between phases 1 and 2, where 0<1<1 and 0<2<1 is given by 1+ 2 =1; similarly, for a triple junction between three phases where 0<i<1 for i=1,2,3, the junction is the domain where 1+ 2+ 3 = 1. The order parameters in phase field modeling can be the either of two types: Conserved Order Parameters Conserved quantities as quite clearly decipherable, are the ones which remain unchanged during the process to be studied. Example can be the concentration of an element or alloy undergoing solidification, because the average concentration is never going to change. The change in conserved order parameters with time is governed by Cahn-Hilliard equation. Non-conserved Order Parameters Non-conserved quantities are the ones that change during a process. Example can be spin or crystalline order. The change in non-conserved order parameters with time is governed by Cahn-Allen equation.
7

[Pick the date]

2.2 GOVERNING EQUATIONS AND MATHEMATICS OF PHASE FIELD MODELLING: Derivations of the important expressions are given in full, on the premise that it is easier for a reader to skip a step than it is for another to bridge the algebraic gap between it is easily shown that and the ensuing equation J.E Hilliard (on the mathematics of their phase field model for spinodal decomposition) As a first requirement for any problem to be modeled by phase field modeling, a free energy functional (for isothermal cases and for non-isothermal cases free entropy functional) has to be defined as a function of order parameter. The general expression of a free energy functional is shown below: F = v [f (, c, T) + (2c/2)*| c|2 + (2/2)*| |2] dv (2)

The first term in the left hand side of the equation is free energy density of the bulk phase as a function of concentration, order parameter and temperature. The second and the third term denote the energy of the interface. The second term denotes the energy due to the gradient present in the concentration and the third term denotes the energy due to the gradient present in the order parameter. After doing a little bit of mathematics (which is intentionally ignored here, considering the point that only the application of these equations shall be sufficient at undergraduate level study), one arrives at two kinds of equation. The first one is for conserved order parameters and the second one is for non-conserved order parameters. Cahn-Hilliard Equation Cahn-Hilliard equation gives the rate of change of conserved order parameter with time. /t = M.2[f/ - 2 2] (3) The above equation is for constant (position-independent) mobility M. is order parameter, is divergence, f is free energy of the bulk, is gradient energy coefficient. As one can quite clearly notice that Cahn-Hilliard equation is nothing but modified form of Ficks second law for transient diffusion.

[Pick the date]

Cahn-Allen Equation - Cahn-Allen equation gives the rate of change of non-conserved order parameter with time. /t = -M [f/ 2 2] (4) Cahn-Allen equation is also known as time-dependent GinsburgLandau equation. Note: While deriving the Cahn-Hilliard and Cahn-Allen equation, an important expression /t = M*F/ was used, which clearly does makes sense because the change in free energy functional with respect to change in order parameter , must be in some relation with change in order parameter with respect to time. 2.3 PSEUDO ALGORITHM TO MODEL A PROCESS VIA PHASE FIELD MODELING: The diffuse interface idea can be extended to almost all systems with an evolution of microstructure or interfacial boundary involved in it. Here is the Pseudo algorithm to approach a problem: Describe the microstructure using a suitable set of field variables (commonly referred as order parameters), some are conserved variables and others are non-conserved. Write the energy of a configuration consistent with the systems thermodynamics. It will have both bulk and gradient energy terms, similar to as equation (2) Write the evolution equations: Cahn-Hilliard equation for conserved variables, and Cahn-Allen equation for nonconserved variables. Discretize the evolution equations via a suitable scheme (a variety of schemes can be found in the literature, each having their own advantages over others) to solve it numerically. Provide the system inputs as well as spacial and time information and then numerically march in time in order to evolve the order parameters. Hope (pray?) that your model will behave nicely!

[Pick the date]

3 EXAMPLES:
As already explained, a diffuse interface approach is very much capable to model almost any kind of problem in the field of material science. The most common problems that are solved are: Spinodal decomposition. Isothermal and non-isothermal solidification. Disorder order phase transformation. Grain growth during recrystallization. Dislocation dynamics. Dendritic growth. A problem involving a combination of above mentioned and many a others to count. In the present report, three kinds of problem i.e Spinodal Decomposition, Isotermal Solidification for pure substance and Dendritic Growth for pure substance are studied, their theory and a detailed method to solve them via phase field modeling is presented 3.1 SPINODAL DECOMPOSITION: Spinodal decomposition is a mechanism by which a solution of two or more components can separate into distinct regions (or phases) with distinctly different chemical compositions and physical properties. This mechanism differs from classical nucleation in that phase separation due to spinodal decomposition is much more subtle, and occurs uniformly throughout the material and not just at discrete nucleation sites. Spinodal decomposition is of interest for two primary reasons. In the first place, it is one of the few phase transformations in solids for which there is any plausible quantitative theory. The reason for this is the inherent simplicity of the reaction. Since there is no thermodynamic barrier to the reaction inside of the spinodal region, the decomposition is determined solely by diffusion. Thus, it can be treated purely as a diffusional problem. From a more practical standpoint, spinodal decomposition provides a means of producing a very finely dispersed microstructure that can significantly enhance the physical properties of the material.
10

[Pick the date]

3.1.1-Mechanism: In a binary mixture spinodal decomposition occurs when delocalized small amplitude fluctuations (concentration waves) grow spontaneously (see figure) as the time after the quenching proceeds. These local concentration Conc. Vs position fluctuations will lead to phase change in the thermodynamically unstable state. The mechanism should not occur in the whole two phase coexistence region of mixture, but rather only inside a smaller region, the boundary of which is given by the spinodal curve (or chemical spinodal). Suppose the plot of free energy vs composition for a binary mixture appears as shown in the figure, then the points of inflection enclose that region in which spinodal decomposition is going to occur at that temperatur. Spinodal curve is plotted by finding these points of inflection at different temperature and then joining them. As shown in the figure, the phase seperation inside the spinodal curve takes place via spinodal decomposition while in between spinodal curve and coexistence curve, it takes place via nucleation mechanism. In this region (outside the spinodal) , the systen is stable against such weak fluctuations and localised large amplitude fluctuations must form in order to start the transformation, such as a formation of nucleus. To be precisely correect, there is always a gradual change from spinodal to nucleation mechanism.
11

[Pick the date]

3.1.2-Phase Field approach: Let us consider a binary alloy of average composition Co occupying the (2D) xy-plane. Let the alloy consist of two phases m and p, that is, it is kept at a temperature (say T) that corresponds to the two phase region in the phase diagram (Fig 3). We assume that the temperature remains a constant, i.e., our present formulation is an isothermal one.

Fig 3 The order parameter in this case is going to be the composition at any point. The microstructure of the system will be completely described by the composition field. Let the composition at any point r in the xy-plane at time t be denoted by c(r, t). Given an initial composition profile, say c(r,0), the composition profile at any future time t can be obtained by
12

[Pick the date]

solving the equation:

following

(Cahn-Hilliard)

non-linear

diffusion

c /t = . M (5) where, M is the mobility, c is the (scaled) composition, t is the time, and is the chemical potential, given by = F /c (6) where /c denotes the variational derivative with respect to composition, and F is the free energy functional. F = v [f ( c) + (2c/2)*| c|2] dv or, F = v [f ( c) + *| c|2] dv (7) where is the gradient energy coefficient, and f(c) is the bulk free energy density, and is parameterized as a function of composition as f(c) = Ac2(1 - c)2 (8) where A is a positive constant indicating the energy barrier between the two equilibrium phases m and p (Fig 4)

Fig 4 Using the expressions (7) and (8) in the definition of the chemical potential, we obtain = h - 22c (9) where, h = f /c = 4Ac(1 - c)(1 - 2c) (10) We assume the mobility M and the gradient energy coefficient to be (scalar) constants: this amounts to assuming the interfacial

13

[Pick the date]

energies and the diffusivities to be isotropic. Using the Equation (9) above, and the fact that M is a constant, we obtain the Cahn-Hilliard equation as follows: c /t = M2(h - 22c) (11) 3.1.3-Numerical Method to solve equation (11): Since Cahn-Hilliard equation is nonlinear, it can only be solved numerically through discretization in space and time. Due to its simplicity and small memory requirement, most of the phase-field simulations in the literature employed the explicit forward Euler method in time and finite-difference in space. To maintain the stability and to achieve high accuracy for the solutions, the time step and spatial grid size have to be very small, which seriously limits the system size and time duration of a simulation. The ability to performing reliable long-time simulation is critical in the fundamental understanding of the scaling behavior of morphological pattern evolution. In this report, we implement an accurate and efficient semi-implicit Fourier-spectral method suggested by L.Q.Chen, J.Shen for solving the phase-field equations. For the time variable semi-implicit scheme is being employed. Thanks to the exponential convergence of the Fourier-spectral discretization, it requires a significantly smaller number of grid points to resolve the solution to within a prescribed accuracy, say 1%. Moreover, the semi-implicit Fourier spectral method is easy to implement and can be extended for systems with position-dependent mobility. Equation (11) is first non-dimensionalised and is then discretised by employing Semi-Implicit Fourier-Spectral Method. The discretised equation is: (k, t+t) = (k, t) tMk2(k,t) 1 + 2tMk4 (12)

In the above equation and represent the fourier transform of composition field and h field, where h is given by equation (10). k = (k1, k2) is a vector in the Fourier space, k = (k12+k22)1/2 is the magnitude of k. Other symbols have their usual meanings.
14

[Pick the date]

Algorithm for microstructural evolution: Given a composition profile at time t = 0, we calculate the h and its Fourier transform as well as the Fourier transform of C. Using and in equation (12), we calculate the composition profile at some future time t + t. The inverse Fourier transform of (t+t) gives the composition profile at time t + t. Repeat the above three steps to march in time for the given number of time steps. A code was developed in MATLAB using the above mentioned algorithm. Periodic boundary conditions were also used. The MATLAB code is being provided in APPENDIX A. The inputs needed for the simulation are as follows: N, M size of the mesh dx, dy distance between the nodes in x & y direction dt length of time step timesteps total number of timesteps A free energy barrier Mob Mobility Kappa gradient energy coefficient C(N,M) Initial composition field information Note: At every node a very small noise is added to its concentration value for starting the simulation. Because this noise is going to imitate the concentration wave happening in the real process. Only those changes (or evolutions) in concentration at the nodes will live which decrease the value of free energy functional equation (7). Hence the evolution of the composition profile will occur. 3.1.4-Observations: (1)-Effect of x, y and t on the stability and accuracy of the solution obtained by the Semi-Implicit Fourier Spectral Method to solve Cahn-Hilliard equation for Spinodal Decomposition:

15

[Pick the date]

All the results correspond to a double well free energy potential function with minima at 0 and 1. And the value of initial composition i.e c_0 = 0.5. Effect of x and y with t = 1.0 - Keeping all the thermodynamic parameters i.e A, kappa, mobility equal to one and a 256x256 mesh size, it was observed that stable solution was obtained in the range of x = y = [0.45, 2.5]. The accuracy of solution increased as the value was increased from 1.0. The increase in accuracy can be attributed to the fact that all the numerical methods to solve a Partial Differential Equation suffer inaccuracy and anomaly when the value of t is larger than x ( or y) beyond a certain value. Effect of t with and y = 1.0 - Keeping all the thermodynamic parameters i.e A, kappa, mobility equal to one and a 256x256 mesh size, it was observed that stable solution was obtained with t as large as 6.0 for the problem being examined. The accuracy of the solution was of course better with smaller value of t. None of the earlier proposed method to solve the corresponding discretized equations could be able to handle such a large value of time step. This is surely one of the major advantages of Semi-Implicit Fourier Spectral Method where the emphasis may be on quantitative analysis rather than precise accuracy. However, to be on the safer side for all the discussions that follow the values chosen are x = y = 1.0, t = 2.5. (2)- Effect of Initial Concentration on the solvability of Cahn-Hilliard equation for Spinodal Decomposition via Semi-Implicit Fourier Spectral method for a double free energy well potential function: All the results correspond to a double well free energy potential function with minima at 0 and 1. When all the thermodynamic parameters were set to one, the composition segregation occurred perfectly fine for initial concentration value c_0 in the range (0.3, 0.7).
16

[Pick the date]

However, when the initial composition value was chosen in the range of (0, 0.3] or [0.7, 1), the composition segregation did not occur and the concentration values at all points moved to the nearest free energy double well potential minima in each case. One possible explanation to this observation can be that when the value of c_0 is too small (or too large) to jump the free energy barrier, the composition segregation becomes impossible to be achieved. As a solution to this problem, the value of A was increased. It was observed that below c_0 = 0.3, value of A=1.3(or greater) was able to show composition segregation. However, the value of A has to be progressively increased as we approach c_0 nearer to 0.0. The same holds true when the value of c_0 is greater than or equal to 0.7 For all the discussions that follow, c_0 = 0.5 will be used in computations. (3) Effect of noise strength on the microstructure evolution during Spinodal Decomposition: Why the noise was introduced into the system? Spinodal Decomposition occurs when in a thermodynamically unstable initial state, long wavelength small amplitude stastical fluctuations grow spontaneously in amplitude as time proceeds. To simulate, these concentration waves (or simply concentration fluctuations), randomness about average composition is introduced by imparting a small noise to the system. Case1 - Noise = 0.5*10-1 Case2 - Noise = 0.5* 10-2

t = 2.5

t = 2.5

17

[Pick the date]

t = 62.5

t = 62.5

t = 150

t = 150

Conclusions drawn from the above results: Before we proceed, we must make a point clear that a noise value of 0.5*10-1 is quite large and is rarely experienced by any system, however just for the sake of comparison we have chosen it against the one that is having far more possibility of occurring. t = 2.5 - As the value of noise in increased, after the first time step, larger concentration segregation is observed in Case1. In Case1 the range of concentration was [0.4854, 0.5149], whereas in the Case2 the range was [0.498, 0.501]. t = 62.5 The concentration segregation was greater in Case1 than Case2 as it can be clearly seen in the images shown above as well. The area covered by the diffuse interface was also lesser for the Case1 then Case2.

18

[Pick the date]

t = 450 At this time, the shape and size of the region occupied by two phases in Case1 became more or less stable and the system seemed to have attained the final state with well segregated regions. However in Case2, the decomposition was not stabilized and the coarsening of the regions would have continued had the system been allowed more time to stand. From the above discussion, few conclusions that can be made are that with larger noise the segregation becomes easy and that system attains equilibrium state relatively quickly. Moreover at any time, the area covered by the diffuse interface in the microstructure is lesser. For the proceeding discussions, the value of noise strength will be 0.5*10-2 unless otherwise stated. (4) Effect of Mobility on the microstructure evolution during Spinodal Decomposition. Since Cahn-Hilliard equation is nothing but a modified form of Ficks law for transient diffusion equation. Thus, the mobility (M) in Cahn-Hilliard equation can be realized as a corresponding for Diffusivity(D) in the diffusion equation. Thus, the value of M shall make a difference in the time taken by the system to achieve equilibrium state. The results obtained after solving the CahnHilliard equation with increasing values of M were in good agreement as well. The following images will show the state of the microstructures with increasing values of M at t = 62.5.

M = 0.5

M = 1.0

19

[Pick the date]

M = 3.0

M = 5.0

Conclusions : The larger the value of M, the lesser the time is taken by the system to move towards equilibrium. As it can be clearly seen from the above images that coarsening of the precipitates is larger for M = 5.0 than other values of M and that too at just t = 62.5. Accuracy It was observed that the accuracy of the results suffered when the value of M was increased above 3. For the case with M = 5, the concentration values at various nodes lies in the range [-0.066, 1.1512]. This clearly shows that the concentration values shooted up from their respective limits. (5) Effect of thermodynamic barrier (A) on the microstructure evolution during Spinodal Decomposition. The change in the value of A led to some interesting results that are summarized as below: A = 0.25 - The concentration segregation could not be achieved. However when a large noise of 0.05 was provided (which is of course too large to occur in any physical example), the composition segregation began to occur. A = 0.50 The concentration segregation could not be achieved until 70 time steps and the concentration value at all the points remained more or less around 0.50. However, the segregation began to occur after 70 time steps and achieved its equilibrium state after number of time steps.

20

[Pick the date]

A = 1.0 - The segregation in this case occurred just fine right from the start and achieved equilibrium after around 300 time steps. A = 2.5 The segregation in this case occurred in almost no time. Actually it was so fast that that both the concentration limits were exceeded and just after 30 time steps, the range of concentration at different mesh nodes were [-0.4506, 1.4828]. This made worm like fluctuations to occur inside the microstructure (image shown below). Thus, much higher values of A are of course going to affect the accuracy.

This completes our discussion on spinodal decomposition.

3.2 - ISOTHERMAL SOLIDIFICATION FOR SINGLE COMPONENT WITH SURFACE ENTROPY ISOTROPY:
For modeling crystal growth from an undercooled pure substance, the system of variables consists of one pure and constant component (c = 1), of the inner energy e, and of an order parameter (x, t), called the phase-field variable. The value of (x, t) characterizes the phase state of the system and its volume
21

[Pick the date]

fraction in space at time t. In contrast to classical sharp interface models, the interfaces are represented by thin diffuse regions in which (x, t) smoothly varies between the values of associated with the adjoining bulk phases. For a solidliquid phase system, a phase-field model may be scaled such that (x, t) = 1 characterizes the region of the solid phase and (x, t) = 0 the region of the liquid phase. The diffuse boundary layer, where 0< (x, t) < 1, and the profile across the interface are schematically drawn in Fig. 5. The darker region is liquid.

Fig. 5 Since the phase field variable in this case is non-conserved, the Cahn-Allen equation will be solved. However, the reader must be informed that the final differential equation will not be same as equation (4) but instead it would be of the same form as (4). This is because phase field modeling is too complex to generalize equations for different processes. The free energy functional is given by:

F() =v (f() (a( ) +(w()/))) dv

(13)

22

[Pick the date]

This equation is similar to the free energy functional shown previously. The second term in the integral (enclosed in the brackets) reflects the thermodynamics of the interface is a small length scale parameter related to the thickness of the diffuse interface. FREE ENERGY FUNCTION: The free energy density uses two functions: a double well function and an interpolating function. Here we chose the two functions w() = 2(1 )2 and f()= ( L(T TM)/TM)* 2(32 ), respectively. This assumption will make (f+w)/ = 0 at both =0 and 1 for all temperatures as shown in Fig. 6.

Fig. 6

23

[Pick the date]

After applying the equation for the energy conservation, a phase field equation /t = M*F/ (where is kinetic mobility) and a little bit of mathematics, the details of which are intentionally avoided here, one arrives at the final partial differential equation for the evolution of the phase field variable .
/t =

a ( ) - w,()/ - f,(T,)/T
,

(14)

where a, , w,, and f, denote the partial derivative with respect to and , respectively. f(T,) =( L(T TM)/TM)* 2(3 2) w() = 2(1 )2 a( ) = ac2 ( )| |2 bulk free energy density double well potential gradient entropy density

where TM is the melting temperature and defines the surface entropy density of the solid liquid interface. For the system of two phases, the gradient entropy density reads a() = ||2. The double well potential is w() = 2(1 )2. And f(T,) =m*2(3 2), where m is a constant bulk energy density related to the driving force of the process, for example, to the isothermal undercooling T, for example, m = m(T). By utilizing the phase field equation and by inserting the values of above mentioned term in equation (14), the modified form of the partial differential equation becomes: t = (2 ) (18 (23 32 + )/) 6m(1 ) is the laplacian of the matrix. (15)

3.2.1 Numerical method to solve equation (15): The most straightforward discretization of finite differences with an explicit time marching scheme is used here by writing the phase-field equation (15) in discrete form and by defining suitable boundary conditions. i,jn+1 = i,j +t/{2((i+1,jn-2i,jn + i-1,jn)/x2 + (i,j+1n- 2i,jn+ i,j1n)/y2) - A*18/2(2(i,jn)3-3)( i,jn)2+ i,jn) B*6m/(i,jn(1- i,jn))} (16)

24

[Pick the date]

Here, A,B {0, 1} are introduced in order to switch on or off the corresponding terms in the phase-field equation for our later case study. Two types of boundary conditions are used in the simulation. Periodic boundary condition: A periodic boundary condition mimics an infinite domain size with a periodicity in the structure. The values of the boundary are set to the value of the neighboring cell from the opposite side of the domain. 1,j = Nx-1,j , Nx,j = 2,j with j = 1, . . . , Ny i,1 = i,Ny-1 , i,Ny = i,2 with 1 = 1, . . . , Nx Neumann Boundary condition: The component of normal to the domain boundary should be zero. In our rectangular domain, this can be realized by copying the value of the neighboring (interior) cell to the boundary cell: 1,j = 2,j , Nx,j = Nx-1,j with j = 1, . . . , Ny i,1 = i,2 , i,Ny = i,Ny-1 with 1 = 1, . . . , Nx Stability Condition: To ensure the stability of the explicit numerical method and to avoid generating oscillations, the following condition for the time step t depending on the spatial discretizations x and y must be fulfilled: t < (1/4)*(1/x2 + 1/y2)-1 A code in Matlab was developed to evolve the phase field variable using equation (16) and suitable boundary conditions. The code is being provided in APPENDIX B. The inputs needed for the simulation are as follows: N, M size of the mesh dx, dy distance between the nodes in x & y direction dt length of time step timesteps total number of timesteps p(N,M) Initial phase field variable information epsilon thickness of the interface m driving force Mob Kinetic Mobility Gamma Surface isotropy density A, B coefficients to switch on or off the respective . terms
25

[Pick the date]

3.2.2 Observations: To investigate the phase-field equation, we switch on the potential entropy contribution w () and the bulk driving force f () by setting the coefficients A = 1 and B = 1. The colorscale indicates = 1(solid) in yellow, = 0(liquid) in black, and the diffuse interface region in varying colors. (1)Diffuse Interface Thickness: A planar solidliquid front is placed in the center of the domain at Nx/2 with a sharp interface profile, with zero driving force m = 0 and with Neumann boundary conditions on each side. The effect of different values of the small length scale parameter: = 1 and = 10 responsible for the thickness of the diffuse interface is shown in Fig. 7a, Fig 7.b

Fig 7.a

Fig 7.b (2)Driving Force: As a next configuration, the three simulations shown in Fig. 8ac were performed with = 1 with different values of the driving force and with the initial configuration of Fig. (a). For m = 0, the initial planar front remains stable, for

26

[Pick the date]

m = 1 the solid phase (yellow color) grows, whereas for m = 1 the solid phase shrinks.

Fig. 8a

Fig. 8b

Fig. 8c

Fig. 8d (3)Phase-Field Simulation of growing nucleus: For the simulation in Fig. 9a, a solid nucleus is set in a 2D domain of NxNy = 100100 grid points, with A = 1, B = 1, with driving force

27

[Pick the date]

m = 2.5 and with Neumann boundary condition. Due to periodic boundary conditions in Fig. 9b the particle grows across the lower boundary and appears at the top boundary.

Fig. 9a

Fig. 9b

3.3 DEDRITIC SOLIDIFICATION:


Dendrites are formed when surface anisotropy is included in the system, which means that now there are going to be some preferred directions for solidification. Though there are many models to simulate dendritic growth in the literature but in the present report, a phase field model suggested by Ryo Kobayashi is being studied. 3.3.1 Phase Field Model: The model includes two variables; one is a phase field (r, t) and the other is a temperature field T(r, t). The variable (r, t) is an ordering parameter at the position r and the time t, = 0 means being liquid and = 1 solid. And the solid/liquid interface is expressed by the steep layer of connecting the values 0 and 1. Fig. 5 shows how the shape of crystal is described by the phase
28

[Pick the date]

field . In order to keep the profile of such form and to move it reasonably, we consider the following Ginzburg-Landau type free energy functional parameter:

similar to equation (2) including m as a

F = v [f (, m) + (2/2)*| |2] dv

(17)

where is a small parameter which determines the thickness of the layer. It is a microscopic interaction length and it also controls the mobility of the interface. f is a double-well potential which has local minimums at = 0 and 1 for each m. Here we take the specific form of f as follows: f(, m) = 1/44 - (1/2 1/3*m)3 + (1/4 -1/2*m)2 (18)

Anisotropy can be introduced by assuming that depends on the direction of the outer normal vector at the interface. So is represented as a function of the vector v = vi satisfying (,v) = (v) for >0. The outer normal vector is represented by - at the interface. Thus, we consider:

F = v [f (, m) + ((- )2/2)*| |2] dv

(19)

From the formula /t = f/ and further simplifying, we have the following evolution equation: /t = -/x(/y)+ /y(/x) + .(2) + (1-)(-0.5+m) (20) where is a small positive constant and /v=(/vi)i. The parameter m gives a thermodynamical driving force. Especially in two dimensional space, we can take = () where is an angle between v and a certain direction (for example the positive direction of the x-axis). means derivative with respect to . Equation (20) gives the evolution of the order parameter or phase field variable with time.

29

[Pick the date]

Here we assume that m is a function of the temperature T, for example, m(T) = (Te - T) where Te is an equilibrium temperature, which means that the driving force of interfacial motion is proportional to the supercooling there. But in the following -l[(Te - T)] where simulations, we used the form m(T)=( and are positive constants; < 1, since this assures |m(T)| < for all values of T. Also m(T) is almost linear for T near Te. To take anisotropy into account, let us specify to be: = 1 + cos(j(-o)) (21)

The parameter means the strength of anisotropy and j is a mode number of anisotropy. Side branching can be stimulated in the dendrites by adding a small random noise in the equation (20). The noise can be of the form a(1-)x, where a is the strength of noise and x is a random number in the range[-0.5,0.5]. The equation for T is derived from the conservation law of enthalpy as: T/t = 2T + Kp/t (22)

T is non-dimensionalized so that the characteristic cooling temperature is 0 and the equilibrium temperature is 1. K is a dimensionless latent heat which is proportional to the latent heat and inversely proportional to the strength of the cooling. For simplicity, the diffusion constant is set to be identical in both of solid and liquid regions. (22) is a heat conduction equation having a heat source along the moving interface, since Kp/t has non-zero value only when the interface passes through the point. 3.3.2 Numerical method to simulate the Dendritic growth: The simplest finite difference scheme with a nine point laplacian is used to solve equations (20) and (22). First, the new value of phase field variable is calculated to substitute it into the temperature field to get the new value of temperature field. Nine point laplacian is required for the stability of the solution. A code in Matlab was developed to evolve the phase field variable and temperature field using equations (20) and (22)
30

[Pick the date]

respectively and periodic boundary conditions. The code is attached along with this report in the CD-Drive. The inputs needed for the simulation are as follows: N, M size of the mesh dx, dy distance between the nodes in x & y direction dt length of time step timesteps total number of timesteps p(N,M) initial phase field variable information T(N,M) initial temperature field information K latent Heat TAU phase field relaxation time EPS interfacial width DELTA strength of anisotropy ANISO mode number of anisotropy ALPHA positive constant GAMMA positive constant 3.3.2 Observations: The code written in Matlab was run using the following inputs: N= 500; M=500; dx=0.03; dy=0.03;dt=0.0003; K=4; TAU=0.0003, EPS= 0.01; GAMMA=10.0; DELTA=0.02; ANISO=4.0;ALPHA=0.9; Noise was not added to the system. The following was the structure of dendrites obtained.

The yellow region indicates solid, black region indicates liquid while the rest in indicated by the interface. Notice that there is directional solidification in four directions because ANISO = 4.0 .

31

[Pick the date]

Effect of latent heat of solidification K: The below shown four images correspond to a dimensionless latent heat K =0.8, 1.0, 1.2 and 1.8 from left to right and top to bottom. Notice that as the value of latent heat is increased, the dendritic structure becomes much defined because release of latent heat is the driving force for the process. Also noise is added to stimulate side branching.

32

[Pick the date]

(4) CONCLUSION
Phase-field models have been successfully applied to various materials processes including solidification, solid-state phase transformations, coarsening, and growth. With the phase-field approach, one can deal with the evolution of arbitrary morphologies and complex microstructures without explicitly tracking the positions of interfaces. This approach can describe different processes such as phase transformations and particle coarsening within the same formulation, and it is rather straightfoward to incorporate the effect of coherency and applied stresses, as well as electrical and magnetic fields. Efforts are being made in Material Science to focus on the exploration of novel applications of the phase-field method to various materials problems, e.g., problems involving simultaneous long-range elastic and electric or magnetic dipole-dipole interactions, lowdimensional systems such as thin films and multilayer structures, and interactions between phase and defect microstructures such as random defects and dislocations. There will also be increasing efforts in establishing schemes to obtain the phase-field parameters directly from more fundamental first-principles electronic structure or atomic calculations. For practical applications, significant additional efforts are required to develop approaches for connecting phase-field models with existing or future thermodynamic, kinetic, and crystallographic databases.

33

[Pick the date]

APPENDIX A
MATLAB code for spinodal decomposition:
%Numerical solution of Cahn%Hilliard Equation via Semi%implicit Fourier Spectral %method for conserved %quantities. %Non-dimensionalisation is done %and isothermal conditions are %considered %For further details, refer to %the research paper"Applications %of semi-implicit Fourier%spectral method to phase field %equations"-L.Q. Chen, Jie Shen comp(j+M*(i-1)) = c_0 + noise_str*(0.5-rand); end end

%The half_N and half_M are %needed for imposing the %periodic boundary conditions half_N = N/2; half_M = M/2; del_kx = (2.0*pi)/(N*del_x); del_ky = (2.0*pi)/(M*del_y); for index = 1:ntmax %calculate g, g is parameterised %as 2Ac(1-c)(1-2c) for i = 1:N for j = 1:M g(j+M*(i-1)) = 2*A*comp(j+M*(i-1))*(1comp(j+M*(i-1)))*(12*comp(j+M*(i-1))); end end %calculate the fourier transform %of composition and g field f_comp = fft(comp); f_g = fft(g); %Next step is to evolve the &composition profile for i1 = 1:N if i1 < half_N kx = i1*del_kx; else kx = (i1-N-2)*del_kx; end kx2 = kx*kx; for i2 = 1:M if i2 < half_M ky = i2*del_ky; else ky = (i2-M2)*del_ky; end ky2 = ky*ky; k2 = kx2 + ky2;

clear clc format long %spatial dimensions -- adjust N %and M to increase or decrease %the size of the computed %solution. N =256; M = 256; del_x = 1.0; del_y = 1.0; %time parameters -- adjust ntmax %to take more time steps, and %del_t to take longer time %steps. del_t = 2.5; ntmax = 500; %thermodynamic parameters A = 1.0; Mob = 1.0; kappa = 1.0; %initial composition and noise %strenght information c_0 = 0.50; noise_str = 0.5*(10^-2); %composition used in %calculations with a noise for i = 1:N for j = 1:M

34

[Pick the date]

k4 = k2*k2; denom = 1.0 + 2.0*kappa*Mob*k4*del_t; f_comp(i2+M*(i1-1)) = (f_comp(i2+M*(i1-1))k2*del_t*Mob*f_g(i2+M*(i11)))/denom; end end %Let us get the composition back %to real space comp = real(ifft(f_comp)); disp(comp); disp(index);

%for graphical display of the %microstructure evolution, %lets store the composition %field into a 256x256 2-d %Matrix. for i = 1:N for j = 1:M U(i,j) = comp(j+M*(i1)); end end %visualization of the output figure(1) image(U*55) colormap(Jet) end disp('done');

APPENDIX B
MATLAB code for isothermal solidification with no anisotropy:
% Isothermal solidifation of a %single component liquid phase %using explicit finite %difference scheme without any %surface aniostropy. % For more details, refer to %chapter 7-Phase Field %Modelling,Britta Nestler % -Computational Materials %Engineering, Bernaad Jansenns. clear clc %input parameters epsilon = 1.0; %thickness %of the interface gamma = 1.0; %surface %anistropy density m = -2.5; %driving %force Mob = 1.0; %kinetic %mobility Nx = 100; Ny = 100; %mesh size dx = 0.1; dy = 0.1; %distance %between two consecutive nodes dt = 0.001; %length of %time step timesteps = 100; %total %number of time steps A = 1.0; %A and B %are the coefiecients to switch B = 1.0; %on and %off the corresponding terms. %Lets generate the initial %picture p=zeros(Nx,Ny); for i1=Nx/2-10:Nx/2+10 for i2=Ny/2-10:Ny/2+10 p(i1,i2)=1; end end %evolution of the order %parameter for index=1:timesteps Lap = (4*del2(p))/(dx^2); %Laplacian of the matrix p for i1=1:Nx for i2=1:Ny p(i1,i2) = p(i1,i2) + (dt/Mob)*((2*gamma*Lap(i1,i2))A*(18*gamma*((2*(p(i1,i2)^3))(3*(p(i1,i2)^2))+p(i1,i2))/(epsi lon^2))-B*(6*m*(p(i1,i2)*(1p(i1,i2)))/epsilon)); end end %periodic boundary condition for i1=1:Nx p(i1,1) = p(i1,Ny-1); p(i1,Ny)= p(i1,2); 1

[Pick the date]

end for i2=1:Ny p(1,i2) = p(Nx-1,i2); p(Nx,i2)= p(2,i2); end %neumann boundary condition %for i1=1:Nx % p(i1,1) = p(i1,2); % p(i1,Ny)= p(i1,Ny-1); %end %for i2=1:Ny % p(1,i2) = p(2,i2); % p(Nx,i2)= p(Nx-1,i2); %end %visualization of the output

disp(p) figure(1) image(p*50) colormap('hot') end

APPENDIX C
MATLAB code for dendritic solidification based on Kobayashis Model:
%Phase Field Modelling of %dendritic growth as suggested %by Ryo Kobayashi %For more details refer to %"Modeling and numerical %simulationsof dendritic crystal %growth-Ryo Kobayashi clear clc K = 4.0; %Latent %heat TAU =0.0003; %PF %relaxation time EPS =0.01; %interfacial %width DELTA= 0.02; %modulation %of the interfacial width ANGLEO =0.0; %orientation %of the anisotropy axis ANISO= 4.0; %anisotropy %2*PI/ANISO ALPHA =0.9; %m(T) = %ALPHA/PI * atan(GAMMA*(TEQ-T)) GAMMA= 10.0; TEQ =1.0; %melting %temperature NX= 500; %size of the %mesh NX*NY NY =500; H= 0.03; %spatial %resolution DT =0.0003; %temporal %resolution timesteps =2000000; %number of %time steps pi =3.14159265358; %intial temperature and phase %field information T = zeros(NY,NX); p = zeros(NY,NX); for i1=1:NY for i2=1:NX if ((i1-NY/2)*(i1NY/2)+(i2-NX/2)*(i2-NX/2)<100) p(i1,i2) = 1.0; else p(i1,i2) = 0.0; end end end

%preallocation of matrices for %faster calculations grad_p_X = zeros(NY,NX); grad_p_Y = zeros(NY,NX); aX = zeros(NY,NX); aY = zeros(NY,NX); eps2 = zeros(NY,NX);

[Pick the date]

angle = zeros(NY,NX); epsilon_prime = zeros(NY,NX); epsilon = zeros(NY,NX); dXdY = zeros(NY,NX); dYdX = zeros(NY,NX); grad_eps2_X = zeros(NY,NX); grad_eps2_Y = zeros(NY,NX); lap_p = zeros(NY,NX); lap_T = zeros(NY,NX);

angle(i1,i2) = atan(grad_p_Y(i1,i2)/grad_p_X(i1 ,i2)); end if (grad_p_X(i1,i2)>0.0 && grad_p_Y(i1,i2)<=0.0) angle(i1,i2) = 2.0*pi + atan(grad_p_Y(i1,i2)/grad_p_X(i1 ,i2)); end if(grad_p_X(i1,i2)<0.0) angle(i1,i2) = pi + atan(grad_p_Y(i1,i2)/grad_p_X(i1 ,i2)); end end end for i1 =1:NY for i2=1:NX epsilon(i1,i2) = EPS*(1.0 + DELTA*cos(ANISO*(angle(i1,i2)ANGLEO))); epsilon_prime(i1,i2) = EPS*ANISO*DELTA*sin(ANISO*(angle (i1,i2)-ANGLEO)); end end for i1 = 1:NY for i2=1:NX aY(i1,i2) = epsilon(i1,i2)*epsilon_prime(i1, i2) * grad_p_Y(i1,i2); aX(i1,i2) = epsilon(i1,i2)*epsilon_prime(i1, i2) * grad_p_X(i1,i2); eps2(i1,i2) = epsilon(i1,i2)*epsilon(i1,i2); end end for i1=1:NY for i2=1:NX ip = mod(i2,NX)+1; im = mod((NX+i22),NX)+1; jp = mod(i1,NY)+1; jm = mod((NY+i12),NY)+1; dXdY(i1,i2) = (aY(i1,ip) - aY(i1,im))/H;

for index=1:timesteps %calculation of all the relevant %matrices for i1=1:NY for i2=1:NX ip = mod(i2,NX)+1; im = mod((NX+i22),NX)+1; jp = mod(i1,NY)+1; jm = mod((NY+i12),NY)+1; grad_p_X(i1,i2) = ((p(i1,ip) - p(i1,im))/H); grad_p_Y(i1,i2) = ((p(jp,i2) - p(jm,i2))/H); lap_p(i1,i2) = (2.0*(p(i1,ip)+p(i1,im)+p(jp,i2) +p(jm,i2))+p(jp,ip)+p(jm,im)+p(j p,im)+p(jm,ip)12.0*p(i1,i2))/(3.0*H*H); lap_T(i1,i2) = (2.0*(T(i1,ip)+T(i1,im)+T(jp,i2) +T(jm,i2))+T(jp,ip)+T(jm,im)+T(j p,im)+T(jm,ip)12.0*T(i1,i2))/(3.0*H*H); end end for i1 = 1:NY for i2=1:NX if (grad_p_X(i1,i2)==0.0 && grad_p_Y(i1,i2)>0.0) angle(i1,i2) = 0.5*pi; end if (grad_p_X(i1,i2)==0.0 && grad_p_Y(i1,i2)<=0.0) angle(i1,i2) = 0.5*pi; end if (grad_p_X(i1,i2)>0.0 && grad_p_Y(i1,i2)>0.0)

[Pick the date]

dYdX(i1,i2) = (aX(jp,i2) - aX(jm,i2))/H; grad_eps2_X(i1,i2) = (eps2(i1,ip) - eps2(i1,im))/H; grad_eps2_Y(i1,i2) = (eps2(jp,i2) - eps2(jm,i2))/H; end end for i1=1:NY for i2=1:NX po = p(i1,i2); m = (ALPHA/pi) * atan(GAMMA*(TEQ-T(i1,i2))); scal= grad_eps2_X(i1,i2)*grad_p_X(i1,i 2)+ grad_eps2_Y(i1,i2)*grad_p_Y(i1,i 2); %evolution of the phase field variable

p(i1,i2) = p(i1,i2)+((dXdY(i1,i2)+dYdX(i1,i 2) + eps2(i1,i2)*lap_p(i1,i2)+ scal + po*(1.0-po)*(po0.5+m))*DT/TAU); %evolution of temperature field T(i1,i2) = T(i1,i2)+(lap_T(i1,i2)*DT) + (K*(p(i1,i2) - po)); end end %visualization of the output disp(p) figure(1) image(p*50) colormap('hot') end

You might also like