Dual-scale phase-field simulation of Mg-Al alloy solidification

Similar documents
Evaluation of dendrite morphology using fractal dimension and dimensionless perimeter in unidirectionally solidified Al-Si Alloys

Thermodynamics and Microstructure: Recent Examples for Coupling of Thermodynamic and Mobility Data to the Software MICRESS

Multi-scale simulation of ductile iron casting

Original. DOI: /

Modeling of microstructure evolution of magnesium alloy during the high pressure die casting process

Refining grain structure and porosity of an aluminium alloy. with intensive melt shearing

Study on rheo-diecasting process of 7075R alloys by SA-EMS melt homogenized treatment

Pre-Course Reading for ESI s Solidification Course

Metallurgy - Lecture (2) Solidification

Surface formation in direct chill (DC) casting of 6082 aluminium alloys

Modelling of long-term phase stability in Ni-based superalloys based on thermodynamic and kinetic CALPHAD calculations

K S T S ' = K L T L ' + vl v

Numerical optimization of solution heat treatments of single crystal nickel-based superalloys: Methods and validation

Effect of solidified structure on hot tear in Al-Cu alloy

Phase field simulation of the columnar dendritic growth and microsegregation in a binary alloy

Thermodynamic Calculations in Mg-alloy Systems

Grain Morphology of As-Cast Wrought Aluminium Alloys

TALAT Lecture Phase Diagrams. 14 pages, 13 Figures. Basic Level

Vibration influence on structure and density of aluminum alloys

Lecture 31: Principles of Solidification of Steel. Key words: Solidification, planar growth, dendritic growth, casting, constitutional supercooling

Numerical modelling of the solidification of ductile iron

Fractal characteristics of dendrite in aluminum alloys

Prediction of mechanical properties of Al alloys with change of cooling rate

Grain Refinement for Improved Lead-Free Solder Joint Reliability

The Effects of Superheating Treatment on Distribution of Eutectic Silicon Particles in A357-Continuous Stainless Steel Composite.

Effects of silicon and chromium additions on glass forming ability and microhardness of Co-based bulk metallic glasses

MICROSTRUCTURE OF RAPIDLY SOLIDIFIED POWDERS

Phase Diagrams. Phases

SIMULATION OF DIFFUSIONAL PROCESSES DURING SOLIDIFICATION IN AUSTENITIC STEELS

Structure and phase composition of the superalloy on the basis of Ni-Al- Cr alloyed by Re and La

Modification of Ohnaka back diffusion equation

Schematic representation of the development of microstructure. during the equilibrium solidification of a 35 wt% Ni-65 wt% Cu alloy

Simulation of Solute Redistribution during Casting and Solutionizing of Multi-phase, Multi-component Aluminum Alloys

ROLE OF SOLUTE AND TRANSITION METALS IN GRAIN REFINEMENT OF ALUMINUM ALLOYS UNDER ULTRASONIC MELT TREATMENT

Diffusion Simulations in High-Entropy Alloys

Evolution of the Specific Solid-Liquid Interface Area in Directional Solidification

Effect of pouring temperature on semi-solid slurry of ZL101 alloy prepared by slightly electromagnetic stirring

Hot Cracking Susceptibility in the TIG Joint of AZ31 Mg-Alloy Plates Produced by the TRC Process with and without Intensive Melt Shearing

Process-scale modelling of microstructure in direct chill casting of aluminium alloys

Grain Refinement of an Al-2 wt%cu Alloy by Al3Ti1B Master Alloy and Ultrasonic Treatment

Solidification and Crystallisation 5. Formation of and control of granular structure

Investigation of Stress Corrosion Cracking Behaviors of an AZ91 Magnesium Alloy in 0.1 kmol/m 3 Na 2 SO 4 Solution Using Slow Strain Rate Test

Microstructure evolution during sand casting in AZ80 Mg alloy

Modeling Macrosegregation with a Multiphase Approach: Examples and Discussions

Phase-field model for mixed mode transformations and its perspective for steel

Numerical Simulation of Dendrite Growth during Solidification

Chapter 9 Phase Diagrams. Dr. Feras Fraige

A.S. Kiran 1, V. Desai 2, Narendranath 2 and P.G. Mukunda 1. R G g

Determination of the cast structure parameter on the basis of micro-segregation analysis

Continuous Rheocasting for Aluminum-Copper Alloys

Numerical Simulation of Solidification Structure Formation during Continuous Casting in Fe 0.7mass%C Alloy Using Cellular Automaton Method

but T m (Sn0.62Pb0.38) = 183 C, so this is a common soldering alloy.

Chapter 10: Phase Diagrams

CALCULATION AND EXPERIMENTS ON THE SOLIDIFICATION BEHAVIOR OF SINGLE CRYSTAL SUPERALLOY CONTAINING Ru

Anomalous Halo Formation in an Arc-Melted ScNi-Sc 2 Ni Off-Eutectic Binary Alloy

Two Components System

Learning Objectives. Chapter Outline. Solidification of Metals. Solidification of Metals

THE EFFECT OF CERIUM ADDITION ON THE MICROSTRUCTURE OF AlSiMgCe ALLOY. Anasyida Abu Seman and Abdul Razak Daud.

the Phase Diagrams Today s Topics

Grain Refinement of Al-Si Alloys by Nb-B Inoculation. Part 1: Concept Development and Effect on Binary Alloys. Part 2: Application to Commercial

Solidification & Binary Phase Diagrams II. Solidification & Binary Phase Diagrams II

Microstructures and mechanical properties of Mg Zn Zr Dy wrought magnesium alloys

27-301, Fall 02, Rollett. Total marks = 100 (20 per question). 1 st Homework, due Sept. 6th

CONVENTIONAL AND SEMI-SOLID A356 ALLOY WITH ADDITION OF STRONTIUM

The peritectic transformation, where δ (ferrite) and L (liquid)

The Effect of Rotation Stirring on Macrosegregation in Bi-Sn Alloy

SEGREGATION DURING SOLIDIFICATION

Point Defects. Vacancies are the most important form. Vacancies Self-interstitials

Development of creep-resistant magnesium casting alloys for high temperature automotive applications

ENGINEERING COUNCIL CERTIFICATE LEVEL ENGINEERING MATERIALS C102 TUTORIAL 3 THERMAL EQUILIBRIUM (PHASE) DIAGRAMS

SOLIDIFICATION MICROSTRUCTURES OBTAINED BY A NOVEL TWIN-SCREW LIQUIDUS CASTING METHOD

EFFECTS OF MAGNESIUM ON NIOBIUM SEGREGATION AND IMPACT TOUGHNESS IN CAST ALLOY 718

Homogenization Model for 7xxx Aluminum Alloys

Phase Diagrams of Pure Substances Predicts the stable phase as a function of P total and T. Example: water can exist in solid, liquid and vapor

Analysis of Yield Rate in Single Crystal Casting Process Using an Engineering Simulation Model

Influence of pouring methods on filling process, microstructure and mechanical properties of AZ91 Mg alloy pipe by horizontal centrifugal casting

Modeling: (i) thermal spray rapid solidification (ii) partially molten particle impact

Module 22. Solidification & Binary Phase Diagrams V. Lecture 22. Solidification & Binary Phase Diagrams V

Partial melting and re-solidification in partially melted zone during gas tungsten arc welding of AZ91 cast alloy

EFFECT OF SILICON CONCENTRATION ON THE CAST MACROSTRUCTURE OF Fe-Si STEELS FOR ELECTRIC APPLICATION

High speed steels are widely used in making highspeed

The effect of alloy elements on the density variation of steel melt at the interdendritic region during solidification

Index Terms Solidification, Zinc-base alloys, thermal parameters, columnar-to-equiaxed transition. Fig. 1: Schematic of experimental setup.

The challenge of service life prediction related corrosion modelling

Recently, more aggressive jet engines and industrial gas

EXPERIMENTAL INVESTIGATION ON COOLING RATE FOR CENTRIFUGAL CASTING Kirti Kanaujiya, Yugesh Mani Tiwari Department of Mechanical Engineering

Hot Deformation Behavior of High Strength Low Alloy Steel by Thermo Mechanical Simulator and Finite Element Method

Title. CitationElectronics Goes Green (EGG 2004) September 6- Issue Date Doc URL

Institute of Materials Science, Helmholtz-Zentrum Geesthacht, Max-Planck-Strasse 1, D Geesthacht, Germany

Solidification microstructure development

J = D C A C B x A x B + D C A C. = x A kg /m 2

Influence of Semi-Solid Isothermal Heat Treatment on Microstructure of Gray Cast Iron

The accuracy of a 2D and 3D dendritic tip scaling parameter in predicting the columnar to equiaxed transition (CET)

Numerical Simulation of Cast Distortion in Gas Turbine Engine Components

Effect of Zn content on microstructure, mechanical properties and fracture behavior of Mg-Mn alloy

EXPERIMENTAL STUDY OF THE Cu-Al-Sn PHASE EQUILIBRIA, CLOSE TO THE COPPER ZONE

The Al-Si cast alloys are widely used in the automotive

Influence of Heat Treatment on the Distribution of Ni 2 Nb and Microsegregation in Cast Inconel 718 Alloy

Diffusional Transformations in Solids

AGING BEHAVIOR IN CO-CR-MO-C ALLOYS

Transcription:

IOP Conference Series: Materials Science and Engineering PAPER OPEN ACCESS Dual-scale phase-field simulation of Mg-Al alloy solidification To cite this article: A Monas et al 2015 IOP Conf. Ser.: Mater. Sci. Eng. 84 012069 View the article online for updates and enhancements. Related content - Microsegregation in multicomponent alloy analysed by quantitative phase-field model M Ohno, T Takaki and Y Shibuta - Detached Melt Nucleation during Diffusion Brazing of a Technical Ni-based Superalloy: A Phase-Field Study B Böttger, M Apel, B Laux et al. - Phase-field modeling of liquid droplet migration in a temperature gradient S Pan, M Zhu and M Rettenmayr This content was downloaded from IP address 37.44.204.3 on 21/11/2017 at 19:42

Dual-scale phase-field simulation of Mg-Al alloy solidification A Monas 1, O Shchyglo 1, D Höche 2, M Tegeler 1 and I Steinbach 1 1 Interdisciplinary Centre for Advanced Materials Simulations (ICAMS), Ruhr University Bochum, Universitätsstraße 150, Bochum, 44801, Germany 2 Institute of Materials Research, Helmholtz-Zentrum Geesthacht, Max-Planck-Straße 1, Geesthacht, 21502, Germany E-mail: alexander.monas@rub.de Abstract. Phase-field simulations of the nucleation and growth of primary α-mg phase as well as secondary β-phase of a Mg-Al alloy are presented. The nucleation model for α- and β-mg phases is based on the free growth model by Greer et al.. After the α-mg phase solidification we study a divorced eutectic growth of α- and β-mg phases in a zoomed in melt channel between α-phase dendrites. The simulated cooling curves and final microstructures of α-grains are compared with experiments. In order to further enhance the resolution of the interdendritic region a high-performance computing approach has been used allowing significant simulation speed gain when using supercomputing facilities. 1. Introduction Magnesium alloys are among the most promising structural materials for lightweight applications due to their low density compared to aluminum and steel. In Mg technology the most applied alloying element for technical alloys like AZ91, AM50 (cast) or AZ31 (wrought) is aluminum. The mechanical properties of such Mg-Al alloys are primarily determined by their micro-structure which typically consists of two main phases: the Mg-rich hexagonal close packed (HCP) α-phase and a near stoichiometric Mg 17 Al 12 β-phase, and further minor secondary phases due to third alloying elements. From a scientific point of view it remains a challenge to simulate the dendritic solidification structure of the primary HCP-α-phase. In particular the anisotropy function of the interface energy between liquid and the α-phase which determines the growth structure of the dendrite, remains under debate [1, 2]. A second question is the morphology of interdendritic eutectic, which is decisive for the corrosion properties of the alloy [3]. In the present work we apply the phase-field method [4, 5, 6, 7] to simulate the microstructure evolution of Mg-Al alloys during solidification. We apply the anisotropy function as given in [8] which produces a dendrite trunk normal to the basal plane of the HCP crystal. While several authors reported 2-dimensional and 3-dimensional phase-field simulations eg.[8, 9, 2] of Mg-alloy solidification, a new challenge is to reveal the eutectic structure, which is assumed to happen in a divorced mode, but is at least one order of magnitude finer than the primary dendritic structure. 2. Experimental details The binary Mg-Al test alloys were prepared by permanent mold indirect chill casting as described in [10]. The alloys were melted in an electric resistance furnace under a protective atmosphere Content from this work may be used under the terms of the Creative Commons Attribution 3.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI. Published under licence by Ltd 1

of 2% SF 6 and Ar. The composition was chosen to be Mg - 9 wt% Al. After mixing, the melt was held at 1020 K for 10 min and then poured into a standard steel mold preheated to 933 K. In addition, the mold containing the melt and a temperature sensor were held at 933 K for 15 min and lowered into water at a rate of 10 mm/s until the top of the melt was in line with the cooling water level. The size of the cast ingot is 250 mm in length and 105 mm in diameter. T e m p e r a t u r e [ K e l v i n ] 1 0 0 0 9 0 0 8 0 0 7 0 0 ( a ) ( b ) ( c ) 6 0 0 0 5 0 0 1 0 0 0 1 5 0 0 2 0 0 0 2 5 0 T i m e [ s e c o n d s ] Figure 1. Experimentally measured cooling curves of Mg-9wt.%Al: (a) - furnace controlled cooling, (b) - air cooling, and (c) - submersion in water. 3. Phase-field model In this study we apply the multi-phase-field method on a meso-scale with locally linearized phase diagram information [8]. In this method each grain of a distinct crystallographic phase is attributed by its own phase-field variable φ p ( x, t) [0, 1]. Three distinct crystallographic phases are considered: liquid melt, Mg-rich α-phase and a near stoichiometric Mg 17 Al 12 β- phase, where each solid grain is attributed by an identifier of the crystallographic phase and its orientation in space (for details see [5, 8]). For the interface between melt and the α-phase an anisotropic interface energy σ pq = σ pq ( n) as a function of the interface normal n with respect to the crystallographic orientation of each grain is used [8]. The interface energy between the melt and β-phase as well as between the solid α- and β-phases assumed isotropic. In order to consider the effect of the external cooling rate on nucleation and growth of the solid phases we solve the heat balance in average over the whole domain with n xyz discrete elements: ρc p T = n xyz p q ψ pq L pq + Q ext, (1) where T is the temperature of the system, Q ext is the heat extraction rate, ρ and c p are the mass density and the specific heat which are taken constant throughout the simulation. ψ pq are the local transformation rates between pairs of phases which sum up to the total transformation rate of individual phases φ p = N q p ψ pq in the interface or a junction point with N phases (for details see [11]). L pq is the latent heat associated with the individual transformations between phases p and q. In the present study we treat all possible transformations between liquid α, liquid β and α β phases. 2

Table 1. Parameters of the linearized Mg-Al phase diagram as shown in Fig. 2 intersection concentration C α liq. 0.00 % slope (liquidus) m liq. α -6.90 K/% intersection temperature T α liq. 923.15 K slope (solidus) m α liq. -27.58 K/% intersection concentration C β liq. 45.64 % slope (liquidus) m liq. β 3.50 K/% intersection temperature T β liq. 760.60 K slope (solidus) m β liq. 9.00 K/% intersection concentration C α β 34.20 % slope (solvus) m α β 22.40 K/% intersection temperature T α β 1301.84 K slope (solvus) m β α -103.00 K/% 4. Phase diagram In order to obtain the phase diagram along with the exact transformation temperatures and slopes for our calculations the CalPhaD (Calculation of Phase Diagrams) method was used via the Thermocalc TCS 4.0 database TCBIN. For convenience the obtained phase diagram was linearized for the simulations as it is shown in Fig. 2 Figure 2. Mg-Al binary phase diagram with linearization (red). All slopes but m liq. α and m α liq. are taken at the eutectic temperature of 709.4K. Solidusand liquidus-slopes are the slopes between the points (923.15K; 0.0%Al) and (709.4K; 31.0%Al) [eutectic point] or (709.4K; 11.6%Al) respectively. This particular version alters the point of maximum solubility in the α from 11.6% to 7.75% to match the area of the original liquid α two-phase area in the phase diagram. Thus the solidus slope (m α liq. ) was adjusted from the original -18.43 K/% to -27.58 K/%. Also C α β and T α β were adjusted to retain m α β. Consequently the original α β intersection-point (1229.84K ; 34.83%Al) was adjusted to (1301.84K ; 34.20%Al). All of the used values for the characterization of the phase diagram are summarized in table 1. 3

Table 2. Nucleation model parameters particle density for α-phase nucleation K α particle density for β-phase nucleation K β a of the normal size-distribution (mean) b of the normal size-distribution (standard deviation) 0.2 10 14 m 3 1.5 10 16 m 3 1.5 10 7 m 1.0 10 7 m 5. Nucleation model In order to capture the dependence of the nucleation density on the cooling rate we employ the so-called free growth model by Greer and coworkers [13]. This model is based on a given distribution of seed particles of different sizes. Assuming perfect wetting conditions between the primary α-phase and the seed particles, the hemispheric cap of the α-phase on the seed particles defines the critical undercooling for free growth of the nucleating phase which is inversely proportional to the size of the seed particles [12]. In our simulations we start from a random distribution of the particles with a given density and size distribution over the entire simulation domain initially filled with liquid melt. The utilized normal distribution is given by N(d) = K 1 b [ 2π exp (d a) 2/ 2b 2]. (2) Each particle is assigned an effective size d which determines the critical undercooling for its activation. The resulting number of particles is typically higher than the actual number of the activated particles (active nucleation sites). Activation happens if a grid cell, which contains a particle, reaches it s critical undercooling, i.e. if the difference between actual temperature and melting temperature dependent on the local alloy concentration exceeds the local critical undercooling. A similar model has also been applied by Eiken et al. [8]. The parameters of the utilized nucleation model are summarized in table 2. Note, that only the tail of a normal distribution is used, thus a is negative, but only seeds with a positive size are created. We also assume a significantly higher nucleation density of the β-phase in the melt channels between the primary α-phase dendrites. 6. Simulation results The simulations are performed on two different scales. Although, the primary solidification in three-dimensional domains can be investigated in order to reveal the effect of the cooling rate on the nucleation density of α-phase dendrites (see Fig. 3), it is computationally demanding to study the entire solidification process in 3D over the temperature range of more than 200K even with high cooling rates. In contrast, in the 2D simulations it is possible to investigate the entire solidification process starting from the primary α-phase dendritic solidification and continuing with the eutectic solidification of the secondary β-phase upon cooling below the eutectic point. It shall be noted, that on a scale of primary α-phase dendrites the eutectic structure in the interdendritic region cannot be resolved and as the result the secondary β-phase builds a bulk structure between the primary dendrites as can be seen in Fig. 4. In order to investigate eutectic growth in detail we zoom into the channel between two primary dendrites upon reaching the eutectic temperature. Here we select the scale of the simulations by a factor of 5 finer than the scale of primary α-phase dendritic solidification simulations. The heat extraction rate is assumed constant, which relates to an applied cooling rate (neglecting release of latent heat) CR = Q ρc p. We note that this cooling rate is different from the cooling rate evaluated from the experiments, due to the difference in definition. We will compare both entities by comparing 4

MCWASP Figure 3. Results of a 3D simulation of primary α-phase nucleation and growth. the solidification time (see below). All simulations of primary solidification are performed in a domain with periodic boundary conditions initially filled with homogeneous melt. The simulations of divorced eutectic growth start from a concentration profile taken around the melt channel of a 2D simulation of primary α-phase growth. Then a second particle distribution for β-phase nucleation is generated using the parameters given in Table 2. Figure 4. 2D simulated microstructures of Mg-Al alloy solidified at different cooling rates. Figure 5. Metallographic images of etched as cast Mg-Al microstructures for different cooling rates. The Mg content is 9 wt.%. The applied cooling rates are (a) furnace controlled cooling, (b) air cooling and (c) submersion in water. The temperature-curves for these experiments are shown in Fig. 1 Simulations of consecutive α- and β-phase nucleation and their divorced eutectic growth were carried out in 2D. The applied cooling rates were varied from 5K/s to 25K/s. The resulting micro-structures are shown in Fig. 4. The solidified microstructure consists of a primary α-phase dendrites (blue) surrounded by the thin channels of β-phase (orange). The volume fraction of the β-phase is about 5% in all simulations. A clear trend of increasing α-phase grain density with increasing cooling rate can be observed which is in agreement with experimental observations (see Fig. 5). This behavior is achieved naturally by the implementation of a free growth model proposed by Greer et al. [13] which allows to reproduce the recoalescence effect responsible for the increased nucleation density at higher cooling rates. The corresponding simulated cooling curves are shown in Fig. 6. 5

900 5K/s 10K/s 15K/s 20K/s 25K/s 850 Temperature [Kelvin] 800 892 750 890 eutectic temperature 700 888 650 886 0 1 2 3 600 0 10 20 30 40 50 60 Time [seconds] Figure 6. Simulated cooling curves. The solidification experiment with water cooling takes approximately 20 seconds from fully liquid to fully solid which matches our simulations with an applied cooling rate of 25 K/s. Thus the simulated microstructure shown in Fig. 4(e) corresponds to the experimentally obtained microstructure shown in Fig. 5(c). Comparison between simulation and experiment shows good agreement in grain density and connectivity. The characteristic grain-size in the simulation is about 150 µm which is in correspondence with the experiment for the same estimated cooling rate. Due to the prohibitive computational expense of slower solidification conditions, the lowest simulated cooling rate was 5 K/s and takes about 110 seconds of simulated time to fully solidify. The experimental results for air or furnace controlled cooling were much slower taking 600 and 1700 seconds respectively. However a clear trend of coarser grains with decreasing cooling rate can be observed. Another important factor is the closed shell β-phase formation which is responsible for increasing the corrosion-resistance of Mg-Al-alloys. Thus good correspondence of our simulation results with experimental observations allows for computationally assisted design of corrosion resistant Mg-alloys. Of course there is a trade off between fine grain structure (better mechanical properties) and low connectivity of primary dendrites (better corrosion properties). The balance between mechanical and corrosion properties can be further tuned and investigated using the proposed method. In order to investigate in detail the β-phase nucleation and growth kinetics and the formation of a divorced eutectic microstructure with the tertiary α-phase nucleation in the residual melt these simulations were performed in 3D. The results of a divorced eutectic solidification simulation are shown in Fig. 7. The simulations starts with a residual melt channel with the eutectic composition of 31% and nearly eutectic α-phase at the bottom and the top of the simulation volume. This system is cooled down starting at 712.0 K, slightly above the eutectic temperature. The α-phase is still slowly growing on this scale. As soon as the undercooling is sufficient, nucleation automatically takes place and generates a large number of β-grains in the residual melt. The high nucleation density of 3 10 13 particles per m 3 was chosen because it is expected for the particles to be aggregated in the residual melt. With an abundant supply of nucleation sites and our chosen cooling rate of 5 K/s we get around 50 nuclei within the channel. A strong tendency of β-phase growing towards and at the α-liquid-interface can be observed. This is because Al is segregated strongly by the growing α-phase into the melt close to the interface, since the solubility for Al in α is only 11.6% at the eutectic temperature of 709.4 K. This leads to a higher concentration 6

MCWASP Figure 7. Zoomed in region of the residual eutectic melt channel in between two α-phase (blue) dendrites at different time-steps. Undercooling causes the β-phase (orange) to nucleate at randomly chosen positions in the melt (a) and grow (b-c) until the α-phase boundary is fully covered (d). Then residual melt is depleted and the α-phase can nucleate and grow again. of Al in the melt close to the slowly growing α-phase. The β-phase, consisting of about 40% Al grows preferably in the area of the increased concentration. Due to this circumstances a rapid, full enclosure of the α-phase by the β-phase can be observed. Thus, further growth of the α-phase is prohibited. Continuing β-phase growth depletes the concentration in the residual melt. It is impossible for the melt at that state to fully solidify as β-phase due to the lack of Al. This leaves the system in a state where a tertiary α-phase nucleation is necessary to fully solidify. Investigations of this behavior will be presented elsewhere. 7. Outlook We have presented 2D simulations of consequent α- and β-phase evolution, shown in Fig. 4. 3D simulations of the same accuracy will require about two orders of magnitude more memory and computational power, which is currently not feasible on a single compute node. Figure 8. 3D benchmark simulation of the eutectic solidification in α-channel shown in Fig. 7. The domain is subdivided into hundreds of blocks shown to the right with the rank-dependent color-coding. Therefore, distributed-memory parallelism based on block-structured domain decomposition has been implemented in OpenPhase, which will be used for further 3D simulations on large grids. Fig. 8 shows a benchmark-simulation of eutectic solidification in α-channels. The computational domain of 300 300 300 grid points is subdivided into hundreds of blocks that are assigned to 128 processors. As the phase-field values can only change in the interface regions, OpenPhase only calculates the updates of the phase-fields in these regions. While this approach drastically 7

Time per step [seconds] 8 4 2 150 x 150 x 150 150 x 150 x 300 150 x 300 x 300 300 x 300 x 300 300 x 300 x 600 8 16 32 64 128 256 512 1024 Processes Figure 9. Benchmark for eutectic solidification in α-channels showing combined strong scaling (solid lines) and weak scaling (dashed lines) results. The colors and symbols represent the different simulated domain sizes. improves the performance, it leads to a very inhomogeneous distribution of the computational load and requires sophisticated load balancing techniques. An over-decomposition of the computational domain allows a graph-partitioner to assign blocks to processors, so that in the worst case the load imbalance is bounded by the computation time on the most expensive block. Therefore, expensive blocks are divided into multiple parts, when the implemented work load model predicts a benefit for the wall clock time for the next time steps. The system is rebalanced at runtime, whenever it is needed. Figure 9 shows performance results on the supercomputer JuRoPA at Forschungszentrum Jülich with up to 1024 processors. Excellent weak scaling is observed: it can be seen that the computation time does not change significantly when the total work load per process is kept constant. This allows the realization of large scale simulations utilizing a parallel cluster without increasing the wall time. For primary solidification perfect weak scaling results were achieved with up to 2048 processes on grids with up to 800 800 1600 grid points. With this new possibilities of large-scale simulations, we hope to set up a full 3D solidification simulation analogous to the 2D calculations shown in Fig. 4. Also, due to an increased resolution due to higher count of grid points, a simulation can be carried out resolving both: the primary as well as the eutectic solidification on their different length-scales. For a realistic setup a grid of about 600 600 600 grid points is needed. Thus the time needed for one timestep is estimated to be about 2 seconds with 4096 processes. A calculation of 25 seconds real time would take approximately 11.5 days.[14] 8. Summary In this paper we presented an approach to simulate the solidification of Mg-Al alloys. A phase diagram based on Thermocalc TCS 4.0 database TCBIN was linearized and adjusted for consistency. The free-growth model was implemented into OpenPhase for realistic nucleation behavior of both α- and β-phases. 3D simulations of the first stage of primary solidification were carried out. A full solidification cycle was simulated in 2D including a primary α- as well as secondary β-phase nucleation and growth in one simulation, with close to 200 K of cooling in between the two events. Resulting cooling curves as well as simulated microstructures were compared to experiments. Remarkable agreement has been obtained for characteristic grain size and β-phase closed shell formation. Due to the insuficient resolution a dual-scale approach was employed to model an interdendritic region of eutectic melt prior to final solidification. Divorced eutectic behavior was reproduced showing a need for tertiary nucleation of α-phase due to full shielding of primary dendrites after a short period of time. The possibility for future highly parallelized computations of large scale 3D domains with high resolution at different scales was 8

implemented and tested. Benchmarks show excellent weak scaling properties. 9. Acknowledgements The authors (AM, OS and IS) would like to thank the Korea Institute of Materials Science (KIMS) for support under the collaboration project Phase-field simulation of Microstructure Evolution in Ni-based Superalloys and Light-weight alloys. IS and MT like to thank the German Research Foundation (DFG) for support under STE 1116/13-1. DH would like to thank the Helmholtz-Association for financial support. References [1] Sun D Y, Mendelev M I, Becker C A, Kudin K, Haxhimali T, Asta M, Hoyt J J, Karma A and Srolovitz D J 2006 Phys. Rev. B 73 024116 [2] Wang M, Xu Y, Zheng Q, Jung T, Wu S and Chawla N 2014 Met. Mat. Trans. A 45 2562 [3] Zhao M C, Liu M, Song G and Atrens A 2008 Corros. Sci. 50 1939 [4] Chen L-Q 2002 Annu. Rev. Mater. Res. 32 113-140 [5] Steinbach I 2009 Mod. Sim. Mater. Sci. Eng. 17 073001-31 [6] Steinbach I 2013 Annu. Rev. Mater. 43 89-107 [7] Steinbach I 2013 JOM 65 1096 [8] Böttger B, Eicken J and Steinbach I 2006 Acta Mater. 54 2697-2704 [9] Böttger B, Eicken J, Ohno M, Klaus G, Fehlbier M., Schmid-Fetzer R, Steinbach I and Bührig-Polaczek A 2006 Adv. Eng. Mater. 8 24147 [10] Elsayed F R, Hort N, Salgado-Ordorica M A and Kainer K 2011 Mater. Sci. Forum 690 65-68 [11] Steinbach I and Pezzolla A 1999 Physica D 134 385-393 [12] Maxwell I and Hellawell A 1975 Acta Metalurgica 23 229-237 [13] Greer A L, Bumm A M, Tronche A,Evans P V and Bristow D J 2000 Acta Mater. 48 2823-2835 [14] Tegeler M, Monas A, and Sutmann G 2015 Proceedings of the Fourth International Conference on Parallel, Distributed, Grid and Cloud Computing for Engineering 107 5 (doi:10.4203/ccp.107.5) 9