Atomistic Determination of Anisotropic Surface ... - ACS Publications

Dec 8, 2017 - Graphic illustration of the slab model for surface energy calculation (a) and typical atomic structures corresponding to the surface ori...
4 downloads 11 Views 7MB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Article Cite This: ACS Omega 2017, 2, 8803−8809

Atomistic Determination of Anisotropic Surface Energy-Associated Growth Patterns of Magnesium Alloy Dendrites Jinglian Du,† Ang Zhang,† Zhipeng Guo,*,† Manhong Yang,† Mei Li,‡ and Shoumei Xiong*,† †

School of Materials Science and Engineering and Laboratory for Advanced Materials Processing Technology, Ministry of Education, Tsinghua University, Beijing 100084, China ‡ Materials Research Department, Research and Innovation Center, Ford Motor Company, MD3182, P.O. Box 2053, Dearborn, Michigan 48121, United States S Supporting Information *

ABSTRACT: Because of the existence of anisotropic surface energy with respect to the hexagonal close-packed (hcp) lattice structure, magnesium alloy dendrite prefers to grow along certain crystallographic directions and exhibits a complex growth pattern. To disclose the underlying mechanism behind the three-dimensional (3-D) growth pattern of magnesium alloy dendrite, an anisotropy function was developed in light of the spherical harmonics and experimental findings. Relevant atomistic simulations based on density functional theory were then performed to determine the anisotropic surface energy along different crystallographic directions, and the corresponding anisotropic strength was quantified via the least-square regression. Results of phase field simulations showed that the proposed anisotropy function could satisfactorily describe the 3-D growth pattern of the α-Mg dendrite observed in the experiments. Our investigations shed great insight into understanding the pattern formation of the hcp magnesium alloy dendrite at an atomic level.



INTRODUCTION Magnesium alloys are extensively applied in the industry because of their light weight, high specific strength, and environmental friendliness.1,2 The α phase or α-Mg dendrite is one of the most important microstructures;3,4 the dendritic size, orientation, and morphological distributions significantly influence the mechanical properties of the magnesium alloys.5−7 Therefore, understanding the dendritic growth pattern, in particular determination of the preferred growth orientation-related three-dimensional (3-D) morphology, is of crucial importance in controlling the solidification microstructure and the associated properties of the alloys. The α-Mg dendrite with an hexagonal close-packed (hcp) lattice structure usually prefers to grow along the highly symmetrical directions,8−10 resulting in a 6-fold symmetry on the basal plane. In this respect, recent experimental investigations have confirmed that the α-Mg dendrite exhibits an 18-primarybranch pattern in 3-D, with 6 along the ⟨1120̅ ⟩ directions on the basal plane and the other 12 along the ⟨112̅x⟩ directions on the nonbasal planes, such as ⟨112̅3⟩ and ⟨224̅5⟩ (i.e., ⟨112̅ 2.5⟩).11−13 During solidification, both thermodynamic and kinetic effects at the solid−liquid interface play crucial roles in determining the microstructure pattern formation.14−20 For a dendritic microstructure, however, the growth orientation and pattern formation are mostly determined by the thermodynamic factors related to the anisotropic surface energy in light of the underlying lattice structure.4,6,8,9 Meanwhile, the atomic © 2017 American Chemical Society

stacking based on the lattice configuration can also be interpreted and understood in terms of the anisotropic surface energy21 and accordingly understanding the 3-D growth pattern formation of the α-Mg dendrite requires a quantitative investigation of the anisotropic surface energy at the atomic level. Relevant studies have revealed that additional elements could also affect the anisotropic surface energy and thus the dendritic growth pattern, leading to complex dendritic microstructures that are not fully understood.22,23 It is well established that dendrite tends to grow along those directions with the maximum surface energy or the minimum surface stiffness.24−26 To quantify this, an explicit surface energy function must be formulated as a prior. For the metallic alloy dendrite with an face-centered-cubic (fcc) lattice structure, such formulation has been well established and extensively employed.5,25 However, this is not the case for the magnesium alloy dendrite with an hcp lattice structure. One of the available models for the hcp α-Mg dendrite was17,27 γ(θ , φ) = γ0[1 + ε20y20 (θ , φ) + ε40y40 (θ , φ) + ε60y60 (θ , φ) + ε66y66 (θ , φ)]

(1)

where γ0 is the average surface energy and ε20, ε40, ε60, and ε66 measure the anisotropic strength along different directions. Received: August 12, 2017 Accepted: November 23, 2017 Published: December 8, 2017 8803

DOI: 10.1021/acsomega.7b01174 ACS Omega 2017, 2, 8803−8809

Article

ACS Omega

Figure 1. Schematic illustration of the spherical and Cartesian coordination systems (a) and the atomic lattice structure of hcp magnesium (b), showing the correlation between the indexes of crystallographic planes and the corresponding orientations.

Equation 1 could perfectly describe the 6-fold symmetry of the α-Mg dendrite at the basal plane, but it could not characterize the dendritic pattern at the nonbasal planes. Consequently, the simulated dendritic growth pattern deviated significantly from that observed in the experiment.12,22 Therefore, it is necessary to develop an anisotropic function to describe the practical pattern formation of the hcp magnesium alloy dendrite. According to the experimental findings on the 3-D growth pattern of the α-Mg dendrite, in this work, we first presented an anisotropy function by combining certain terms of the spherical harmonics. Relevant atomistic simulations based on the hcp lattice structure and density functional theory (DFT)28 were then performed for the first time to quantify the associated anisotropic strength along different directions. The influence of both solute type and solute concentration on the anisotropic parameters was also investigated and discussed.

y53 (θ , φ) =

nx = sin θ cos φ , ny = sin θ sin φ , nz = cos θ

flm ylm (θ , φ)

1 4

y53 (θ , φ) =

1 32

5 ·(3nz2 − 1) π

(7)

770 ·(nx3 − 3nxny2) ·(9nz2 − 1) π

(nx3 − 3nx ·ny 2)2 (9nz2 − (1 + ε3))2 ]

(2)

(8)

γ(θ , φ) = γ0·[1 + α20·y20 2 (θ , φ) + α53·y532 (θ , φ)]

5 ·(3 cos2 θ − 1) π

(9)

where εi’s (i = 1, 2, and 3) are the anisotropic parameters weighing the anisotropic strength along different directions. In particular, ε1 relates to dendritic growth along the direction normal to the basal plane (i.e., ⟨0001⟩), whereas ε2 and ε3 relate to that along both the basal plane and nonbasal plane directions (i.e., ⟨112̅x⟩ and ⟨112̅0⟩). The magnitudes of ε1, ε2, ε3, and γ0 in eq 9 were quantitatively determined via the DFT-based atomistic simulations by constructing a slab model with a periodic boundary condition. The surface unit cell repeats along the crystallographic orientation normal to crystallographic plane

where f lm is the expansion coefficient, ylm(θ, φ) is the spherical harmonics function, and θ ∈ [0, π] and φ ∈ [0, 2π] are the spherical coordinate angles. We employed the y20-square and y53-square terms to describe the 18-primary-branch pattern of the α-Mg dendrite

1 4

y20 (θ , φ) =

γ(θ , φ) = γ0·[1 + ε1·(3nz2 − 1)2 + ε2 ·

l = 0 m =−l

y20 (θ , φ) =

(6)

On substituting eqs 7 and 8 into eq 3, the resultant surface energy formulation is expressed as

l

∑∑

(5)

where nx, ny, and nz are the components of surface normal unit vector n⃗ along the x, y, and z directions, respectively, as shown in Figure 1a. It is worth stressing that the y53-square term could satisfactorily illustrate the 18-primary-branch pattern (including both basal and nonbasal directions) of the magnesium alloy dendrite observed in the experiments.12 Accordingly, eqs 4 and 5 can be written as

RESULTS AND DISCUSSION According to the experimental observations, the α-Mg dendrite exhibits an 18-primary-branch pattern in 3-D.12,13 A typical way of characterizing such a growth pattern is by formulating an anisotropic function on the basis of spherical harmonics17,29 f (θ , φ ) =

385 3iφ 3 ·e ·sin θ· (9 cos2 θ − 1) π

where α20 and α53 are the anisotropic parameters and





−1 32

(3)

(4) 8804

DOI: 10.1021/acsomega.7b01174 ACS Omega 2017, 2, 8803−8809

Article

ACS Omega

Figure 2. Graphic illustration of the slab model for surface energy calculation (a) and typical atomic structures corresponding to the surface orientations ⟨101̅0⟩ (b) and ⟨112̅0⟩ (c) of the Mg−1.56 atom % Al alloy. (Note that the red spheres and the blue spheres denote the aluminum atoms and the magnesium atoms, respectively.)

Table 1. Average Value of Surface Energy γ0 (J/m2) and Anisotropy Parameters ε1, ε2, and ε3 of Mg-Based Alloys, Obtained by Least-Square Regression according to eq 9 and the Resultant Data from DFT-Based Calculations composition

γ0

ε1

ε2

ε3

Mg−1.56 atom % Sn Mg−1.56 atom % Ba Mg−1.56 atom % Ca Mg−1.56 atom % Y Mg−1.56 atom % Al Mg−3.12 atom % Al Mg−6.25 atom % Al Mg−7.0 atom % Al Mg−11.5 atom % Al Mg−2.0 atom % Zn Mg−1.33 atom % Zn Mg−0.67 atom % Zn 100 atom % Mg

0.7919 0.7605 0.7989 0.8232 0.8011 0.8003 0.8067 0.8202 0.8188 0.8094 0.81 0.8108 0.7986 0.1222 0.0899

−0.0552 −0.0457 −0.059 −0.0594 −0.0548 −0.0508 −0.0479 −0.0629 −0.0467 −0.0317 −0.033 −0.0331 −0.0056

0.0741 0.0956 0.0422 0.044 0.0805 0.0812 0.0736 0.0742 0.1344 0.0757 0.0765 0.0715 0.0593

−0.0770 −0.084 −0.5843 −0.3839 0.0594 −0.0345 −0.1688 −0.1046 0.0401 −0.4284 −0.4631 −0.5014 −0.3174

{hkil}.28,30 For the hcp lattice structure,31,32 the indexes of the crystallographic plane and the perpendicular orientation (e.g., ⟨112̅k⟩ ⊥ ⟨112̅x⟩) are correlated via k = 2x(c/a)2/3 (see Figure 1b). As shown in Figure 2a, this model includes a vacuum region and a slab with several atomic layers; the bottom atomic layers in the slab were fixed, whereas the top ones were relaxed during optimization. Accordingly, the surface energy along 12 crystallographic directions related to the preferred growth orientations of the magnesium alloy dendrite, including {0001}, {101̅0}, {101̅1}, and {112̅k} with k ranging from 0 to 8 (see Supporting Table SI), was first determined. These values were then used to calculate the magnitude of ε1, ε2, ε3, and γ0 on the basis of the least-square regression method.8,33 The atomic structure of binary Mg-based alloys was simulated using the solid solution model, where a certain number of solvent atoms was replaced by the solute atoms.34,35 It has been confirmed that the position of additional atoms does not affect significantly the variation trend of the calculated surface energy with respect to the surface orientations.36,37 The concentration of the additional elements for the magnesium alloys was chosen on the basis of their solubility in the solid or the maximum solute concentration in the Mg matrix.36 We optimized the atomic structure of bulk materials and then constructed the

(γ112̅0 − γ101̅0)/2γ0

(γ112̅0 − γ0001)/2γ0

0.1114 0.0888 0.0460 0.0664 0.1237 0.1248 0.1216 0.0790 0.0785 0.0873 0.0849 0.0813 0.0534

0.134 0.1265 0.0962 0.1086 0.1431 0.1429 0.1353 0.1273 0.1619 0.1396 0.1365 0.1335 0.0728

0.18 ± 0.08

1.2 ± 0.7

reference this work this work this work this work this work this work this work this work this work this work this work this work this work ref 17 ref 27

surface slab model with respect to the growth orientations of the α-Mg dendrite. In terms of an n-layer surface slab model with respect to the {hkil} plane, the surface energy can be obtained via30,33 γ (θ , φ ) =

n Eslab − n·E bulk 2S

(10)

Enslab

where is the total energy per primitive slab unit cell, Ebulk is the total energy per primitive bulk unit cell, S is the surface area per primitive slab unit cell, and the factor of 2 accounts for two equivalent surfaces of a particular slab model. Details on the construction of the atomic structure model and surface energy calculations for these magnesium alloys can be found elsewhere.37 Accordingly, the surface anisotropic parameters were deduced from these calculated data of orientationdependent surface energy. The atomistic simulations in this work were performed using the Vienna ab initio simulation package within the DFT framework.28,38 The exchange and correlation interaction was described by local density approximation with projectoraugmented wave potentials.39 After convergence tests, a plane-wave cutoff energy of 450 eV was used for pure Mg and 420 eV for binary Mg-based alloys, including Mg−Al, Mg− 8805

DOI: 10.1021/acsomega.7b01174 ACS Omega 2017, 2, 8803−8809

Article

ACS Omega

Figure 3. Scope and distribution of the determined anisotropic parameters ε1, ε2, and ε3 for (a) Mg−Al alloys and (b) Mg−Zn alloys. Inset maps show the corresponding polar diagram of the anisotropic surface energy.

Figure 4. Schematic illustration for the surface anisotropy function of the (a) Mg−1.56 atom % Al alloy dendrite and (b) Mg−2.0 atom % Zn alloy dendrite. The anisotropic parameters determined via the DFT-based calculations and the least-square regression are (a) ε1 = −0.0547, ε2 = 0.0805, and ε3 = 0.0594 and (b) ε1 = −0.0317, ε2 = 0.0757, and ε3 = −0.4284, respectively.

Ba, Mg−Sn, Mg−Ca, Mg−Y, and Mg−Zn. The k-point separation in the Brillouin zone of the reciprocal space was set as 0.01 Å−1 for each surface unit cell.40 Meanwhile, the total energy was converged to 5 × 10−7 eV/atom with respect to electronic, ionic, and unit cell degrees of freedom. The anisotropic surface energy was obtained from accurate calculations in accordance with our representative slab models. Figure 2b,c shows the typical atomic structures corresponding to the ⟨101̅0⟩ and ⟨112̅0⟩ surface orientations of Mg−1.56 atom % Al alloy, respectively. The calculated surface energy along different orientations is listed in Supporting Table SI. The estimated surface energy of the {0001} basal plane, that is, the most close-packed crystallographic plane, was minimum, which agreed well with the previous reported results.17,27 It is worth mentioning that the method for the anisotropic surface

energy calculation of hcp Mg-based alloys could be employed for other alloy systems as well. However, a specific slab model must be developed to accommodate the influence of the lattice structure for other alloys, for example, the fcc lattice structure for Ni-based alloys. Depending on the DFT-based calculations and the leastsquare regression according to eq 9, the γ0, ε1, ε2, and ε3 were evaluated and the corresponding values are listed in Table 1, together with those from the literature. The results indicated that for all considered alloys, ε1 was always negative, indicating that the α-Mg dendrite preferred not to grow along ⟨0001⟩, which agreed well with the experimental observations.11,12,23 On the contrary, ε2 was positive, signifying that the α-Mg dendrite preferred to grow along both ⟨112̅0⟩ and ⟨112̅x⟩. Moreover, the value of ε3 was either negative or positive, 8806

DOI: 10.1021/acsomega.7b01174 ACS Omega 2017, 2, 8803−8809

Article

ACS Omega

Figure 5. Growth patterns of an α-Mg dendrite simulated using the phase field model on the basis of the quantified anisotropy function in light of relevant DFT-based calculations (a0−a3) and a comparison with those simulated in a previous work (b0−b3).12 (a0) and (b0) show the 3-D dendritic morphology, (a1−a3) and (b1−b3) show the dendritic morphology viewed from the ⟨0001⟩, ⟨101̅0⟩, and ⟨112̅0⟩ directions, and (a4, a5) and (b4, b5) show the patterns by cutting the dendrite along the {0001} and {101̅0} sections, respectively.

indicating that the growth tendency of the α-Mg dendrite along ⟨112̅0⟩ changed according to local conditions. Figure 3a,b shows the scope and distribution of anisotropic parameters ε1, ε2, and ε3 for binary Mg−Al and Mg−Zn alloys, respectively, and the insert maps show the corresponding polar diagrams of anisotropic surface energy. The orientation-dependent surface energy of the other Mg-based alloys, including Mg−Ba, Mg− Sn, Mg−Ca, and Mg−Y, exhibited similar behavior as that of the Mg−Al and Mg−Zn alloys, as shown in Supporting Figure S1. The average value of γ(θ, φ) for these Mg-based alloys was ∼0.8 J/m2 and changed as the additional solute elements. For all considered Mg−Al alloys, ε1 varied from −0.0467 to −0.0629, ε2 from 0.0736 to 0.1344, and ε3 from −0.1688 to 0.0594, as shown in Figure 3a. This result signified that the Mg−Al alloy dendrite tended to grow along ⟨112̅0⟩ in the basal plane and along ⟨112̅x⟩ in the nonbasal planes and the growth tendency along the ⟨112̅x⟩ direction was dependent on the magnitude of Al concentration. For different Mg−Zn alloys, ε1 changed from −0.0331 to −0.033, ε2 from 0.0715 to 0.0765, and ε3 from −0.5014 to −0.4284, as shown in Figure 3b. In comparison to that for the Mg−Al alloy dendrite, the value of ε3 for the Mg−Zn alloy dendrite tended to be more negative, indicating a weaker growth tendency along ⟨112̅0⟩ in the basal plane. The growth tendency along ⟨112̅x⟩ in the nonbasal planes of the Mg−Zn alloy dendrite was comparable to that of

the Mg−Al alloy dendrite. The underlying reason behind such difference can be understood by comparing the atomic interaction between the solvent and solute, as indicated by the enthalpy of mixing, (ΔH).41 The predicted growth tendency of the α-Mg dendrite according to eq 9 in terms of the anisotropic parameters was in agreement with previous experimental findings.13,22 Figure 4a,b shows typical graphic illustrations of the surface anisotropy function of the Mg−1.56 atom % Al alloy and the Mg−2.0 atom % Zn alloy dendrites, respectively. In Figure 4a, ε1 = −0.0547, ε2 = 0.0805, ε3 = 0.0594, and γ0 = 0.8011 J/m2, and in Figure 4b, ε1 = −0.0317, ε2 = 0.0757, ε3 = −0.4284, and γ0 = 0.8094 J/m2, that is, according to the DFT-based calculations and the least-square estimated values shown in Figure 3a,b, respectively. Then, the 3-D growth pattern of the hcp α-Mg dendrite can be quantitatively determined with the evaluated anisotropic parameters. However, the present surface anisotropy function is still limited in terms of evaluating the angle between the ⟨112̅0⟩ and ⟨112̅x⟩ directions, that is, it could never exceed 45°, which is different from some of the values measured in the experiments.12,13 Further investigations are thus required to resolve this uncertainty. Nevertheless, the proposed anisotropy function is still able to capture the 18primary-branch pattern of the hcp α-Mg dendrite observed in the experiments, which has not been achieved by other models. 8807

DOI: 10.1021/acsomega.7b01174 ACS Omega 2017, 2, 8803−8809

Article

ACS Omega The DFT-based atomistic simulations in the present work were performed under equilibrium conditions, which is different from the practical experimental cases. During solidification, the way the atoms are absorbed from the liquid phase to the solid phase is highly dependent on the lattice structure of the formed crystal, and the atoms tend to attach these locations with either the minimum surface stiffness or the maximum surface energy,42−45 both being more concerned with the lattice structure rather than the kinetic behavior at the liquid−solid interface. Accordingly, we believed that the crystallographic lattice structure is the most fundamental factor that determines dendritic pattern formation. Remarkably, even with such difference, the predicted results on the growth pattern of the magnesium alloy dendrite agreed quite well with the experimental findings.36,37 However, more investigation, in particular relevant atomistic simulations at elevated temperatures, should be performed to explore the exact influence of temperature on the dendritic growth pattern. On the basis of the quantified anisotropic parameters, the growth pattern of the α-Mg dendrite was simulated by incorporating eq 9 into the phase field model. Figure 5 shows the simulated morphology of a Mg−Sn alloy dendrite and a comparison with those reported in ref 12 where the anisotropic parameters were guessed values. The results confirmed that the proposed anisotropy function could reasonably describe the 3-D growth pattern of the α-Mg dendrite. It is noted that these guessed anisotropic parameters in ref 12 were significantly larger than the ones determined on the basis of DFT calculations, and such difference could generate artificial effects on the predictions of the 3-D dendritic morphology. Besides the anisotropy function, meaningful simulations of the hcp α-Mg dendrite also require a thorough investigation of other important parameters, including partition coefficient, supercooling, release of heat of fusion, and so on. Relevant studies along these directions are still in progress, and the detailed results will be reported in a forthcoming paper.



CONCLUSIONS



ASSOCIATED CONTENT



Orientation-dependent surface energy of binary Mgbased alloys with different additional elements, including Mg−Al, Mg−Ba, Mg−Sn, Mg−Ca, Mg−Y, and Mg−Zn (PDF)

AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected] (Z.G.). *E-mail: [email protected] (S.X.). ORCID

Zhipeng Guo: 0000-0003-2899-0870 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was financially supported by the National Key Research and Development Program of China (2016YFB0301001), the National Natural Science Foundation of China (51701104), the Postdoctoral Science Foundation of China (2017M610884), the Tsinghua University Initiative Scientific Research Program (20151080370), and the U.K. Royal Society Newton International Fellowship Scheme. The authors acknowledge the support from Shanghai Synchrotron Radiation Facility for the provision of beam time and the National Laboratory for Information Science and Technology in Tsinghua University for access to supercomputing facilities.



REFERENCES

(1) Sandlöbes, S.; Pei, Z.; Friák, M.; Zhu, L. F.; Wang, F.; Zaefferer, S.; et al. Ductility improvement of Mg alloys by solid solution: Ab initio modeling, synthesis and mechanical properties. Acta Mater. 2014, 70, 92−104. (2) Hidalgo-Manrique, P.; Robson, J. D.; Pérez-Prado, M. T. Precipitation strengthening and reversed yield stress asymmetry in Mg alloys containing rare-earth elements: A quantitative study. Acta Mater. 2017, 124, 456−467. (3) Eiken, J. Dendritic growth texture evolution in Mg-based alloys investigated by phase-field simulation. Int. J. Cast Met. Res. 2009, 22, 86−89. (4) Wang, M. Y.; Xu, Y. J.; Jing, T.; Peng, G. Y.; Fu, Y. N.; Chawla, N. Growth orientations and morphologies of α-Mg dendrites in Mg−Zn alloys. Scr. Mater. 2012, 67, 629−632. (5) Haxhimali, T.; Karma, A.; Gonzales, F.; Rappaz, M. Orientation selection in dendritic evolution. Nat. Mater. 2006, 5, 660−664. (6) Gránásy, L.; Pusztai, T.; Börzsönyi, T.; Warren, J. A.; Douglas, J. F. A general mechanism of polycrystalline growth. Nat. Mater. 2004, 3, 645−650. (7) Danilov, D.; Nestler, B. Dendritic to globular morphology transition in ternary alloy solidification. Phys. Rev. Lett. 2004, 93, No. 215501. (8) Becker, C. A.; Olmsted, D.; Asta, M.; Hoyt, J. J.; Foiles, S. M. Atomistic underpinnings for orientation selection in alloy dendritic growth. Phys. Rev. Lett. 2007, 98, No. 125701. (9) Zhang, L.; Liu, X.; Geng, C.; Fang, H.; Lian, Z.; Wang, X.; et al. Hexagonal crown-capped zinc oxide micro rods: hydrothermal growth and formation mechanism. Inorg. Chem. 2013, 52, 10167−10175. (10) Xia, Z. G.; Sun, D. Y.; Asta, M.; Hoyt, J. J. Molecular dynamics calculations of the crystal-melt interfacial mobility for hexagonal closepacked Mg. Phys. Rev. B. 2007, 75, No. 012103. (11) Pettersen, K.; Lohne, O.; Ryum, N. Dendritic Solidification of Magnesium Alloy AZ91. Metall. Trans. A 1990, 21, 221−230. (12) Yang, M.; Xiong, S.; Guo, Z. Characterisation of the 3-D dendrite morphology of magnesium alloys using synchrotron X-ray tomography and 3-D phase-field modelling. Acta Mater. 2015, 92, 8− 17.

In summary, an anisotropy function based on the spherical harmonics and experimental findings was proposed to describe the 3-D growth pattern of the hcp magnesium alloy dendrite. Relevant atomistic simulations in light of density functional theory and the hcp lattice structure were performed to quantify the surface energy along different growth directions. Accordingly, the anisotropic parameters in the surface energy formulation and the average values of surface energy (∼0.8 J/ m2) for different Mg-based alloys were predicted via the leastsquare regression method. The results showed that the proposed surface anisotropy function in terms of the anisotropic parameters could satisfactorily describe the practical 3-D morphology of the magnesium alloy dendrite observed in the experiments. The present work offers an important insight for a better understanding on the 3-D growth pattern of the hcp magnesium alloy dendrite and thus provides a theoretical guidance for further exploring the microstructure formation of those metallic alloy dendrites with a hexagonal symmetrical structure.

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.7b01174. 8808

DOI: 10.1021/acsomega.7b01174 ACS Omega 2017, 2, 8803−8809

Article

ACS Omega

(36) Du, J.; Guo, Z.; Zhang, A.; Yang, M.; Li, M.; Xiong, S. Correlation between crystallographic anisotropy and dendritic orientation selection of binary magnesium alloys. Sci. Rep. 2017, 7, No. 13600. (37) Du, J.; Guo, Z.; Yang, M.; Xiong, S. Growth pattern and orientation selection of magnesium alloy dendrite: From 3-D experimental characterization to theoretical atomistic simulation. Mater. Today Commun. 2017, 13, 155−162. (38) Kresse, G.; Marsman, M.; Furthüller, J. VASP the Guide. http:// cmsmpiUnivieacat/vasp/. (39) Blöchl, P. E. Projector augmented-wave method. Phys. Rev. B. 1994, 50, 17953−17979. (40) Monkhorst, H. J.; Pack, J. D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188−5192. (41) Du, J.; Wen, B. Composition-structure-property correlations of complex metallic alloys described by the “cluster-plus-glue-atom” model. Appl. Mater. Today 2017, 7, 13−46. (42) Glicksman, M. E.; Singh, N. B. Effects of crystal-melt interfacial energy anisotropy on dendritic morphology and growth kinetics. J. Cryst. Growth 1989, 98, 277−284. (43) Jaafar, M. A.; Rousse, D. R.; Gibout, S.; Bedecarrats, J. P. A review of dendrite growth during solidification: Mathematical modeling and numerical simulations. Renewable Sustainable Energy Rev. 2017, 74, 1064−1079. (44) Krumwiede, T. R.; Schulze, T. P. Dendrite growth shapes in kinetic Monte Carlo models. Modell. Simul. Mater. Sci. Eng. 2017, 25, No. 025009. (45) Hoyt, J. J.; Asta, M. Atomistic computation of liquid diffusivity, solid-liquid interfacial free energy, and kinetic coefficient in Au and Ag. Phys. Rev. B 2002, 65, No. 214106.

(13) Yang, M.; Xiong, S.; Guo, Z. Effect of different solute additions on dendrite morphology and orientation selection in cast binary magnesium alloys. Acta Mater. 2016, 112, 261−272. (14) Takaki, T. Phase-field Modeling and Simulations of Dendrite Growth. ISIJ Int. 2014, 54, 437−444. (15) Eustathopoulos, N. Energetics of Solid/Liquid Interfaces of Metals and Alloys. Int. Met. Rev. 1983, 28, 189−210. (16) Van der Planken, J.; Deruyttere, A. A scanning electron microscope study of vapour grown magnesium. J. Cryst. Growth 1971, 11, 273−279. (17) Asadi, E.; Asle Zaeem, M. The anisotropy of hexagonal closepacked and liquid interface free energy using molecular dynamics simulations based on modified embedded-atom method. Acta Mater. 2016, 107, 337−344. (18) Tang, S.; Wang, J. C.; Li, J. J.; Wang, Z. J.; Guo, Y. L.; Guo, C.; Zhou, Y. H. Phase-field crystal investigation of the morphology of a steady-state dendrite tip on the atomic scale. Phys. Rev. E 2017, 95, No. 062803. (19) Sun, D. K.; Zhu, M. F.; Wang, J.; Sun, D. B. Lattice Boltzmann modeling of bubble formation and dendritic growth in solidification of binary alloys. Int. J. Heat Mass Transfer 2016, 94, 474−487. (20) Lee, D. N. Orientations of dendrite growth during solidification. Met. Mater. Int. 2017, 23, 320−325. (21) Luo, Y.; Qin, R. Surface energy and its anisotropy of hexagonal close-packed metals. Surf. Sci. 2014, 630, 195−201. (22) Guo, E.; Phillion, A. B.; Cai, B.; Shuai, S.; Kazantsev, D.; Jing, T.; et al. Dendritic evolution during coarsening of Mg-Zn alloys via 4D synchrotron tomography. Acta Mater. 2017, 123, 373−382. (23) Casari, D.; Mirihanage, W. U.; Falch, K. V.; Ringdalen, I. G.; Friis, J.; Schmid-Fetzer, R.; et al. α-Mg primary phase formation and dendritic morphology transition in solidification of a Mg-Nd-Gd-ZnZr casting alloy. Acta Mater. 2016, 116, 177−187. (24) Morris, J. R.; Song, X. The anisotropic free energy of the Lennard-Jones crystal-melt interface. J. Chem. Phys. 2003, 119, 3920− 3925. (25) Hoyt, J. J.; Asta, M.; Karma, A. Method for computing the anisotropy of the solid-liquid interfacial free energy. Phys. Rev. Lett. 2001, 86, 5530−5533. (26) Du, D.; Zhang, H.; Srolovitz, D. Properties and determination of the interface stiffness. Acta Mater. 2007, 55, 467−471. (27) Sun, D. Y.; Mendelev, M. I.; Becker, C. A.; Kudin, K.; Haxhimali, T.; Asta, M.; et al. Crystal-melt interfacial free energies in hcp metals: A molecular dynamics study of Mg. Phys. Rev. B. 2006, 73, No. 024116. (28) Payne, M. C.; Teter, M. P.; Allan, D. C.; Arias, T. A.; Joannopoulos, J. D. Iterative minimization techniques for ab initio total-energy calculations: molecular dynamics and conjugate gradients. Rev. Mod. Phys. 1992, 64, 1045−1097. (29) Kara, M.; Kurki-suonio, K. Symmetrized Multipole Analysis of Orientational Distributions. Acta Crystallogr., Sect. A: Cryst. Phys., Diffr., Theor. Gen. Crystallogr. 1981, 37, 201−210. (30) Du, J.; Wen, B.; Melnik, R. Mechanism of hydrogen production via water splitting on 3C-SiC’s different surfaces: A first-principles study. Comput. Mater. Sci. 2014, 95, 451−455. (31) Du, J.; Guo, Z.; Yang, M.; Xiong, S. Multiscale Simulation of αMg Dendrite Growth via 3D Phase Field Modeling and Ab Initio First Principle Calculations, Proceedings of the Fourth World Congress on Integrated Computational Materials Engineering, Ypsilanti; Mason, P., Ed., 2017; pp 263−272. (32) Miller, W. A.; Chadwick, G. A. The Equilibrium Shapes of Small Liquid Droplets in Solid-Liquid Phase Mixtures: Metallic h.c.p. and Metalloid Systems. Proc. R. Soc. A 1969, 312, 257−276. (33) Fitzgibbon, A.; Pilu, M.; Fisher, R. B. Direct Least Square Fitting of Ellipses. IEEE Trans. Pattern Anal. Mach. Intell. 1999, 21, 476−480. (34) Liu, P. L.; Siao, Y. J. Ab initio study on preferred growth of ZnO. Scr. Mater. 2011, 64, 483−485. (35) Mikkelsen, J. C.; Boyce, J. B. Atomic-Scale Structure of Random Solid Solutions: Extended X-Ray-Absorption Fine-Structure Study of Ga1−xInxAs. Phys. Rev. Lett. 1982, 49, 1412−1415. 8809

DOI: 10.1021/acsomega.7b01174 ACS Omega 2017, 2, 8803−8809