You are on page 1of 13

Computational Materials Science 184 (2020) 109883

Contents lists available at ScienceDirect

Computational Materials Science


journal homepage: www.elsevier.com/locate/commatsci

Reliability of molecular dynamics interatomic potentials for modeling of T


titanium in additive manufacturing processes

S. Alireza Etesamia, Mohamed Laradjib, Ebrahim Asadia,
a
Department of Mechanical Engineering, The University of Memphis, Memphis, TN 38152, USA
b
Department of Physics and Materials Science, The University of Memphis, Memphis, TN 38152, USA

A R T I C LE I N FO A B S T R A C T

Keywords: We present an approach to study the transferability of molecular dynamics (MD) interatomic potentials (IPs) for
Direct metal laser sintering materials modeling during metal additive manufacturing (MAM) processes, which involves rapid and cyclic
Molecular dynamics melting, solidification, vaporization, and solid phase transformations. We apply the approach to the ten available
Interatomic potentials IPs of titanium (Ti) as an example, which resembles the MAM process modeling of commercially pure Ti and Ti-
Metals
base alloys. We consider the capability to produce solid phase change behavior of the alloy in equilibrium and
nonequilibrium thermodynamics as the primary required characteristics of an IP for the MAM process modeling.
Two-phase α-β, β-liquid, and liquid-vapor coexistence properties are used as the secondary criteria. Finally, we
use other relevant single-phase properties as the tertiary criteria, such as the elastic constants at high tem-
peratures, liquid structure factor, self-diffusivity, and shear viscosity. We show there are only three IPs that
demonstrate reasonable characteristics for MAM process modeling.

Metal additive manufacturing (MAM) is an advanced manufacturing behavior, including solid-state phase transformations.
technique that has been increasingly adopted in many industrial sec- MD simulations can be employed as a tool to elucidate the physical
tors, including the biodevice, aerospace, and automobile industries, to processes at the microscopic level, which are involved during the MAM,
manufacture consolidated, complex, and customized parts [1,2]. Most thereby reducing the amount of trial and error experiments that are
MAM techniques are based on the layer-by-layer scanning of the part needed to determine the process variables [10]. However, most MD
geometry using a localized heat source, such as a laser or an electron simulations of additive manufacturing processes have been carried out
beam [3]. Among the major obstacles to the widespread utilization of on nonmetallic materials. For instance, Qiu et al. [11] used MD simu-
MAM is the lack of understanding of the material behavior during the lations to investigate the melting behavior of nanosized α–Al2O3 and
MAM processes [4]. Recent advances in computational materials sci- used the results to appropriately select the wavelength of the laser or
ence, such as integrated computational materials engineering (ICME), electron beam in the process. In another study, Wang et al. [12] em-
can be leveraged to understand the materials behavior during MAM at ployed MD simulations to study the binder spraying volume and its
the microscopic level [5–7]. Molecular dynamics (MD) using empirical mechanics in bone scaffolds during a three-dimensional (3D) printing
and semiempirical interatomic potentials (IPs) is among those fre- process. Qin et al. [13] used MD simulations to investigate the me-
quently and efficiently used computational techniques within the ICME chanics of porous graphene-based materials used in 3D printing. For
concept. Today’s computer power easily allows the MD simulation of metallic one, Zhao et al. [14] simulated the MAM of amorphous Cu
large systems with up to millions of atoms and, in some instances, up to pillars with a particular temperature controlling procedure and then
billions of atoms [8,9]. Hybrid techniques combining MD with Monte they studied the stress induced transformation during the tensile test.
Carlo methods further extend the simulation time scales, thereby al- MD IPs are typically developed based on accurate predictions of 0 K
lowing simulations of systems relevant to MAM processes. The aim of and low temperature properties, such as the cohesive energy, lattice
the present study is to propose an approach for gauging the reliability of parameters and elastic constants. Such an approach is commonly used
IPs for MD modeling of materials during MAM processes and a roadmap to develop the embedded-atom method (EAM) [15] and modified em-
for the development/verification of novel IPs for this purpose. As an bedded-atom method (MEAM) IPs [16,17], which are commonly used
application of our approach, we choose titanium (Ti), which is a widely IPs for metals. In addition to the 0 K and low temperature properties,
used metallic element, to demonstrate the generality in the material’s one or more key high temperature properties are usually added to the


Corresponding author.
E-mail address: easadi@memphis.edu (E. Asadi).

https://doi.org/10.1016/j.commatsci.2020.109883
Received 9 December 2019; Received in revised form 21 April 2020; Accepted 13 June 2020
0927-0256/ © 2020 Elsevier B.V. All rights reserved.
S.A. Etesami, et al. Computational Materials Science 184 (2020) 109883

optimization procedure to increase the reliability of the IPs for high-


temperature simulations. Such properties include the melting point
(TM) [18,19], liquid structure factor [20–23], high temperature elastic
constants [24–26], and liquid self-diffusion coefficients [27]. However,
there are no IPs in the literature that are either fully verified or de-
veloped for materials modeling during the MAM process. This im-
portant consideration is mandated due to the unique thermal history of
the material during the MAM process, which is briefly reviewed next.
Powder bed fusion (PBF) and directed energy deposition (DED) are
the two MAM techniques that are most commonly used for the fabri-
cation and repair of 3D components, respectively [28]. In PBF, the
moving heat source creates a moving melt-pool by selectively scanning
a thin layer of powders (powder bed), while the metal feedstock, in the
form of powder or wire, is added to the moving heat source spot in the
DED. Therefore, MAM can be considered a “microwelding” material
joining process that lays single tracks of weldments adjacent to each
other to scan the component area within the layers and repeats this
process layer-by-layer to fabricate the 3D component [29]. The scan
strategy is the path of single tracks to scan the component area within
the layers. The stripe scan method is one of the most commonly used Fig. 2. Schematic of the thermal history of alloys during MAM: the build
platform temperature (TB), the melting point (TM), and the solid-state phase
scan strategies in PBF. Fig. 1 depicts the simplified version of the stripe
transformation temperature (TS).
scan strategy to provide illustrations for the following MAM material
thermal history discussion.
Although quantifying the thermal history of alloys during a complex transformation at TS depending on the thermodynamics of the alloy.
MAM process, as illustrated in Fig. 1, is a very challenging task both Therefore, the MD IP should also account for two-phase coexistence
numerically and experimentally, due to the multiphysics nature of the properties, such as TM, α → β transition temperature (Tα→β), energy
process [30–32], there have been significant advances in recent years differences and volume expansion coefficients at TM and Tα→β.
[29,33–36]. The real-time thermography of the PBF process at the 3) Cyclic heating to descending temperatures smaller than Tmax and
National Institute of Standard and Technology (NIST) [37] confirms the cooling due to the heat conduction and overlapping of single tracks
qualitative expected thermal history (schematically shown in Fig. 2 for and repeating of the whole cycle due to scanning of the adjacent
the stripe scan strategy) and validates several successful numerical stripes and the next layers. While the cooling rates for the first track
predictions of the thermal history [38–40]. In summary, the thermal vary in the order of 104 C°/s and 106 C°/s depending on the selected
history of materials in the MAM, as depicted in Fig. 2, involves laser power and scanning speed [41,42], the cooling rates descend
and the bounding temperatures vary for each subsequent cycle of
1) Heating from TB to a Tmax higher than TM of the alloy to melt the the thermal history. Therefore, the time-dependent properties of
metal feedstock and form the melt-pool with the possibility of va- solidification and solid-state phase transformation, such as tem-
porization. Therefore, the MD IP for the MAM should consider perature-time-transformation (TTT) diagram, must be considered as
temperature-dependent liquid and solid properties, such as the li- well for the IP development/verification for the MAM process.
quid structure factor, solid and liquid densities and enthalpies, li-
quid self-diffusion, and liquid shear viscosity, and liquid–vapor Clearly, the MD IP for MAM process modeling must exhibit various
properties, such as the liquid–vapor interfacial tension. solid states as well as liquid and vapor forms of the material as a pre-
2) Cooling down until the adjacent welding track reaches the spot, condition to the above requirements. Developing classical IPs, such as
typically resulting in solidification given the proper selection of EAM and MEAM, to exhibit all the above listed characteristics is very
scanning parameters, such as the laser power and scanning speed; challenging, especially for alloy systems. Therefore, these character-
furthermore, it is possible to observe a solid-state phase istics must be prioritized, and a roadmap or approach should be

Fig. 1. Stripe scan strategy illustration in the PBF-MAM process: the stripe width (s), hatch spacing (h), stripe gap (g), and layer-to-layer rotation of stripes (θ).

2
S.A. Etesami, et al. Computational Materials Science 184 (2020) 109883

developed that considers the targeted phenomena to study. Providing ¯


F (N , V , T ) = ⎡ ⎛ hω ⎞
the first attempt to design such a roadmap for MD IP development/ ⎢3NkB T0 ln kB T0 + ΔFirreversible
⎜ ⎟

⎣ ⎝ ⎠
verification for MAM process modeling is the main goal of the present
study. Furthermore, such a roadmap can be implemented in the de- N 2πkB T0 ⎞1.5⎤ ⎤ T
+ kB T0 ln ⎡ ⎛
velopment of recently emerged machine learning IPs [43–47] to cap- ⎢ V ⎝ Nmω2 ⎠ ⎥ ⎥ T0
⎣ ⎦⎦
ture all of the characteristics for MAM process modeling. 1 1 T
In the present study, which is based on a large set of systematic MD − T ⎡a⎛ − ⎞ + b⎛ln ⎞ + c (T − T0) ⎤.
⎜ ⎟ ⎜ ⎟
⎢ ⎝ T0 T⎠ ⎝ T0 ⎠ ⎥ (2)
simulations, we first determine the solid-state phase change behavior of ⎣ ⎦
the available ten IPs for Ti, which can be found at the NIST IP
where N is number of atoms, V is the volume, T is the temperature, T0 is
Repository (www.ctcms.nist.gov/potentials), from the free energies of
an arbitrary reference temperature, kB is the Boltzmann constant, h̄ is
the body-centered-cubic (bcc), face-centered-cubic (fcc), and hex-
Planck’s constant and ω is the frequency of the oscillations of the re-
agonal-close-packed (hcp) phases as a function of temperature. Second,
ference ideal Einstein model. Parameters a, b and c are determined from
we determine the TTT diagrams using these IPs to determine their
a fit of the enthalpy with a quadratic order polynomial of temperature,
predictive capability of the solid-state phase transformation under
i.e., h (T ) = a + bT + cT 2. The parameter ΔFirreversible is calculated using
nonequilibrium thermodynamics conditions. Third, we assess these IP
a hybrid Hamiltonian that allows for gradually switching between the
predictions of the two-phase coexistence properties, including the α-β,
system Hamiltonian from the used IP’s Hamiltonian and the ideal Ein-
β-liquid, and liquid-vapor properties. These two-phase coexistence
stein crystal Hamiltonian [59]. Here, we chose T0 = 900 K. The en-
properties correspond to the melting point (TM), α → β transition
thalpy polynomial parameters a, b and c are determined from MD si-
temperature (Tα→β), energy differences and volume expansion coeffi-
mulations using the Nosé-Hoover chain thermostat and the Parrinello-
cients at TM and Tα→β, and the liquid-vapor surface tension. Fourth, we
Rahman barostat [60] at zero pressure and at temperatures ranging
determine the IP predictions of various single-phase structural proper-
between 900 and 1200 K at intervals of 25 K. In each simulation, the
ties. These predictions correspond to the elastic constants of the solid
system is run for 0.15 ns for relaxation followed by 0.015 ns for sta-
phase in the temperature range [Tα→β, TM], the structure factor of the
tistical averaging. The lattice constant is determined at T0 from MD
liquid phase, the temperature dependence of the liquid and solid den-
simulations of the used IPs in the isothermal–isobaric (NPT) ensemble.
sities and enthalpies, the liquid self-diffusivity, and the liquid shear
The hybrid switching (EAM or MEAM)-Einstein crystal simulations are
viscosity. Comparison between the numerically obtained results from
performed in the canonical (NVT) ensemble. The number of atoms in
the available IPs with their available experimental counterparts are
the switching simulations is 23 328, 31 250, and 8 788 for the hcp, bcc,
then used to discuss the transferability of these IPs, their capabilities
and fcc phases, respectively. The systems are first equilibrated for 0.1 ns
and limitations for modeling the MAM processes of Ti and its alloys.
before switching, and the duration of each switching simulation is 2 ns.
To numerically obtain the TTT diagram for an IP, a system of atoms
is first equilibrated in the liquid state at 1.1TM. Then, the system is
1. Methods
thermally quenched to various temperatures below the TM, ranging
between 200 and 1250 K, where the equilibrium state is either the α or
The present study includes the investigation of six EAM IPs for Ti,
β phase. We extracted TTT diagrams from MD simulations in the NPT
corresponding to Ti1.eam.fs [22], Ti2.eam.fs [22], Ti3.eam.fs [22],
ensemble at 0 pressure for 15 ns of systems consisting of 5 324 atoms.
Ti.set [48], Ti-v2.eam.fs [49], and Zope-Ti-Al-2003.eam.alloy [50].
The time dependence of the internal energy, following a quench, is used
According to the general EAM formalism, the total energy of a unary
to monitor the kinetics of crystallization to obtain the crystallization
system (Etot ) is defined by [15]
onset time. The percentage of the hcp, bcc, fcc and other structures are
characterized using the polyhedral template matching method [61] at
⎡ 1 ⎤ the end of each run.
Etot = ∑ ⎢F (ρ¯i ) + 2
∑ ∅ (rij )⎥
i ⎣ j (≠ i) ⎦ (1) Tα→β are determined from the free energy calculations in Section
2.1. TM of Ti from the β phase is determined using the solid-liquid
where F (ρ¯i ) and ∅ (rij ) are the embedding energy and pair interaction, coexistence method described in reference [19]. In this method, a slab
respectively [15]. The embedding energy is the required energy to add with a bcc structure coexists at some initial guess of the temperature TM
one atom into the background electron density. The total energy of a with a liquid slab and the system is then allowed to relax while mon-
specific system can be calculated based on fitting EAM parameters to itoring the total energy of the system. The guess of TM is repeatad until
the experimental and/or ab initio calculations. no change in the total energy of the system is observed as a function of
In addition, we investigate all the available MEAM IPs for Ti, cor- time. In other words, until the number of atoms in the β and liquid
responding to NiTi.meam [51], Ti.Kim.meam [52], Ti.Dickel.meam phases are constant. ΔHα→β and ΔVα→β are determined by performing
[53], and Ti.meam.spline [54]. MEAM is a semiempirical IP formula- NPT ensemble simulations in the α phase (with 8 788 atoms) and the β
tion similar to the EAM formulation, with the additional contribution of phase (with 4 394 atoms) at Tα→β of the IP. In each simulation, the
angular momentum into the electron density. system is equilibrated for 1.5 ns, followed by 0.75 ns for data collection.
All the MD simulations in this study are carried out with the large- Likewise, the calculations of the volume expansion (ΔVβ→M) and latent
scale atomic/molecular massively parallel simulator (LAMMPS) mole- heat (ΔHβ→M) are carried out through MD simulations in the NPT en-
cular dynamics simulator [55]. Post-processing analysis and visualiza- semble on systems composed of 6 750 atoms at TM of the IP. The density
tion were performed with the open visualization tool OVITO [56,57]. and excess enthalpy as a function of temperature are determined from
The Helmholtz free energy is calculated using the Gibbs-Duhem similar MD simulations in the NPT ensembles at various temperatures
thermodynamic integration method (TI) [58], approximating the en- in intervals of 25 K.
thalpy with a quadratic polynomial of temperature, The liquid–vapor interfacial tension is calculated through the si-
mulation of coexisting liquid and vapor phases of Ti in the NVT en-
semble. In this method, a system of 3 840 atoms is first equilibrated in
the NPT ensemble for 0.4 ns. The simulation box is then extended by a
factor of 2.5 along the z-axis, with the additional volume corresponding
to the vapor phase. The system is then equilibrated in the NVT en-
semble for 2 ns. The diagonal components of the pressure tensor, Pxx,

3
S.A. Etesami, et al. Computational Materials Science 184 (2020) 109883

Pyy and Pzz, are then calculated over an additional 1 ns using the Irving-
Kirkwood formalism [62]. The liquid–vapor interfacial tension is cal-
culated according to
Lz
γ= [〈Pzz 〉 − 0.5(〈Pxx 〉 + 〈Pyy 〉)],
2 (3)

where Lz is the length of the simulation box in the z-direction.


The self-diffusion coefficient, D, of Ti in the liquid phase is calcu-
lated using the Einstein relation,
1 d
D= lim 〈Δr (t )2〉,
6 t →∞ dt
where 〈Δr (t )2〉 is the mean-square displacement of the atoms at time t
and the average is over initial time t0 at equilibrium. The systems of
atoms are first equilibrated for 0.5 ns in the NPT ensemble. The systems
are further equilibrated for another 1 ns in the NVT ensemble. Then,
additional simulations in the NVT ensemble over 9 ns are used for the
extraction of D. All simulations to calculate D are performed on three
system size consisting of 7 200, 3 840 and 1 584 atoms, and negligible
size effects were observed.
The shear viscosity of the liquid phase is calculated using the Green-
Kubo relation [63,64]
V ∞
η=
kB T
∫0 〈Pxy (0) Pxy (t ) 〉dt ,
(5)

where Pxy (t) is the off-diagonal component of the stress tensor at time t.
In Eq. (5), the average is over the initial times at equilibrium.

2. Results

2.1. Solid-State α → β phase transformation

The thermodynamic stability of the hcp, bcc and fcc phases is de-
termined from their Gibbs free energies as calculated using each of the
ten EAM and MEAM IPs in the MD simulations. Fig. 3(a and b) depicts
Fig. 3. Calculated Gibbs free energy differences for all the considered Ti IPs
the calculated Gibbs free energy differences (in eV/atom) between the
along with the SGTE unary database [65]: (a) bcc and hcp phases and (b) fcc
bcc and hcp phases and fcc and hcp phases, respectively. The data for and hcp phases.
each IP is reported until the melting point of the hcp phase, which is
calculated based on the solid–liquid coexisting method [19]. The ver-
tical and horizontal gray dashed lines in Fig. 3(a and b) mark the ex- 2.3. Two-Phase α-β, β-Liquid, and Liquid-Vapor coexistence characteristics
perimental Tα→β for Ti (1155 K) and Gibbs free energy difference cor-
responding to a phase transition (0 eV/atom), respectively. The black MD calculations of the α-β transition temperature (Tα→β), melting
dashed curves in Fig. 3(a and b) are the Calphad data based on the point (TM,) change in enthalpy (ΔHα→β), change in volume (ΔVα→β),
Scientific Group Thermodata Europe (SGTE) unary database [65]. Since latent heat (ΔHβ→M), and expansion in melting (ΔVβ→M) using all six IPs
stabilizing the fcc phase for the Ti2.eam.fs and Ti3.eam.fs IPs between are listed in Table 1. The calculated liquid–vapor interfacial tensions
900 K and 1200 K was not possible, we only present the free energy using all six IPs over a range of temperatures relevant to MAM are
difference between the hcp and bcc phases for these two potentials. depicted in Fig. 7.
Additionally, the bcc phase is unstable in the same range of tempera-
tures for Ti-v2.eam.fs and is unstable for T < 1100 K for Ti.me-
am.spline; thus, Fig. 3 excludes the corresponding free energy differ- 2.4. High-temperature elastic constants of the solid phases
ences.
We recently calculated the elastic constants using the fluctuation
technique from MD simulations of the α phase of Ti in the range
2.2. Temperature-Time-Transformation diagram
100–1150 K [26]. Here, we use the same method to calculate the elastic
constants of the α and β phases using all six IPs considered for tem-
Fig. 4 depicts the TTT diagrams for each of the six IPs considered
peratures ranging from 900 K until the TM. The MD calculations for the
hereafter that demonstrated the α → β solid-state transition. Three
five independent elastic constants, C11, C12, C13, C33 and C44, of the α
independent runs are performed for each temperature since the initial
phase and the three independent elastic constants, C11, C12, and C44, of
velocities and positions of individual atoms have some influence on the
the β phase are shown in Fig. 8(a–e). Also shown in the figure are the
obtained TTT diagrams. The structure of the crystallized solid following
experimental results for the α phase by Fisher et al. [66] and by Ogi
the quench at each temperature and for each independent run is
et al. [67] and for the β phase by Ogi et al. [67]. We note that the results
characterized next. Fig. 5 shows the percentages of the different solid
from Fisher et al. for C11 C13, C33 and C44 (experiment 1) were obtained
phases versus temperature for each of the six IPs. The standard devia-
using pulse-echo measurements up to 1156, 923, 1083 and 1156 K,
tion of the percentages is determined using the results obtained from
respectively. The experimental data in Ref. [66] near the Tα→β were
the independent runs at each temperature. The final snapshot of the
obtained from extrapolation, and the researcher’s accuracy was ques-
solidified structure at low and high temperatures for different potentials
tioned by Ogi et al. [67].
is shown in Fig. 6.

4
S.A. Etesami, et al. Computational Materials Science 184 (2020) 109883

Fig. 4. MD-calculated TTT diagrams using various Ti IPs.

2.5. Density and enthalpy variations by temperature 2.6. Liquid phase characteristics

The MD calculations of density and excess enthalpy (ΔH = H(T)-H Fig. 10(a) depicts MD calculations of the liquid structure factor, S
(2 9 9)) of the α and β solid phases are shown in Fig. 9(a) and (b), (k), using all six IPs along with experiments, using the procedure ex-
respectively. Fig. 9(c) and (d) depicts the MD calculations of the density plained in our previous publication [70]. Fig. 10(b) depicts the self-
and excess enthalpy (ΔH = H(T)-H(2 9 3)) of the liquid phase from diffusion coefficient in the liquid phase as a function of temperature for
1800 to 2250 K, respectively, along with the two experimental values all six IPs, along with available experimental data [27]. Fig. 10(c) de-
[68,69]. picts the calculated shear viscosity along with its experimental coun-
terparts [68]. It is worth mentioning that D and η are interrelated

5
S.A. Etesami, et al. Computational Materials Science 184 (2020) 109883

Fig. 5. Volume fractions of the hcp, bcc, fcc and other structures as a function of the quench temperature for various Ti IPs.

through the Stokes-Einstein relationship, D ~ kBT/η. Fig. 10(d) depicts 3. Discussions


Dη/T as calculated from the six IPs as a function of temperature. All IPs
predict an almost independent Dη/T from the temperature in the liquid 3.1. Primary Properties: Equilibrium and nonequilibrium phase stabilities
phase, in agreement with the Stokes-Einstein relation.
During the MAM process, materials undergo cyclic transformations
between various phases with variable cooling/heating rates, as ex-
plained in Fig. 2. Thus, the MD modeling of the MAM process requires
the use of IPs that reliably predict these phase transitions in equilibrium
and nonequilibrium thermodynamics. We consider the exhibition of the

6
S.A. Etesami, et al. Computational Materials Science 184 (2020) 109883

Fig. 6. The final snapshot of the solidified structure at low and high temperatures for different potentials: a) Ti1.eam.fs, b) Ti2.eam.fs, c) Ti3.eam.fs, d) NiTi.meam, e)
Ti.Dickel.meam, f) Ti.meam.spline. Each color relates to a crystal structure (fcc: green, hcp: red, bcc: blue, simple cubic: purple, other Structure: white). (For
interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

different solid phases of the material in equilibrium thermodynamics determined by calculating the free energy variations versus tempera-
through MD-calculated free energies as the first primary property. The ture for the competing phases (e.g. bcc or hcp). The phases with the
MD-calculated TTT diagram, as a representative nonequilibrium ther- lower free energy at each temperature are determined as the thermo-
modynamic characteristic of the material, is considered as the second dynamically stable phase. Notably, the Ti3.eam.fs IP almost exactly
primary property. matches the Calphad calculations of phase stabilities in Fig. 3(a), and
The α-β solid-state phase transformation is the basis for the design Ti1.eam.fs is the closest IP to the Calphad calculations of phase stabi-
of the MAM and post-MAM heat treatment processes for Ti alloys that lities in Fig. 3(b). In addition, Ti.set and Ti.Kim.meam IPs show that the
have a significant influence on the performance of the alloy [71,72]. hcp phase is the only stable phase in the temperature range of 900 K to
Therefore, we consider the capability of the IP to simply exhibit such a TM. Therefore, there are seven IPs, corresponding to Ti1.eam.fs,
transition as the primary requirement for use of the IP for MAM process Ti2.eam.fs, Ti3.eam.fs, NiTi.meam, Zope-Ti-Al-2003.eam, Ti.Dick-
modeling. The phase stability in equilibrium thermodynamics was el.meam and Ti.meam.spline, that exhibit a phase transition between

Table 1
MD calculations of the α-β and β-M transition temperatures, change in enthalpies, and change in volumes using six IPs, along with the experimental results [79].
Property Ti1.eam.fs Ti2.eam.fs Ti3.eam.fs NiTi.meam Ti.Dickel.meam Ti.meam.spline Exp.

Tα→β [K] 1155 1153 1152 1510 1163 1271 1155


TM [K] 1921 1322 1210 1719 1904 1881 1941
ΔHα→β [eV/atom] 0.022 0.032 0.042 0.032 0.028 0.037 0.044
ΔHβ→M[eV/atom] 0.128 0.097 0.083 0.093 0.102 0.093 0.157
ΔVα→β/Vα [%] −0.664 −0.009 0.713 0.520 0.873 0.314 —
ΔVβ→M/V β [%] 2.198 2.211 2.0987 1.132 1.079 0.734 —

7
S.A. Etesami, et al. Computational Materials Science 184 (2020) 109883

indexes as the secondary consideration for the development/verifica-


tion of MD IPs. For the Ti alloy, the α-β, β-liquid, and liquid–vapor
coexistences cyclically occur during the MAM process, and the related
secondary properties include the transition temperature, change in
enthalpy, change in volume, and liquid–vapor tension. Clearly, addi-
tional coexistence properties may be added to this list to enhance the
confidence in the MD IPs.
The calculated Tα→β values (Table 1) for all three EAM IPs are very
close to the experimental value (1155 K). The MEAM IP Ti.Dickel.meam
also predicts Tα→β within a 1% error from experiment. In contrast, the
error margins from the two other MEAM IPs, NiTi.meam and Ti.me-
am.spline, are 31% and 10%, respectively. This result is expected since
Tα→β was included in the IP development process for the three EAM and
Ti.Dickel.meam IPs. Furthermore, Ti3.eam.fs and Ti.meam.spline are
closest to the experimental values of ΔHα→β, with errors corresponding
to 4.5% and 15.9%, respectively. Ti1.eam.fs predictions of TM and
ΔHβ→M are closest to experiments, with errors corresponding to 1% and
10%, respectively. Ti.Dickel.meam shows a reasonable agreement with
Fig. 7. MD calculations of the vapor–liquid interfacial tension using six IPs,
experiments with errors of 11% and 35% for TM and ΔHβ→M, respec-
along with two experimental data sets corresponding to Experiment 1 (solid
tively. Here, we note that Ti1.eam.fs is unable to show the spontaneous
circles) [68] and Experiment 2 (open circles) [69].
transformation of a liquid to the solid α at lower temperatures, as
shown in Fig. 3(a). Furthermore, the MD-calculated expansion during
the hcp (α) and bcc (β) phases. The calculated TM and Tα→β by the melting using Ti1.eam.fs is approximately twice that of the MD calcu-
Zope-Ti-Al-2003.eam is 1535 K and 1520 K, respectively. Due to the lation using Ti.Dickel.meam.
very short range of temperatures between the Tα→β and the TM for this Ti2.eam.fs and Ti.Dickel.meam are the only IPs that predict a li-
IP, we exclude this IP from the list of the IPs that exhibit the α → β quid–vapor interfacial tension within a 5% deviation from the experi-
solid-state transition. This primary consideration leaves out four of the mental values (Fig. 7). The other four IPs overestimate (Ti3.eam.fs and
ten IPs for Ti from further evaluations. These four IPs are not able to NiTi.meam) and underestimate (Ti1.eam.fs and Ti.meam.spline) the
demonstrate experimentally expected solid-state transformation from experimental values by more than 5%. We note that the liquid–vapor
the α phase at low temperatures to the β phase at high temperatures. interfacial tension is not included as one of the fitting properties for the
These IPs have been developed for modeling different phenomena and parameterization of any of the considered Ti IPs.
processes, and the conclusions drawn from the present study are spe-
cific to the MAM process modeling capability. 3.3. Tertiary Properties: Single-phase properties
All six IPs predict typically expected TTT diagrams (Fig. 4), char-
acterized by long crystallization onset times at both low and high The variation in several single-phase properties as a function of
temperatures and relatively short crystallization times at intermediate temperature plays an important role in the predictive modeling of the
temperatures. These MD-calculated TTT diagrams are representative MAM process. For instance, in the liquid state, the shear viscosity in-
nonequilibrium properties of materials related to homogenous nuclea- fluences the dynamics of the flow within the melt-pool, and self-diffu-
tion, which can be compared to their experimental counterparts in the sivity influences the anisotropy in the distribution of the alloying ele-
development/verification process of IPs for the MAM process modeling. ments. In the solid-state, elastic constant variations with temperature
Given the availability of experimental TTT diagrams, the influence of influence the development of residual stresses during the MAM process.
the system size on the MD-calculated TTT diagrams should also be in- Therefore, the development/verification of MD IPs for MAM process
vestigated for robust comparisons. modeling should include a range of these properties as the tertiary
The percentages of solid phases depicted at Fig. 5 show that consideration.
Ti1.eam.fs results in the crystallization of the liquid into the bcc phase Fig. 8(a) shows that the predictions of Ti1.eam.fs, Ti2.eam.fs and
only, regardless of the temperature. Likewise, the Ti3.eam.fs results in Ti.meam.spline of C11 for the α phase are in good agreement with the
the crystallization of the liquid into the hcp phase only. In contrast, experimental results in the range 900–1150 K. Experiment 2 shows the
Ti2.eam.fs, NiTi.meam, Ti.Dickle.meam and Ti.meam.spline yield to elastic hardening of C11 near Tα→β, where Ti2.eam.fs depicts the same
the crystallization of the liquid into the hcp or the bcc phases at low or behavior. Experiment 2 also shows an increase in the value of C11 at
high temperatures, respectively. Since the MAM process modeling of Ti Tα→β. This trend is correctly observed using Ti1.eam.fs, Ti2.eam.fs and
requires an IP that correctly predicts stable hcp and bcc phases at low Ti3.eam.fs IPs. However, only Ti.Dickel.meam can accurately re-
and high temperatures, respectively, both Ti1.eam.fs and Ti3.eam.fs produce the C11 of the β phase from Tα→β to TM. The calculated C12 for
may not be suitable for MAM modeling of Ti. The fact that free energy the α phase by Ti2.eam.fs and Ti.meam.spline are in good agreement
calculations based on these two IPs predict thermodynamically stable with Experiment 2, while NiTi.meam and Ti.Dickel.meam predict C12 in
bcc phase at high temperatures, as shown by Fig. 3(a and b), implies the agreement with Experiment 1. For C12 of the β phase, the best results
existence of a relatively high energy barrier, preventing the sponta- are obtained by Ti3.eam.fs from Tα→β to 1250 K and Ti.meam.spline
neous transformation of the hcp phase to the bcc phase at high tem- from 1300 K to 1850 K. We note that all the IPs except Ti1.eam.fs show
peratures. It should be noted here that the above conclusions are drawn approximately no variation in C12 for the β phase with respect to
without studying the influence of the system size on the MD-calculated temperature in accord with the experimental results. The calculated
volume fractions and in the absence of experimental counterpart for values of C13 using NiTi.meam, Ti.Dickel.meam and Ti1.eam.fs (at high
these quantities. temperatures) for the α phase are the closest to the experimental values.
NiTi.meam and Ti.Dickel.meam predict a steadily increasing C13 with
3.2. Secondary Properties: Two-phase coexistence properties temperature without a steeper increase near Tα→β, while Experiment 2
shows considerable hardening near Tα→β. NiTi.meam and Ti.Dick-
The cyclic two-phase coexistence of various states of metals during el.meam predict C33 values that are close to the experimental values,
the MAM process secures the consideration of the related quantitative while Ti1.eam.fs and Ti2.eam.fs overestimate the experimental values

8
S.A. Etesami, et al. Computational Materials Science 184 (2020) 109883

Fig. 8. MD-calculated elastic constants for Ti in the α and β phases, along with the experimental results corresponding to Experiment 1 [66] and Experiment 2 [67].

and Ti3.eam.fs and Ti.meam.spline underestimate the experimental with both experiments and the Calphad calculations [73]. Note that all
values. C44 elastic softening is predicted for the α phase by all the IPs, the IPs predictions of the density of the α phase are within approxi-
whereas NiTi.meam and Ti.meam.spline predict values that are closest mately 1% from experiments. The deviation of the calculated density of
to the experimental values and the predictions from Ti2.eam.fs and the β phase is also small, i.e., within approximately 2% from the ex-
Ti3.eam.fs outperform other EAM IPs. All the IPs overestimate the value perimental values. However, the slope of the density of the β phase
of C44 of the β phase. Notably, crossing Tα→β accompanies a decreasing versus temperature, as predicted by all three MEAM IPs, is lower than
value of C44 that is only predicted by the EAM IPs. the experimental slope. For MD calculations of the excess enthalpy for
In regard to the density of the α and β solid phases shown in Fig. 9, the α and β phases, two EAM IPs, Ti3.eam.fs and Ti2.eam.fs, and two
all the considered IPs predict a decreasing density with increasing MEAM IPs, NiTi.meam and Ti.Dickel.meam, predict results that are
temperature, in qualitative agreement with the experimental results. within 10% margins from the experiments. However, all IPs predict a
The slope of the density of the α phase is found to be practically the slope of the enthalpy versus temperature of the α phase that is practi-
same for all IPs, except for Ti2.eam.fs and Ti3.eam.fs, in agreement cally equal to the experimental slope. All IPs, except Ti2.eam.fs, predict

9
S.A. Etesami, et al. Computational Materials Science 184 (2020) 109883

Fig. 9. MD calculations, using six Ti IPs, of (a) density of the α and the β phases, along with the experimental results: Experiment 1 [74], Experiment 2 [75],
Experiment 3 [76], and the Calphad calculations [73], (b) excess enthalpy of the α and the β phases, along with the experimental results [77], (c) liquid density along
with Experiment 1 [68] and Experiment 2 [69] and (d) excess enthalpy along with Experiment 1 [68] and Experiment 2 [69].

a deviation of the enthalpy of the β phase that is higher than 10%. For 4. Conclusion
the liquid phase, all six IPs predict a liquid density of Ti that is within a
5% error margin from the experimental values while all six IPs un- We presented an approach for the development and/or verification
derestimate the experimental enthalpy by more than 10% for all tem- of the IPs for the MD modeling of MAM processes and applied the
peratures. procedure for the ten available Ti IPs, which were mostly developed
Fig. 10(a) shows that almost all six Ti IPs exhibit a shift in the lo- considering 0 K properties and limited high temperature properties. The
cation of the first three peaks compared to the experimental results. approach considers the thermal history of the material during the MAM
Ti1.eam.fs predicts the location of the peaks closest to the experimental process, which includes the cyclic variable-rate thermal processing of
value. However, the values of the calculated S(k) for the first peak by metal alloys between room temperature and temperatures higher than
Ti2.eam.fs and Ti3.eam.fs are closest to the experimental values. We their melting points. The MAM thermal process results in the cyclic
note that there is a considerable deviation of the second peak of S(k), as transformation of the alloy among the various solid phases, liquid
predicted by all the IPs. The MD calculations of self-diffusion coefficient phase, and vapor phase. Thus, we considered the ability of the IP to
in the liquid phase, shown in Fig. 10(b), follows the expected Arrhenius demonstrate the equilibrium and nonequilibrium phase stability of the
form, D = D0exp(-Ea/kBT), where Ea is the activation energy. NiTi.- alloy as the primary consideration. Of the ten Ti IPs, only six demon-
meam is the only IP that can reproduce the experimental values of D strate the β → α solid-state phase transition under equilibrium condi-
within a 5% error margin. Ti1.eam.fs is the only EAM IP that can re- tions and were used for the remainder of the study. TTT diagrams were
produce the experimental values of D within a 10% error margin. In considered representative of the nonequilibrium phase stability prop-
regard to liquid shear viscosity calculations, shown in Fig. 10(c), the erties and were determined for the six Ti IPs. In the absence of an ex-
NiTi.meam IP predicts the lowest error margin ranging between 10% at perimental counterpart, no conclusion was drawn in this regard.
low temperatures and 5% at high temperatures. Increasing the tem- However, Ti2eam.fs, NiTi.meam, Ti.Dickel.meam, and Ti.meam.spline
perature results in a significant decrease in the error margin of the shear were the IPs that exhibit a spontaneous β → α transition in these si-
viscosity determined using Ti1.eam.fs, ranging from 39% at 1800 K to mulations.
approximately 10% at 2200 K. We considered the two-phase coexistence properties as the sec-
ondary consideration for the development/verification of IPs for MAM
process modeling. These properties include the melting point, the β →
α transition temperature, and the energy/volume differences associated

10
S.A. Etesami, et al. Computational Materials Science 184 (2020) 109883

Fig. 10. MD calculations, using six Ti IPs, of (a) liquid structure factor along with the experimental results at 1973 K [78], (b) self-diffusivity along with the
experimental results [27], (c) the shear viscosity along with the experimental results [68], and (d) Dη/T.

with these transitions as well as the liquid–vapor surface tension. quantitative and qualitative modeling of metal alloys during the MAM
Finally, we considered other relevant single-phase properties as the process considering their limitations, there is a need to develop so-
tertiary properties. These properties include the elastic constants, phisticated IPs, such as machine learning-based IPs, for higher accuracy
density, and enthalpy variations with temperature at the solid phase in the MD modeling of the MAM process for alloy systems.
and the structure factor, density and enthalpy, self-diffusion, and shear
viscosity variations with temperature at the liquid phase. Overall, the 5. Data availability
comparison of these MD-calculated secondary and tertiary properties to
experiments show that Ti2.eam.fs, Ti.Dickel.meam and Ti.meam.spline All the scripts and protocols to reproduce the data in this article are
are the best performing IPs for Ti; however, they may fall short in freely available upon request from the corresponding author.
predicting one or more of these properties in reasonable agreement
with experiments.
Author Contributions
In summary, we determined that Ti2.eam.fs, Ti.Dickel.meam, and
Ti.meam.spline IPs are suitable for the MD simulations of MAM process
S.A.E. wrote the manuscript with significant contributions from E.A.
modeling considering their few deficiencies in predicting some of the
and contributions from M.L. E.A. designed and supervised the research.
secondary and tertiary properties in agreement with the experimental
S.A.E. implemented the methods with some helps from E.A. and M.L.
results. Furthermore, there is a lack of experimental data for the
nonequilibrium thermodynamics of metal alloys, as this data relate to
the MAM process, e.g., the phase stability and kinetics at high cooling Declaration of Competing Interest
rates. Using such experimental data in the proposed approach will in-
crease the confidence in the capability of MD IPs for MAM process The authors declare that they have no known competing financial
modeling. interests or personal relationships that could have appeared to influ-
Given the significant added complexity for the development/ver- ence the work reported in this paper.
ification of IPs for metal alloy systems, more deviations with experi-
mental results are more likely to be observed in the development/ver- Acknowledgements
ification of classical IPs, such as the EAM and MEAM for MAM process
modeling. Although such classical IPs will still be effective in the This work is partially supported by the National Aeronautics and

11
S.A. Etesami, et al. Computational Materials Science 184 (2020) 109883

Space Administration (NASA) under Grant NNX16AT77G issued [23] M.I. Mendelev, G.J. Ackland, Development of an interatomic potential for the si-
through the NASA/Marshall Space Flight Center. The authors are also mulation of phase transformations in zirconium, Philos. Mag. Lett. 87 (2007)
349–359, https://doi.org/10.1080/09500830701191393.
grateful for the computer time allocation provided by the University of [24] S.A. Etesami, E. Asadi, Molecular dynamics for near melting temperatures simula-
Memphis’s High-Performance Computing Center. tions of metals using modified embedded-atom method, J. Phys. Chem. Solids. 112
(2018) 61–72 (accessed November 21, 2017), https://www.sciencedirect.com/
science/article/pii/S0022369717312039.
References [25] S.A. Etesami, M.I. Baskes, M. Laradji, E. Asadi, Thermodynamics of solid Sn and
PbSn liquid mixtures using molecular dynamics simulations, Acta Mater. 161
[1] N. Guo, M.C. Leu, Additive manufacturing: technology, applications and research (2018) 320–330, https://doi.org/10.1016/J.ACTAMAT.2018.09.036.
needs, Front. Mech. Eng. 8 (2013) 215–243, https://doi.org/10.1007/s11465-013- [26] S.A. Etesami, M. Laradji, E. Asadi, Transferability of interatomic potentials in pre-
0248-8. dicting the temperature dependency of elastic constants for titanium, zirconium and
[2] G. Tapia, A. Elwany, A Review on Process Monitoring and Control in Metal-Based magnesium, Model. Simul. Mater. Sci. Eng. 27 (2019) 025005, , https://doi.org/10.
Additive Manufacturing, J. Manuf. Sci. Eng. 136 (2014) 060801, , https://doi.org/ 1088/1361-651X/aaf617.
10.1115/1.4028540. [27] J. Horbach, R.E. Rozas, T. Unruh, A. Meyer, Improvement of computer simulation
[3] W.E. Frazier, Metal Additive Manufacturing: A Review, J. Mater. Eng. Perform. 23 models for metallic melts via quasielastic neutron scattering: A case study of liquid
(2014) 1917–1928, https://doi.org/10.1007/s11665-014-0958-z. titanium, Phys. Rev. B. 80 (2009) 212203, , https://doi.org/10.1103/PhysRevB.80.
[4] J. Smith, W. Xiong, W. Yan, S. Lin, P. Cheng, O.L. Kafka, G.J. Wagner, J. Cao, 212203.
W.K. Liu, Linking process, structure, property, and performance for metal-based [28] ASTM International, F2792–12a - Standard Terminology for Additive
additive manufacturing: computational approaches with experimental support, Manufacturing, Technologies (2013) 10–12, https://doi.org/10.1520/F2792-
Comput. Mech. 57 (2016) 583–610, https://doi.org/10.1007/s00466-015-1240-4. 12A.2.
[5] M.F. (Mark F. Horstemeyer, Integrated computational materials engineering (ICME) [29] W.J. Sames, K.A. Unocic, R.R. Dehoff, T. Lolla, S.S. Babu, Thermal effects on mi-
for metals : case studies, n.d. https://books.google.com/books?hl=en&lr=& crostructural heterogeneity of Inconel 718 materials fabricated by electron beam
id=OkJLDwAAQBAJ&oi=fnd&pg=PR19&dq=info:boGEtcixIcwJ:scholar.google. melting, J. Mater. Res. 29 (2014) 1920–1930, https://doi.org/10.1557/jmr.2014.
com&ots=FGmR8xAHsE&sig=soiE6SLjxIYcH8YMmqR0AZ7AvZo#v=onepage&q 140.
&f=false (accessed June 8, 2018). [30] G. Tapia, A.E.-J. of Manufacturing, undefined 2014, A review on process mon-
[6] M. Horstemeyer, Integrated Computational Materials Engineering (ICME) for me- itoring and control in metal-based additive manufacturing, Asmedigitalcollection.
tals: using multiscale modeling to invigorate engineering design with science, 2012. Asme.Org. (n.d.). http://biomechanical.asmedigitalcollection.asme.org/article.as-
https://books.google.com/books?hl=en&lr=&id=Pr-SzR5cosUC&oi=fnd& px?articleid=1905940 (accessed June 5, 2018).
pg=PR13&dq=Integrated+Computational+Materials+Engineering+(ICME) [31] S.K. Everton M. Hirsch P. Stravroulakis R.K. Leach A.T. Clare Review of in-situ
+for+metals:+using+multiscale+modeling+to+invigorate+engineering+de- process monitoring and in-situ metrology for metal additive manufacturing 95 2016
sign+with+science&ots=pe8uD7_BVc&sig=VxQA38P9ZBm_wMlZx_i8C7y9F6g 431 445 http://creativecommons.org/licenses/by/4.0/ (accessed January 21,
(accessed June 8, 2018). 2019).
[7] M. Seifi, A. Salem, J. Beuth, O. Harrysson, J.J. Lewandowski, Overview of Materials [32] O. Zinovieva, A. Zinoviev, V. Ploshikhin, Three-dimensional modeling of the mi-
Qualification Needs for Metal Additive Manufacturing, JOM. 68 (2016) 747–764, crostructure evolution during metal additive manufacturing, Comput. Mater. Sci.
https://doi.org/10.1007/s11837-015-1810-0. 141 (2018) 207–220, https://doi.org/10.1016/J.COMMATSCI.2017.09.018.
[8] Y. Shibuta, K. Oguchi, T. Takaki, M. Ohno, Homogeneous nucleation and micro- [33] H. Krauss T. Zeugner M.F. Zaeh Thermographic process monitoring in powderbed
structure evolution in million-atom molecular dynamics simulation, Sci. Rep. 5 based additive manufacturing 2015 in 10.1063/1.4914608 177 183.
(2015) 13534, https://doi.org/10.1038/srep13534. [34] M. Mahesh, B. Lane, A. Donmez, S. Feng, S. Moylan, R. Fesperman, NISTIR 8036
[9] Y. Shibuta, S. Sakane, E. Miyoshi, S. Okita, T. Takaki, M. Ohno, Heterogeneity in Measurement Science Needs for Real-time Control of Additive Manufacturing
homogeneous nucleation from billion-atom molecular dynamics simulation of so- Powder Bed Fusion Processes NISTIR 8036 Measurement Science Needs for Real-
lidification of pure metal, Nat. Commun. 8 (2017) 10, https://doi.org/10.1038/ time Control of Additive Manufacturing Powder Bed Fusion Processes Mahesh
s41467-017-00017-5. Mani, (n.d.). doi:10.6028/NIST.IR.8036.
[10] S. Ghosh, Predictive modeling of solidification during laser additive manufacturing [35] S. Clijsters, T. Craeghs, S. Buls, K. Kempen, J.-P. Kruth, In situ quality control of the
of nickel superalloys: recent developments, future directions, Mater. Res. Express. 5 selective laser melting process using a high-speed, real-time melt pool monitoring
(2018) 012001, , https://doi.org/10.1088/2053-1591/aaa04c. system, Int. J. Adv. Manuf. Technol. 75 (2014) 1089–1101, https://doi.org/10.
[11] H. Qiu, D. Yang, D. Luo, L. Hu, Y. Shang, L. Zhang, Y. Liu, W. Liang, Y. Teng, K. Zhu, 1007/s00170-014-6214-8.
Z. Yu, R. Liu, Y. Ma, L. Wei, Simulation study on the melting process of nano- [36] C. Lough, X. Wang, C. Smith, O. Adeniji, R. Landers, Use of SWIR Imaging to
corundum, Ceram. Int. 44 (2018) 5932–5938, https://doi.org/10.1016/J. Monitor Layer-to-Layer Part Quality during SLM of 304L Stainless Steel,
CERAMINT.2017.12.159. Sffsymposium.Engr.Utexas.Edu. (n.d.). http://sffsymposium.engr.utexas.edu/sites/
[12] Y. Wang, Q. Wei, F. Pan, M. Yang, S. Wei, Molecular dynamics simulations for the default/files/2018/182 UseofSWIRImagingtoMonitorLayerToLayerPart.pdf (ac-
examination of mechanical properties of hydroxyapatite/ poly α-n-butyl cyanoa- cessed June 18, 2019).
crylate under additive manufacturing, Biomed. Mater. Eng. 24 (2014) 825–833, [37] B. Lane, S. Moylan, E.P. Whitenton, L. Ma, Thermographic measurements of the
https://doi.org/10.3233/BME-130874. commercial laser powder bed fusion process at NIST, Rapid Prototyp. J. 22 (2016)
[13] Z. Qin, G.S. Jung, M.J. Kang, M.J. Buehler, The mechanics and design of a light- 778–787, https://doi.org/10.1108/RPJ-11-2015-0161.
weight three-dimensional graphene assembly, Sci. Adv. 3 (2017) e1601536, , [38] T. Keller, G. Lindwall, S. Ghosh, L. Ma, B.M. Lane, F. Zhang, U.R. Kattner, E.A. Lass,
https://doi.org/10.1126/sciadv.1601536. J.C. Heigel, Y. Idell, M.E. Williams, A.J. Allen, J.E. Guyer, L.E. Levine, Application
[14] Z. Zhao, J. Liu, A.K. Soh, C. Tang, Temperature-mediated fabrication, stress-induced of finite element, phase-field, and CALPHAD-based methods to additive manu-
crystallization and transformation: atomistic simulations of additively manu- facturing of Ni-based superalloys, Acta Mater. 139 (2017) 244–253, https://doi.
factured amorphous Cu pillars, Model. Simul. Mater. Sci. Eng. 27 (2019) 075012, , org/10.1016/J.ACTAMAT.2017.05.003.
https://doi.org/10.1088/1361-651X/AB3758. [39] Z. Gan, Y. Lian, S.E. Lin, K.K. Jones, W.K. Liu, G.J. Wagner, Benchmark Study of
[15] M.S. Daw, M.I. Baskes, Embedded-atom method: Derivation and application to Thermal Behavior, Surface Topography, and Dendritic Microstructure in Selective
impurities, surfaces, and other defects in metals, Phys. Rev. B. 29 (1984) Laser Melting of Inconel 625, Integr. Mater. Manuf. Innov. 8 (2019) 178–193,
6443–6453, https://doi.org/10.1103/PhysRevB.29.6443. https://doi.org/10.1007/s40192-019-00130-x.
[16] M.I. Baskes, Modified embedded-atom potentials for cubic materials and impurities, [40] S. Ghosh, L. Ma, L.E. Levine, R.E. Ricker, M.R. Stoudt, J.C. Heigel, J.E. Guyer,
Phys. Rev. B. 46 (1992) 2727–2742, https://doi.org/10.1103/PhysRevB.46.2727. Single-Track Melt-Pool Measurements and Microstructures in Inconel 625, JOM. 70
[17] B.-J. Lee, M.I. Baskes, Second nearest-neighbor modified embedded-atom-method (2018) 1011–1016, https://doi.org/10.1007/s11837-018-2771-x.
potential, Phys. Rev. B. 62 (2000) 8564–8567, https://doi.org/10.1103/PhysRevB. [41] J.C. Heigel, B.M. Lane, MEASUREMENT OF THE MELT POOL LENGTH DURING
62.8564. SINGLE SCAN TRACKS IN A COMMERCIAL LASER POWDER BED FUSION
[18] E. Asadi, M. Zaeem, S. Nouranian, M. Baskes, Two-phase solid–liquid coexistence of PROCESS, n.d. https://ws680.nist.gov/publication/get_pdf.cfm?pub_id=922273
Ni, Cu, and Al by molecular dynamics simulations using the modified embedded- (accessed June 17, 2019).
atom method, Acta Mater (2015). [42] J.C. Heigel, B.M. Lane, THE EFFECT OF POWDER ON COOLING RATE AND MELT
[19] E. Asadi, M. Asle Zaeem, S. Nouranian, M.I. Baskes, Quantitative modeling of the POOL LENGTH MEASUREMENTS USING IN SITU THERMOGRAPHIC TECNIQUES,
equilibration of two-phase solid-liquid Fe by atomistic simulations on diffusive time n.d. https://sffsymposium.engr.utexas.edu/sites/default/files/2017/Manuscripts/
scales, Phys. Rev. B. 91 (2015) 024105, , https://doi.org/10.1103/PhysRevB.91. TheEffectofPowderonCoolingRateandMeltPool.pdf (accessed June 18, 2019).
024105. [43] W. Mizukami, S. Habershon, D.P. Tew, A compact and accurate semi-global po-
[20] S.R. Wilson, M.I. Mendelev, A unified relation for the solid-liquid interface free tential energy surface for malonaldehyde from constrained least squares regression,
energy of pure FCC, BCC, and HCP metals, J. Chem. Phys. 144 (2016) 144707, , J. Chem. Phys. 141 (2014) 144310, , https://doi.org/10.1063/1.4897486.
https://doi.org/10.1063/1.4946032. [44] A. Glielmo, P. Sollich, A. De Vita, Accurate interatomic force fields via machine
[21] D.Y. Sun, M.I. Mendelev, C.A. Becker, K. Kudin, T. Haxhimali, M. Asta, J.J. Hoyt, learning with covariant kernels, Phys. Rev. B. 95 (2017) 214302, , https://doi.org/
A. Karma, D.J. Srolovitz, Crystal-melt interfacial free energies in hcp metals: A 10.1103/PhysRevB.95.214302.
molecular dynamics study of Mg, Phys. Rev. B. 73 (2006) 024116, , https://doi.org/ [45] V. Botu, R. Ramprasad, Learning scheme to predict atomic forces and accelerate
10.1103/PhysRevB.73.024116. materials simulations, Phys. Rev. B. 92 (2015) 094306, , https://doi.org/10.1103/
[22] M.I. Mendelev, T.L. Underwood, G.J. Ackland, Development of an interatomic po- PhysRevB.92.094306.
tential for the simulation of defects, plasticity, and phase transformations in tita- [46] V. Botu, R. Ramprasad, Adaptive machine learning framework to accelerate ab initio
nium, J. Chem. Phys. 145 (2016) 154102, , https://doi.org/10.1063/1.4964654. molecular dynamics, Int. J. Quantum Chem. 115 (2015) 1074–1083, https://doi.

12
S.A. Etesami, et al. Computational Materials Science 184 (2020) 109883

org/10.1002/qua.24836. Dependent Phenomena. II. Irreversible Processes in, Fluids, J. Chem. Phys. 22
[47] G.P. Purja Pun, R. Batra, R. Ramprasad, Y. Mishin, Physically informed artificial (1954) 398–413, https://doi.org/10.1063/1.1740082.
neural networks for atomistic modeling of materials, (n.d.). doi:10.1038/s41467- [64] R. Kubo, Statistical-Mechanical Theory of Irreversible Processes. I. General Theory
019-10343-5. and Simple Applications to Magnetic and Conduction Problems, J. Phys. Soc. Japan.
[48] X.W. Zhou, R.A. Johnson, H.N.G. Wadley, Misfit-energy-increasing dislocations in 12 (1957) 570–586, https://doi.org/10.1143/JPSJ.12.570.
vapor-deposited CoFe/NiFe multilayers, Phys. Rev. B. 69 (2004) 144113, , https:// [65] A.A.T.A. Dinsdale, SGTE data for pure elements, Calphad. 15 (1991) 317–425,
doi.org/10.1103/PhysRevB.69.144113. https://doi.org/10.1016/0364-5916(91)90030-N.
[49] G.J. Ackland, Theoretical study of titanium surfaces and defects with a new many- [66] E.S. Fisher, C.J. Renken, Single-Crystal Elastic Moduli and the hcp → bcc
body potential, Philos. Mag. A. 66 (1992) 917–932, https://doi.org/10.1080/ Transformation in Ti, Zr, and Hf, Phys. Rev. 135 (1964) A482–A494, https://doi.
01418619208247999. org/10.1103/PhysRev.135.A482.
[50] R.R. Zope, Y. Mishin, Interatomic potentials for atomistic simulations of the Ti-Al [67] H. Ogi, S. Kai, H. Ledbetter, R. Tarumi, M. Hirao, K. Takashima, Titanium’s high-
system, Phys. Rev. B. 68 (2003) 024102, , https://doi.org/10.1103/PhysRevB.68. temperature elastic constants through the hcp–bcc phase transformation, Acta
024102. Mater. 52 (2004) 2075–2080, https://doi.org/10.1016/J.ACTAMAT.2004.01.002.
[51] W.-S. Ko, B. Grabowski, J. Neugebauer, Development and application of a Ni-Ti [68] K. Zhou, H.P. Wang, J. Chang, B. Wei, Experimental study of surface tension,
interatomic potential with high predictive accuracy of the martensitic phase tran- specific heat and thermal diffusivity of liquid and solid titanium, Chem. Phys. Lett.
sition, Phys. Rev. B. 92 (2015) 134107, , https://doi.org/10.1103/PhysRevB.92. 639 (2015) 105–108, https://doi.org/10.1016/J.CPLETT.2015.09.014.
134107. [69] J.J. Wessing, J. Brillo, Density, Molar Volume, and Surface Tension of Liquid Al-Ti,
[52] Y.-M. Kim, B.-J. Lee, M.I. Baskes, Modified embedded-atom method interatomic Metall. Mater. Trans. A. 48 (2017) 868–882, https://doi.org/10.1007/s11661-016-
potentials for Ti and Zr, Phys. Rev. B. 74 (2006) 014101, , https://doi.org/10.1103/ 3886-8.
PhysRevB.74.014101. [70] E. Asadi, M. Asle Zaeem, S. Nouranian, M.I. Baskes, Two-phase solid–liquid coex-
[53] D. Dickel, C.D. Barrett, R.L. Carino, M.I. Baskes, M.F. Horstemeyer, Mechanical istence of Ni, Cu, and Al by molecular dynamics simulations using the modified
instabilities in the modeling of phase transitions of titanium, Model. Simul. Mater. embedded-atom method, Acta Mater. 86 (2015) 169–181, https://doi.org/10.
Sci. Eng. 26 (2018) 065002, , https://doi.org/10.1088/1361-651X/aac95d. 1016/j.actamat.2014.12.010.
[54] R.G. Hennig, T.J. Lenosky, D.R. Trinkle, S.P. Rudin, J.W. Wilkins, Classical potential [71] B. Fotovvati, S.A. Etesami, E. Asadi, Process-property-geometry correlations for
describes martensitic phase transformations between the α, β, and ω titanium additively-manufactured Ti–6Al–4V sheets, Mater. Sci. Eng. A. 760 (2019)
phases, Phys. Rev. B. 78 (2008) 054121, , https://doi.org/10.1103/PhysRevB.78. 431–447, https://doi.org/10.1016/J.MSEA.2019.06.020.
054121. [72] W.J. Sames, F.A. List, S. Pannala, R.R. Dehoff, S.S. Babu, F.A. List, S. Pannala,
[55] S. Plimpton, Fast Parallel Algorithms for Short-Range Molecular Dynamics, J. R.R. Dehoff, S.S.B. The, W.J. Sames, F.A. List, S. Pannala, R.R. Dehoff, S.S. Babu,
Comput. Phys. 117 (1995) 1–19, https://doi.org/10.1006/jcph.1995.1039. The metallurgy and processing science of metal additive manufacturing The me-
[56] A. Stukowski, Visualization and analysis of atomistic simulation data with tallurgy and processing science of metal additive manufacturing 6608 (2016),
OVITO–the Open Visualization Tool, Model. Simul. Mater. Sci. Eng. 18 (2010) https://doi.org/10.1080/09506608.2015.1116649.
015012, , https://doi.org/10.1088/0965-0393/18/1/015012. [73] Xiao-Gang Lu Malin Selleby Bo Sundman Assessments of molar volume and thermal
[57] http://ovito.org/, (n.d.). expansion for selected bcc, fcc and hcp metallic elements Calphad 29 1 2005 68 89
[58] D. Frenkel, B. Smit, Understanding molecular simulation: from algorithms to, ap- 10.1016/j.calphad.2005.05.001 https://linkinghub.elsevier.com/retrieve/pii/
plications (2001). S0364591605000362.
[59] R. Freitas, M. Asta, M. de Koning, Nonequilibrium free-energy calculation of solids [74] J. Spreadborough, J.W. Christian, The Measurement of the Lattice Expansions and
using LAMMPS, Comput. Mater. Sci. 112 (2016) 333–341, https://doi.org/10. Debye Temperatures of Titanium and Silver by X-ray Methods, Proc. Phys. Soc. 74
1016/J.COMMATSCI.2015.10.050. (1959) 609–615, https://doi.org/10.1088/0370-1328/74/5/314.
[60] M. Parrinello, A. Rahman, Polymorphic transitions in single crystals: A new mole- [75] P. Schmitz-Pranghe, N.,, Dünner, Congenital Atresia of Esophagus with Esophago-
cular dynamics method, J. Appl. Phys. 52 (1981) 7182, https://doi.org/10.1063/1. Tracheal Fistula, J Med Res. 59 (1968) 377.
328693. [76] R.H. Willens, Vacuum X-Ray Diffractometer for High Temperature Studies of Metals
[61] P.M. Larsen, S. Schmidt, J. Schiøtz, Robust structural identification via polyhedral Sensitive to Contamination by Oxygen and Nitrogen, Rev. Sci. Instrum. 33 (1962)
template matching, Model. Simul. Mater. Sci. Eng. 24 (2016) 055007, , https://doi. 1069–1076, https://doi.org/10.1063/1.1717685.
org/10.1088/0965-0393/24/5/055007. [77] M.W. Chase, NIST-JANAF Thermochemical Tables, Washington, D. C, 1998.
[62] J.H. Irving, J.G. Kirkwood, The Statistical Mechanical Theory of Transport [78] Y. Waseda, The structure of non-crystalline materials: liquids and amorphous solids,
Processes. IV. The Equations of Hydrodynamics, J. Chem. Phys. 18 (1950) 817–829, McGraw_Hill, New York, 1980.
https://doi.org/10.1063/1.1747782. [79] A.I. Efimov, L. P. Belorukova, L. V Vasilkova, Svoistva Neorganicheskikh
[63] M.S. Green, Markoff,, Random Processes and the Statistical Mechanics of Time- Soedinenii, Himiia, Leningrad, 1983.

13

You might also like