Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
License: CC BY 4.0
arXiv:2309.05806v2 [cond-mat.dis-nn] 09 Jan 2024

Simulated multi-component CuZr(Al) metallic glasses akin to experiments

Rene Alvarez-Donado11{}^{1}start_FLOATSUPERSCRIPT 1 end_FLOATSUPERSCRIPT, Silvia Bonfanti11{}^{1}start_FLOATSUPERSCRIPT 1 end_FLOATSUPERSCRIPT, Mikko Alava1,212{}^{1,2}start_FLOATSUPERSCRIPT 1 , 2 end_FLOATSUPERSCRIPT 11{}^{1}start_FLOATSUPERSCRIPT 1 end_FLOATSUPERSCRIPT NOMATEN Centre of Excellence, National Center for Nuclear Research, ul. A. Soltana 7, 05-400 Swierk/Otwock, Poland
22{}^{2}start_FLOATSUPERSCRIPT 2 end_FLOATSUPERSCRIPT Aalto University, Department of Applied Physics, PO Box 11000, 00076 Aalto, Espoo, Finland
Abstract

We study a three-component CuZrAl metallic glass system by means of a combined Monte Carlo and Molecular Dynamics simulations scheme. This hybrid method allows us to generate equilibrated samples at temperatures below the conventional glass transition for the first time, achieving a more stable glassy regime. By using a realistic potential for the interactions of metallic species, we explore the kinetics, thermodynamics, and rheology of a CuZrAl glass, and then compare these findings with those of the ubiquitous CuZr. Remarkably, the resulting sheared glassy configurations show an abrupt stress drop corresponding to the shear band, akin to experimental observations. Our results pave the way for theoretical studies of complex metallic glasses and offer comparisons with experiments.

Metallic glasses (MGs) are an important class of materials constituted by an amorphous atomic structure, which forms when the high-temperature metallic liquid is rapidly cooling down to room temperature [1, 2, 3]. The unique disordered structure thereby formed is the key to many outstanding properties, such as unprecedented mechanical features, including high strength and high elastic limit. Hence, MGs have great potential to outperform conventional materials faor various technological and industrial applications [2, 4, 5]. Nevertheless, the major limitation that prevents their use on a large scale, is the difficulty to avoid the crystallization of the samples during the cooling process [6]. Recent advancement in synthesis techniques has come from a method different from conventional cooling: vapor deposition. This technique allows the production of MGs with enhanced stability, known as ultrastable metallic glasses (UMGs) [7, 8, 9, 10, 11]. A resultant shift in the glass transition temperature, indicative of this enhanced stability, has been observed in various UMGs such as for Cu4646{}_{46}start_FLOATSUBSCRIPT 46 end_FLOATSUBSCRIPTZr4646{}_{46}start_FLOATSUBSCRIPT 46 end_FLOATSUBSCRIPTAl88{}_{8}start_FLOATSUBSCRIPT 8 end_FLOATSUBSCRIPT MG in Ref. [9]. These glasses are thought to be equivalent to conventionally liquid quenched MGs that have undergone aging for thousands of years [12]. This trend was initially discovered by Ediger and co-workers but in the class of organic glasses [13]. In general, ultrastable glasses are characterized by extraordinary thermodynamic and kinetic stability and exceptional mechanical properties [13, 14, 15, 16]

The molecular mechanisms underlying the glass-forming ability (GFA), shift of glass transition temperature, and formation of enhanced stable states in MGs are far from being understood, a challenge that also characterizes the broad spectrum of glasses [17, 18]. Molecular simulations have been fundamental to investigate the properties of glasses from the microscopic point of view [19, 20]. However, simulations suffer from unrealistic high-cooling rates ([101000]similar-toabsentdelimited-[]101000\sim[10-1000]∼ [ 10 - 1000 ] K/ns), resulting in computer-generated metallic glasses with properties that differ significantly from those observed in experiments [21, 22, 23, 24, 25]. Here, we solve this gap by obtaining in silico samples of realistic CuZr-based ultrastable MGs similar to experiments, using an efficient hybrid Molecular Dynamics and Monte Carlo (MD+MC) simulation approach. Our study focus on two-component Cu5050{}_{50}start_FLOATSUBSCRIPT 50 end_FLOATSUBSCRIPTZr5050{}_{50}start_FLOATSUBSCRIPT 50 end_FLOATSUBSCRIPT MGs as a starting point, and ternary component Cu4646{}_{46}start_FLOATSUBSCRIPT 46 end_FLOATSUBSCRIPTZr4646{}_{46}start_FLOATSUBSCRIPT 46 end_FLOATSUBSCRIPTAl88{}_{8}start_FLOATSUBSCRIPT 8 end_FLOATSUBSCRIPT MGs, both known for their exceptional GFA, outstanding mechanical properties and significant potential in engineering applications [26, 27, 28, 29, 9].

Recently, the limitation of timescale of standard simulations has been overcome through the development of a very efficient method, called Swap Monte Carlo (SMC)  [30, 31, 32]. Here the equilibration time increases less rapidly than the relaxation time ταsubscript𝜏𝛼\tau_{\alpha}italic_τ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT as the temperature decreases [33, 34], allowing for the computation of the equilibrium configurational entropy below the experimental Tgsubscript𝑇𝑔T_{g}italic_T start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT [35, 36, 37]. The SMC algorithm has seemed to be the sole means to investigate the properties of ultrastable glasses in silico. Despite the accelerated dynamics facilitated by swap moves, the crystallization effects commonly observed in conventional simulations continue to occur in SMC, but can be resolved by enforcing a degree of polydispersity, which effectively prevents crystallization and enables the generation of deeply cooled equilibrium liquid configurations, even at temperatures below the glass transition Tgsubscript𝑇𝑔T_{g}italic_T start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT, comparable with experiments. So far SMC has been applied only to a simplify model of metallic glass, a ternary mixture Lennard-Jones, that does not consider the specific nature of the constituents [38], therefore limiting our knowledge of the chemical composition landscape for MGs.

An intriguing solution to accelerate the dynamics sampling in the chemical configuration space [39] comes from the field of multicomponent crystalline alloys and combines MC methods, MD, and realistic short range potential, e.g. Embedded Atom Method (EAM) [40]. Only very recently, this techniques has been applied in glasses to obtain more stable realistic samples of binary CuZr MG in computer simulations [41], where it was shown that shear transformation zones are limited to small clusters of particles. Here, we demonstrate how this approach is suitable for tackling various concentrations of different species and that it is applicable to produce more stable realistic samples of multicomponent MGs in silico. We introduce for the first time equilibrium configurations of ternary metallic MGs, specifically Cu4646{}_{46}start_FLOATSUBSCRIPT 46 end_FLOATSUBSCRIPTZr4646{}_{46}start_FLOATSUBSCRIPT 46 end_FLOATSUBSCRIPTAl88{}_{8}start_FLOATSUBSCRIPT 8 end_FLOATSUBSCRIPT, approaching experiments. We examine their thermodynamic and more stable states and compare the mechanical properties. Our work presents a first example of generating computationally realistic models for any type of metallic glass consisting of several atomic species.

Simulation methods — The interatomic interactions are simulated through an Embedded Atom Method (EAM) as developed by [27]. For thermodynamics and kinetics, we performed simulations consisting of N𝑁Nitalic_N=1500 atoms in a cubic box with periodic boundary conditions in three dimensions. Our simulations are performed with LAMMPS [42], using a time step ΔtΔ𝑡\Delta troman_Δ italic_t=1 fs. The glass state is obtained through quenching in the isobaric-isothermal ensemble (NpT𝑁𝑝𝑇NpTitalic_N italic_p italic_T) from the liquid at 2000 K to 300 K using a cooling rate κ=1012𝜅superscript1012\kappa=10^{12}italic_κ = 10 start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT K/s. We save configurations every 50 K during the cooling process for further analysis. The cooling process is performed by integrating the Nosé-Hoover equations with damping parameters τTsubscript𝜏𝑇\tau_{T}italic_τ start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT=2 fs and τpsubscript𝜏𝑝\tau_{p}italic_τ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT=5 ps for the thermostat and barostat. All results are obtained keeping the external pressure p𝑝pitalic_p=0.
Hybrid Molecular Dynamics-Monte Carlo (MD+MC) algorithm — In order to generate glasses that attempt to emulate the mechanical behavior observed in experiments, we employ a hybrid Molecular Dynamics-Monte Carlo (MD+MC) scheme under the variance-constrained semi-grand canonical ensemble (VC-SGC) [39]. Differently from traditional SMC, this VC-SGC MC scheme allows exploring the configurational degrees of freedom by randomly selecting an atom and attempting to change its type, while also calculating the corresponding energy and concentration changes. It allows to target specific concentration ranges while maintaining fixed total number of particles and the volume. Acceptance of these transmutations follows the Metropolis criterion, ensuring the preservation of detailed balance. On the other hand, the relaxation processes are accounted for by the MD integration steps. To maintain the desired concentration within the system [39], we set the variance parameter κ𝜅\kappaitalic_κ=1033{}^{3}start_FLOATSUPERSCRIPT 3 end_FLOATSUPERSCRIPT. Furthermore, we evaluate the differences in chemical potential relative to Zr using hybrid MD+MC simulations under the semi-grand canonical ensemble at a temperature of 2000 K. The specific set of parameters that minimize the composition errors in relation to the desired concentration can be found in the Appendix.
Quenching Protocol — The quenching is performed starting from molten metals at high temperature well above the melting point using a fixed cooling rate of 1012superscript101210^{12}10 start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT K/s at fixed pressure. For the MD+MC scheme, every 20 MD steps, a MC cycle consisting of N/4𝑁4N/4italic_N / 4 attempts is performed. For a comparison between different cooling rates with MD and MD+MC, see Appendix.

Results — First, we compare the thermal evolution of the potential energy per particle between a conventional MD simulation and our MD+MC scheme for the two MG systems. Figure 1 reports the potential energy per atom (U/N𝑈𝑁U/Nitalic_U / italic_N) as a function of the temperature T𝑇Titalic_T during quenching, obtained by averaging over 5 independent simulations.

Refer to caption
Figure 1: Thermal evolution of the potential energy per atom of the binary Cu5050{}_{50}start_FLOATSUBSCRIPT 50 end_FLOATSUBSCRIPTZr5050{}_{50}start_FLOATSUBSCRIPT 50 end_FLOATSUBSCRIPT and the ternary Cu4646{}_{46}start_FLOATSUBSCRIPT 46 end_FLOATSUBSCRIPTZr4646{}_{46}start_FLOATSUBSCRIPT 46 end_FLOATSUBSCRIPTAl88{}_{8}start_FLOATSUBSCRIPT 8 end_FLOATSUBSCRIPT MGs with a cooling rate of 1012superscript101210^{12}10 start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT K/s using both a pure MD and a hybrid MD+MC algorithm (red lines). At high temperatures, both methods exhibit similar results. However, as the temperature decreases, the behavior diverges significantly. The MC+MD method remains in equilibrium for lower temperatures compared to the pure MD approach.

Notably, the hybrid algorithm allows us to quench the liquid at lower temperatures, increasing the chance of reaching more stable configurations and properties similar to those observed in experiments. Note that both methods, MD and MC+MD, exhibit the same trend at high temperatures, as expected.
After the quenching we evaluate the relaxation time of the two different dynamics: conventional MD and MC+MD, for the ternary CuZrAl system. For each temperature we switch from NpT𝑁𝑝𝑇NpTitalic_N italic_p italic_T to NVT𝑁𝑉𝑇NVTitalic_N italic_V italic_T ensemble and run equilibration runs to compute the self-part of the intermediate scattering function [32]. For the case of MC+MD we employ the VC-SGC method as described above. The relaxation time, ταsubscript𝜏𝛼\tau_{\alpha}italic_τ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT, defined as the value at which the intermediate scattering function is equal to 1/e, is reported in Fig. 2 for the standard MD (circles) and MC+MD dynamics (squares).

Refer to caption
Figure 2: Comparison of relaxation time ταsubscript𝜏𝛼\tau_{\alpha}italic_τ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT for standard MD simulations and hybrid MD+MC with VC-SGC method for the ternary CuZrAl MG. The speedup offered by the MD+MC dynamics is evident.

The speedup of the dynamics given by our MD+MC scheme is shown as about two orders of magnitude at the lowest temperature.

It is well known that the GFA of CuZr MG improves when a small percentage of Al is included [7]. In a previous study of the same MG [43], we estimated TgMDsubscriptsuperscript𝑇𝑀𝐷𝑔T^{MD}_{g}italic_T start_POSTSUPERSCRIPT italic_M italic_D end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT to be 623623623623 K and 713713713713 K for Cu5050{}_{50}start_FLOATSUBSCRIPT 50 end_FLOATSUBSCRIPTZr5050{}_{50}start_FLOATSUBSCRIPT 50 end_FLOATSUBSCRIPT and Cu4646{}_{46}start_FLOATSUBSCRIPT 46 end_FLOATSUBSCRIPTZr4646{}_{46}start_FLOATSUBSCRIPT 46 end_FLOATSUBSCRIPTAl88{}_{8}start_FLOATSUBSCRIPT 8 end_FLOATSUBSCRIPT, respectively. By following a similar protocol, we observe the same trend in the MD+MC algorithm, indicating that the method mimics the melt-quench procedure but provides equilibrium configurations for temperatures below the conventional Tgsubscript𝑇𝑔T_{g}italic_T start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT, meaning that it creates more stable glasses. Throughout the quenching process, in both pure MD and MD+MC, we carefully analyzed the structure and found no evidence of crystallization in the alloys.

Thermodynamics — Once we successfully create glasses approaching experimental conditions, our focus shifts to exploring their thermodynamic properties. To being with, we determine the configurational entropy of the glass formed by MD+MC by the formula Sconf=StotSvibsubscript𝑆𝑐𝑜𝑛𝑓subscript𝑆𝑡𝑜𝑡subscript𝑆𝑣𝑖𝑏S_{conf}=S_{tot}-S_{vib}italic_S start_POSTSUBSCRIPT italic_c italic_o italic_n italic_f end_POSTSUBSCRIPT = italic_S start_POSTSUBSCRIPT italic_t italic_o italic_t end_POSTSUBSCRIPT - italic_S start_POSTSUBSCRIPT italic_v italic_i italic_b end_POSTSUBSCRIPT, where Stotsubscript𝑆𝑡𝑜𝑡S_{tot}italic_S start_POSTSUBSCRIPT italic_t italic_o italic_t end_POSTSUBSCRIPT represents the total entropy and Svibsubscript𝑆𝑣𝑖𝑏S_{vib}italic_S start_POSTSUBSCRIPT italic_v italic_i italic_b end_POSTSUBSCRIPT stands for the vibrational contribution. We obtain both entropies using the reversible scaling (RS) method [44], with the Uhlenbeck-Ford potential and the Einstein crystal serving as reference systems, respectively [45, 46, 47] (see Supplementary Information). Figure 3 displays the configurational entropy (Sconf)subscript𝑆𝑐𝑜𝑛𝑓(S_{conf})( italic_S start_POSTSUBSCRIPT italic_c italic_o italic_n italic_f end_POSTSUBSCRIPT ) plotted against temperature for both the binary Cu5050{}_{50}start_FLOATSUBSCRIPT 50 end_FLOATSUBSCRIPTZr5050{}_{50}start_FLOATSUBSCRIPT 50 end_FLOATSUBSCRIPT and ternary Cu4646{}_{46}start_FLOATSUBSCRIPT 46 end_FLOATSUBSCRIPTZr4646{}_{46}start_FLOATSUBSCRIPT 46 end_FLOATSUBSCRIPTAl88{}_{8}start_FLOATSUBSCRIPT 8 end_FLOATSUBSCRIPT alloys. As anticipated, Sconfsubscript𝑆𝑐𝑜𝑛𝑓S_{conf}italic_S start_POSTSUBSCRIPT italic_c italic_o italic_n italic_f end_POSTSUBSCRIPT decreases with decreasing temperature until it freezes in at lower temperatures due to the glass transition. The measured residual entropy values, representing the constant Sconfsubscript𝑆𝑐𝑜𝑛𝑓S_{conf}italic_S start_POSTSUBSCRIPT italic_c italic_o italic_n italic_f end_POSTSUBSCRIPT value at the occurrence of the glass transition, were 0.049 and 0.035 J/mol-K for Cu5050{}_{50}start_FLOATSUBSCRIPT 50 end_FLOATSUBSCRIPTZr5050{}_{50}start_FLOATSUBSCRIPT 50 end_FLOATSUBSCRIPT and Cu4646{}_{46}start_FLOATSUBSCRIPT 46 end_FLOATSUBSCRIPTZr4646{}_{46}start_FLOATSUBSCRIPT 46 end_FLOATSUBSCRIPTAl88{}_{8}start_FLOATSUBSCRIPT 8 end_FLOATSUBSCRIPT, respectively. Note in particular that the Sconfsubscript𝑆𝑐𝑜𝑛𝑓S_{conf}italic_S start_POSTSUBSCRIPT italic_c italic_o italic_n italic_f end_POSTSUBSCRIPT of Cu4646{}_{46}start_FLOATSUBSCRIPT 46 end_FLOATSUBSCRIPTZr4646{}_{46}start_FLOATSUBSCRIPT 46 end_FLOATSUBSCRIPTAl88{}_{8}start_FLOATSUBSCRIPT 8 end_FLOATSUBSCRIPT are lower than the binary alloy, the reason for that is because the ternary alloy possesses a high GFA allowing to reach deeper values before the freeze-in at Tgsubscript𝑇𝑔T_{g}italic_T start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT (which is higher). The reported value of Sconfsubscript𝑆𝑐𝑜𝑛𝑓S_{conf}italic_S start_POSTSUBSCRIPT italic_c italic_o italic_n italic_f end_POSTSUBSCRIPT in the Cu5050{}_{50}start_FLOATSUBSCRIPT 50 end_FLOATSUBSCRIPTZr5050{}_{50}start_FLOATSUBSCRIPT 50 end_FLOATSUBSCRIPT alloy is just a 20%percent2020\%20 % from the one reported in situ by Smith et al. [48]. We are dealing with glass configurations with glass transition temperatures lower than the one obtained in conventional glasses, thus it is expected to have a lower value of the residual entropy. A comparison with experiments for the ternary system is currently not available.

Refer to caption
Figure 3: Configurational entropy as a function of temperature obtained through RS method for Cu5050{}_{50}start_FLOATSUBSCRIPT 50 end_FLOATSUBSCRIPTZr5050{}_{50}start_FLOATSUBSCRIPT 50 end_FLOATSUBSCRIPT (blue) and Cu4646{}_{46}start_FLOATSUBSCRIPT 46 end_FLOATSUBSCRIPTZr4646{}_{46}start_FLOATSUBSCRIPT 46 end_FLOATSUBSCRIPTAl88{}_{8}start_FLOATSUBSCRIPT 8 end_FLOATSUBSCRIPT (red).
Refer to caption
Figure 4: Heat capacity as a function of temperature obtained through RS method for a)a)italic_a ) Cu4646{}_{46}start_FLOATSUBSCRIPT 46 end_FLOATSUBSCRIPTZr4646{}_{46}start_FLOATSUBSCRIPT 46 end_FLOATSUBSCRIPTAl88{}_{8}start_FLOATSUBSCRIPT 8 end_FLOATSUBSCRIPT and b)b)italic_b ) Cu5050{}_{50}start_FLOATSUBSCRIPT 50 end_FLOATSUBSCRIPTZr5050{}_{50}start_FLOATSUBSCRIPT 50 end_FLOATSUBSCRIPT. The insets show the potential energy of the binary and ternary alloys, respectively, as a function of temperature during the heating process using MD (thin lines) and MD+MC (dashed lines).

Following experiments on ultra stable MGs using vapor-deposition method, a common approach to studying their stability is by performing calorimetric measurements during a heating process. In Figure 4, we compare the behavior of glasses prepared using conventional MD with those created using the MD+MC algorithm. We heat the glass at the same rate of 1012superscript101210^{12}10 start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT K/s and save the configuration every 20202020 K. Then, we compute the heat capacity using its statistical definition: (E2E2)/kBT2delimited-⟨⟩superscript𝐸2superscriptdelimited-⟨⟩𝐸2subscript𝑘𝐵superscript𝑇2(\langle E^{2}\rangle-\langle E\rangle^{2})/k_{B}T^{2}( ⟨ italic_E start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ - ⟨ italic_E ⟩ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) / italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, where E𝐸Eitalic_E is the energy of the system, and kBsubscript𝑘𝐵k_{B}italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT is the Boltzmann constant. The insets of Figures 4a)a)italic_a ) and 4b)b)italic_b ) with the potential energy of the binary and ternary alloys, respectively, illustrate as well via the difference of the two kinds of glasses as a function of temperature. Clearly, the devitrification process of the glass created through the hybrid MD+MC happens at a higher temperature, approximately 18%percent1818\%18 % above Tgsubscript𝑇𝑔T_{g}italic_T start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT [49, 50, 15]. In vapor-deposited experiments [7], this temperature is called the “onset” temperature Tosubscript𝑇𝑜T_{o}italic_T start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT, and it is generally between 6666 to 10%percent1010\%10 % above the conventional Tgsubscript𝑇𝑔T_{g}italic_T start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT, reflecting the much larger kinetic stability reached by this method. Since our MD+MC glass produces similar results, we can confirm that the Cu4646{}_{46}start_FLOATSUBSCRIPT 46 end_FLOATSUBSCRIPTZr4646{}_{46}start_FLOATSUBSCRIPT 46 end_FLOATSUBSCRIPTAl88{}_{8}start_FLOATSUBSCRIPT 8 end_FLOATSUBSCRIPT and Cu5050{}_{50}start_FLOATSUBSCRIPT 50 end_FLOATSUBSCRIPTZr5050{}_{50}start_FLOATSUBSCRIPT 50 end_FLOATSUBSCRIPT alloys produced by this method are more stable configurations.

Rheology — For mechanical tests we perform athermal quasistatic simulations [51] starting from configurations with N𝑁Nitalic_N=100000 atoms, at T𝑇Titalic_T=300 K cooled down to 0 K, following the work of Ref. [52]. We shear the simulation box along the x𝑥xitalic_x direction by the amount δγ𝛿𝛾\delta\gammaitalic_δ italic_γ. We perform energy minimization after each strain step with the fast inertial relaxation engine (FIRE) [53] until the system reaches mechanical equilibrium, until a maximum force of 101010{}^{-10}start_FLOATSUPERSCRIPT - 10 end_FLOATSUPERSCRIPTeV/Å  is reached.

Refer to caption
Refer to caption
Figure 5: Scaled shear-modulus stress-strain curves for MD and MD+MC schemes (upper panel). Strain inhomogeneity maps for a strain of 0.125 for the CuZrAl cases. The color bar indicates the non affine displacements from the unstrained configuration, from 0% (blue) to 20% (red).

Figure 5 (top) shows how the composition and the history influence the mechanical behavior. The addition of aluminum increases by a few percent the shear modulus [54], and to compare the glasses at peak stress we scale away the dependence of on the modulus. These are typical stress-strain curves, where after an initial elastic part we see the typical glass response: strain-softening close to the maximum, fluctuations, and a drop in stress before plastic flow. In particular, as in model glasses [41] we observe that the more stable glasses exhibit drastic stress drops which corresponds to the the appearance of a system-spanning shear band (Inset). The well-prepared glasses do so at slightly larger strains than the basic ones, and in scaled units they are noticeably stronger, here about 20%. The final state of flow one finds may be summarized by the examples of the strain fields in the Fig. 5 (bottom), with the features associated with the post-yield shear bands depending on the nature of the stress-peak: for glasses with enhanced stability we find a well-defined shear band whereas for the reference glass the band is more diffuse [38].

Summary & Conclusions — In order to resolve the physics of glasses in silico - a big challenge due to the complexity of glasses and their history-dependence - novel ideas are needed. In this work, we show for the first time how a combination of Molecular Dynamics and Monte Carlo with VC-SGC method is able to tackle this issue for realistic multi-component MGs.

We present results that clearly show a convergence to the experimental observations as regards the thermodynamics of these systems, we also demonstrate, that such glasses have the ultrastable character, aligning them closely with vapor-deposited experimental counterparts. Finally, this fact is seen to make a difference in the mechanical properties. The next steps to follow are a through exploration of the relaxation times the MC+MD approach leads to and the subsequent changes in the yielding.

In conclusion, our results provide a substantial contribution to the field of glass physics, particularly with regard to multicomponent MGs with more than two component species [55, 56]. Using advanced simulations to obtain ultrastable glassy states, we provide an essential theoretical framework for further studies, shedding light on the underlying physical mechanisms governing the complex behaviors and properties of these amorphous materials. Our work prompts further theoretical investigations aimed at probing the microscopic mechanisms of complex MGs at a fundamental level, such as their relaxation dynamics and facilitation [57]. The methodologies employed in this work not only serve as a benchmark for experimental comparison but also suggest new avenues for optimizing the structure of MGs via cooling and processing techniques, for a range of applications.

Acknowledgements: We gratefully thank Anshul Parmar and Misaki Ozawa for fruitful discussions. Funding: RAD, SB, and MA are supported by the European Union Horizon 2020 research and innovation program under grant agreement no. 857470 and from the European Regional Development Fund via the Foundation for Polish Science International Research Agenda PLUS program grant No. MAB PLUS/2018/8.

Appendix A APPENDIX

Appendix B Setting VC-SGC parameters

In the variance-constrained semi-grand canonical (VC-SGC) ensemble[39], the system explores configurational degrees of freedom by randomly selecting an atom and attempting to change its type. The change in energy and concentration is then computed, and transmutations are accepted or rejected based on the Metropolis criterion, ensuring detailed balance and equilibrium sampling of the phase space. To determine the difference in chemical potential species (Δμ)Δ𝜇(\Delta\mu)( roman_Δ italic_μ ), which fixes the desired composition and minimizes error, we conduct a series of semi-grand-canonical (SGC) simulations at a temperature of T=2000𝑇2000T=2000italic_T = 2000 K.
Figure 6 depicts the behavior of ΔμΔ𝜇\Delta\muroman_Δ italic_μ with regard to Zr for a)a)italic_a ) Cu and b)b)italic_b ) Al for a series of SGC simulations. The blue line is obtained expanding ΔμΔ𝜇\Delta\muroman_Δ italic_μ as a polynomial in concentration as follows:

Δμ(XZr,p,T)=Tln(XZr1XZr)+i=0nAiXZri,Δ𝜇subscript𝑋Zr𝑝𝑇𝑇subscript𝑋Zr1subscript𝑋Zrsuperscriptsubscript𝑖0𝑛subscript𝐴𝑖superscriptsubscript𝑋Zr𝑖\Delta\mu(X_{\rm{Zr}},p,T)=T\ln\left(\frac{X_{\rm{Zr}}}{1-X_{\rm{Zr}}}\right)+% \sum_{i=0}^{n}A_{i}X_{\rm{Zr}}^{i},roman_Δ italic_μ ( italic_X start_POSTSUBSCRIPT roman_Zr end_POSTSUBSCRIPT , italic_p , italic_T ) = italic_T roman_ln ( divide start_ARG italic_X start_POSTSUBSCRIPT roman_Zr end_POSTSUBSCRIPT end_ARG start_ARG 1 - italic_X start_POSTSUBSCRIPT roman_Zr end_POSTSUBSCRIPT end_ARG ) + ∑ start_POSTSUBSCRIPT italic_i = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_X start_POSTSUBSCRIPT roman_Zr end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT , (1)

where XZrsubscript𝑋ZrX_{\rm{Zr}}italic_X start_POSTSUBSCRIPT roman_Zr end_POSTSUBSCRIPT is the Zr composition and the coefficient Aisubscript𝐴𝑖A_{i}italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are fitting parameters.

Refer to caption
Figure 6: Chemical potential as a function of the composition for a)a)italic_a ) Zr-Cu and b)b)italic_b ) Zr-Al from SGC simulations at T=2000𝑇2000T=2000italic_T = 2000 K. The blue solid line is a fitting obtained by Eq. 1.

By utilizing Eq (1), we can effectively determine the chemical potential at any desired composition. For both the binary Cu5050{}_{50}start_FLOATSUBSCRIPT 50 end_FLOATSUBSCRIPTZr5050{}_{50}start_FLOATSUBSCRIPT 50 end_FLOATSUBSCRIPT and ternary Cu4646{}_{46}start_FLOATSUBSCRIPT 46 end_FLOATSUBSCRIPTZr4646{}_{46}start_FLOATSUBSCRIPT 46 end_FLOATSUBSCRIPTAl88{}_{8}start_FLOATSUBSCRIPT 8 end_FLOATSUBSCRIPT alloys, the calculated chemical potential differences were ΔμZrCu=2.69854Δsubscript𝜇ZrCu2.69854\Delta\mu_{\rm Zr-Cu}=-2.69854roman_Δ italic_μ start_POSTSUBSCRIPT roman_Zr - roman_Cu end_POSTSUBSCRIPT = - 2.69854 eV, and ΔμZrCu=2.90351Δsubscript𝜇ZrCu2.90351\Delta\mu_{\rm Zr-Cu}=-2.90351roman_Δ italic_μ start_POSTSUBSCRIPT roman_Zr - roman_Cu end_POSTSUBSCRIPT = - 2.90351 eV, ΔμZrAl=0.55781Δsubscript𝜇ZrAl0.55781\Delta\mu_{\rm Zr-Al}=-0.55781roman_Δ italic_μ start_POSTSUBSCRIPT roman_Zr - roman_Al end_POSTSUBSCRIPT = - 0.55781 eV, respectively. Once the composition is defined, we proceed with the hybrid MC+MD simulation under the VC-SGC ensemble. This combination of MD+MC efficiently allows us to create in silico glasses with properties similar to the ones obtained in experiments.

Appendix C Cooling rate comparison

Refer to caption
Figure 7: Potential energy for various cooling rates for the MD vs. the hybrid algorithm for a)a)italic_a ) Zr-Cu and b)b)italic_b ) Zr-Al.

In Figure 7, we compare the thermalization efficiency of the hybrid MC+MD algorithm with conventional MD simulations. Starting from high temperatures well above the melting point, we rapidly cool the liquid using fixed cooling rates of 1014superscript101410^{14}10 start_POSTSUPERSCRIPT 14 end_POSTSUPERSCRIPT, 1013superscript101310^{13}10 start_POSTSUPERSCRIPT 13 end_POSTSUPERSCRIPT, and 1012superscript101210^{12}10 start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT through MD simulation. With the hybrid MC+MD approach, we observe the same cooling rate-dependent trend at higher temperatures. However, when employing the MC+MD algorithm, the liquid deviates from equilibrium at lower temperatures. The agreement among all the curves at high temperatures confirms the proper utilization of the MC+MD approach in our simulations.

Appendix D Computing the configurational entropy

We have numerically calculated the entropy by means of the thermodynamic relation S=(HG)/T𝑆𝐻𝐺𝑇S=(H-G)/Titalic_S = ( italic_H - italic_G ) / italic_T, where H and G are the enthalpy and Gibbs free energy of the system, respectively. Since free energy is a thermal property [58] we are not able to compute it directly from convectional MD simulations. Thus, we have used a combination of the adiabatic switching (AS)[59, 60] and reversible scaling (RS)[61, 62] methods, which in principle allows the estimation of the free energy for a wide range of temperatures in a single simulation. We estimated the free energy (and the entropy) by carrying out the following steps:

  • (i)

    The preparation of the initial state in which we want to calculate the free energy. This state is defined as:

    MEAM(Γα)=𝒯(Γα)+UMEAM(Γα),subscriptMEAMsubscriptΓ𝛼𝒯subscriptΓ𝛼subscript𝑈MEAMsubscriptΓ𝛼\mathcal{H}_{\rm MEAM}(\Gamma_{\alpha})=\mathcal{T}(\Gamma_{\alpha})+U_{\rm MEAM% }(\Gamma_{\alpha}),caligraphic_H start_POSTSUBSCRIPT roman_MEAM end_POSTSUBSCRIPT ( roman_Γ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ) = caligraphic_T ( roman_Γ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ) + italic_U start_POSTSUBSCRIPT roman_MEAM end_POSTSUBSCRIPT ( roman_Γ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ) , (2)

    where MEAMsubscriptMEAM\mathcal{H}_{\rm MEAM}caligraphic_H start_POSTSUBSCRIPT roman_MEAM end_POSTSUBSCRIPT is the Hamiltonian of the alloy described by EAM potential, 𝒯𝒯\mathcal{T}caligraphic_T and UEAMsubscript𝑈EAMU_{\rm EAM}italic_U start_POSTSUBSCRIPT roman_EAM end_POSTSUBSCRIPT are the kinetic and potential energy, respectively. ΓαsubscriptΓ𝛼\Gamma_{\alpha}roman_Γ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT is a phase space vector with information of all momenta and positions as a function of time, and the subindex α𝛼\alphaitalic_α holds to crystal (g)𝑔(g)( italic_g ) or liquid (l)𝑙(l)( italic_l ).

  • (ii)

    Once defined the initial state, we estimate its absolute free energy through AS method. The AS is a variation of the thermodynamic integration method[63] where the Hamiltonian of the initial state is coupled with a reference system whose free energy is known beforehand. However, the work done is obtained dynamically in one single simulation. Thus, the free energy is obtained by means of:

    GEAM0=GRefα+𝒲AS,superscriptsubscript𝐺EAM0superscriptsubscript𝐺𝑅𝑒𝑓𝛼subscript𝒲𝐴𝑆G_{\rm EAM}^{0}=G_{Ref}^{\alpha}+\mathcal{W}_{AS},italic_G start_POSTSUBSCRIPT roman_EAM end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT = italic_G start_POSTSUBSCRIPT italic_R italic_e italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT + caligraphic_W start_POSTSUBSCRIPT italic_A italic_S end_POSTSUBSCRIPT , (3)

    where GEAM0superscriptsubscript𝐺EAM0G_{\rm EAM}^{0}italic_G start_POSTSUBSCRIPT roman_EAM end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT is the free energy of the alloy in the initial state, GRefαsuperscriptsubscript𝐺𝑅𝑒𝑓𝛼G_{Ref}^{\alpha}italic_G start_POSTSUBSCRIPT italic_R italic_e italic_f end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT is the free energy of the reference system, and 𝒲ASsubscript𝒲𝐴𝑆\mathcal{W}_{AS}caligraphic_W start_POSTSUBSCRIPT italic_A italic_S end_POSTSUBSCRIPT is the quasistatic work obtained through AS-method.

  • (iii)

    Using the value of GEAM0superscriptsubscript𝐺EAM0G_{\rm EAM}^{0}italic_G start_POSTSUBSCRIPT roman_EAM end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT and the RS method we obtain the alloy-free energy for a wide range of temperatures. This method is based on writing the initial Hamiltonian in the parametric form[61]:

    EAM(Γα,ζ(t))=𝒯+ζ(t)UEAM.subscriptEAMsubscriptΓ𝛼𝜁𝑡𝒯𝜁𝑡subscript𝑈EAM\mathcal{H}_{\rm EAM}(\Gamma_{\alpha},\zeta(t))=\mathcal{T}+\zeta(t)U_{\rm EAM}.caligraphic_H start_POSTSUBSCRIPT roman_EAM end_POSTSUBSCRIPT ( roman_Γ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT , italic_ζ ( italic_t ) ) = caligraphic_T + italic_ζ ( italic_t ) italic_U start_POSTSUBSCRIPT roman_EAM end_POSTSUBSCRIPT . (4)

    Here each value of ζ𝜁\zetaitalic_ζ corresponds to a particular state of HEAMsubscript𝐻EAMH_{\rm EAM}italic_H start_POSTSUBSCRIPT roman_EAM end_POSTSUBSCRIPT in a particular temperature and we use 𝒯𝒯\mathcal{T}caligraphic_T and UEAMsubscript𝑈EAMU_{\rm EAM}italic_U start_POSTSUBSCRIPT roman_EAM end_POSTSUBSCRIPT instead of 𝒯(Γα)𝒯subscriptΓ𝛼\mathcal{T}(\Gamma_{\alpha})caligraphic_T ( roman_Γ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ) and UEAM(Γα)subscript𝑈EAMsubscriptΓ𝛼U_{\rm EAM}(\Gamma_{\alpha})italic_U start_POSTSUBSCRIPT roman_EAM end_POSTSUBSCRIPT ( roman_Γ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ) to simplify the notation. One can show that using Eq. (4) the free energy of the alloy at any temperature is given by the equation[61, 44]:

    GEAM(T)=GEAM0ζ+32nkBT0lnζζ++1ζ0tsim𝑑tdζdtUEAM,subscript𝐺EAM𝑇superscriptsubscript𝐺EAM0𝜁32𝑛subscript𝑘𝐵subscript𝑇0𝜁𝜁1𝜁superscriptsubscript0subscript𝑡𝑠𝑖𝑚differential-d𝑡𝑑𝜁𝑑𝑡subscript𝑈EAM\begin{split}G_{\rm EAM}(T)=\frac{G_{\rm EAM}^{0}}{\zeta}+\frac{3}{2}nk_{B}T_{% 0}\frac{\ln\zeta}{\zeta}+\\ +\frac{1}{\zeta}\int_{0}^{t_{sim}}dt\frac{d\zeta}{dt}U_{\rm EAM},\end{split}start_ROW start_CELL italic_G start_POSTSUBSCRIPT roman_EAM end_POSTSUBSCRIPT ( italic_T ) = divide start_ARG italic_G start_POSTSUBSCRIPT roman_EAM end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT end_ARG start_ARG italic_ζ end_ARG + divide start_ARG 3 end_ARG start_ARG 2 end_ARG italic_n italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT divide start_ARG roman_ln italic_ζ end_ARG start_ARG italic_ζ end_ARG + end_CELL end_ROW start_ROW start_CELL + divide start_ARG 1 end_ARG start_ARG italic_ζ end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t start_POSTSUBSCRIPT italic_s italic_i italic_m end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_d italic_t divide start_ARG italic_d italic_ζ end_ARG start_ARG italic_d italic_t end_ARG italic_U start_POSTSUBSCRIPT roman_EAM end_POSTSUBSCRIPT , end_CELL end_ROW (5)

    where T0subscript𝑇0T_{0}italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the temperature at which GEAM0subscriptsuperscript𝐺0EAMG^{0}_{\rm EAM}italic_G start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_EAM end_POSTSUBSCRIPT was estimated in item (ii) and t sim is the simulation time.

  • (iv)

    Finally, the configurational entropy is obtained as:

    Sconf=SlSg,subscript𝑆𝑐𝑜𝑛𝑓subscript𝑆𝑙subscript𝑆𝑔S_{conf}=S_{l}-S_{g},italic_S start_POSTSUBSCRIPT italic_c italic_o italic_n italic_f end_POSTSUBSCRIPT = italic_S start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT - italic_S start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT , (6)

    where Slsubscript𝑆𝑙S_{l}italic_S start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT and Sgsubscript𝑆𝑔S_{g}italic_S start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT are the liquid and glass entropies, respectively.

The reference free energy GEAM0subscriptsuperscript𝐺0EAMG^{0}_{\rm EAM}italic_G start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_EAM end_POSTSUBSCRIPT in Eq. (3) is obtained by transforming the MEAM interaction into a reference system by means of AS method. For the solid phase at T=300𝑇300T=300italic_T = 300 K, we used the Einstein crystal with a spring constant α=5eV/Å2𝛼5𝑒𝑉superscriptitalic-Å2\alpha=5eV/\AA^{2}italic_α = 5 italic_e italic_V / italic_Å start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT for all the particles of the system[46, 64]. The liquid phase, on the other hand, GEAM0subscriptsuperscript𝐺0EAMG^{0}_{\rm EAM}italic_G start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_EAM end_POSTSUBSCRIPT was obtained at T=2800𝑇2800T=2800italic_T = 2800 K using as a reference the Uhlenbeck-Ford model (UFM) defined as[45, 65]:

UUFM(r)=pβln(1e(r/σ)2),subscript𝑈UFM𝑟𝑝𝛽1superscript𝑒superscript𝑟𝜎2U_{\rm UFM}(r)=-\frac{p}{\beta}\ln\left(1-e^{-(r/\sigma)^{2}}\right),italic_U start_POSTSUBSCRIPT roman_UFM end_POSTSUBSCRIPT ( italic_r ) = - divide start_ARG italic_p end_ARG start_ARG italic_β end_ARG roman_ln ( 1 - italic_e start_POSTSUPERSCRIPT - ( italic_r / italic_σ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ) , (7)

where β=1/(kBT)𝛽1subscript𝑘𝐵𝑇\beta=1/(k_{B}T)italic_β = 1 / ( italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T ) is the inverse of the temperature times the Boltzmann constant kBsubscript𝑘𝐵k_{B}italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT , and p𝑝pitalic_p and σ𝜎\sigmaitalic_σ are the softness and length scaling parameters, respectively. Following Paula-Leite et al.[45] a value of p=50𝑝50p=50italic_p = 50 gives a liquid-like behavior for the UFM system. The quasistatic work WASsubscript𝑊𝐴𝑆W_{AS}italic_W start_POSTSUBSCRIPT italic_A italic_S end_POSTSUBSCRIPT is estimating and evaluating the irreversible work during two adiabatic switching trajectories. One transforming the system from the EAM to the reference system (WEAMrefirrsubscriptsuperscript𝑊𝑖𝑟𝑟EAMrefW^{irr}_{\rm EAM\rightarrow ref}italic_W start_POSTSUPERSCRIPT italic_i italic_r italic_r end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_EAM → roman_ref end_POSTSUBSCRIPT) and the other one in the inverse process (WrefEAMirrsubscriptsuperscript𝑊𝑖𝑟𝑟refEAMW^{irr}_{\rm ref\rightarrow EAM}italic_W start_POSTSUPERSCRIPT italic_i italic_r italic_r end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_ref → roman_EAM end_POSTSUBSCRIPT). The irreversible work done in the AS method is calculated by Wirr=0tsim𝑑tdλdtU(Γα)superscript𝑊𝑖𝑟𝑟superscriptsubscript0subscript𝑡𝑠𝑖𝑚differential-d𝑡𝑑𝜆𝑑𝑡𝑈subscriptΓ𝛼W^{irr}=\int_{0}^{t_{sim}}dt\frac{d\lambda}{dt}U(\Gamma_{\alpha})italic_W start_POSTSUPERSCRIPT italic_i italic_r italic_r end_POSTSUPERSCRIPT = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t start_POSTSUBSCRIPT italic_s italic_i italic_m end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_d italic_t divide start_ARG italic_d italic_λ end_ARG start_ARG italic_d italic_t end_ARG italic_U ( roman_Γ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ), where U(Γα)𝑈subscriptΓ𝛼U(\Gamma_{\alpha})italic_U ( roman_Γ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ) is the hybrid potential connecting UEAMsubscript𝑈EAMU_{\rm EAM}italic_U start_POSTSUBSCRIPT roman_EAM end_POSTSUBSCRIPT and Urefsubscript𝑈𝑟𝑒𝑓U_{ref}italic_U start_POSTSUBSCRIPT italic_r italic_e italic_f end_POSTSUBSCRIPT and λ𝜆\lambdaitalic_λ is the AS coupling parameter. Thus,

𝒲AS=12(WEAMrefirrWrsfEAMirr).aformulae-sequencesubscript𝒲𝐴𝑆12superscriptsubscript𝑊EAMref𝑖𝑟𝑟superscriptsubscript𝑊rsfEAM𝑖𝑟𝑟𝑎\mathcal{W}_{AS}=\frac{1}{2}\left(W_{\rm EAM\rightarrow ref}^{irr}-W_{\rm rsf% \rightarrow EAM}^{irr}\right).acaligraphic_W start_POSTSUBSCRIPT italic_A italic_S end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( italic_W start_POSTSUBSCRIPT roman_EAM → roman_ref end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i italic_r italic_r end_POSTSUPERSCRIPT - italic_W start_POSTSUBSCRIPT roman_rsf → roman_EAM end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i italic_r italic_r end_POSTSUPERSCRIPT ) . italic_a (8)

Equation (8) is only valid if the dynamical work is performed slowly enough for holding the system in the linear response regime[44, 61].

Refer to caption
Figure 8: Free energy and Entropy as a function of temperature obtained by the RS method of the ternary Cu4646{}_{46}start_FLOATSUBSCRIPT 46 end_FLOATSUBSCRIPTZr4646{}_{46}start_FLOATSUBSCRIPT 46 end_FLOATSUBSCRIPTAl88{}_{8}start_FLOATSUBSCRIPT 8 end_FLOATSUBSCRIPT alloy.

In Fig. 8 we show the absolute values of the free energy and the entropy obtained by the RS method of the ternary Cu4646{}_{46}start_FLOATSUBSCRIPT 46 end_FLOATSUBSCRIPTZr4646{}_{46}start_FLOATSUBSCRIPT 46 end_FLOATSUBSCRIPTAl88{}_{8}start_FLOATSUBSCRIPT 8 end_FLOATSUBSCRIPT.

References

  • Greer [1995] A. L. Greer, Metallic glasses, Science 267, 1947 (1995).
  • Inoue [2000] A. Inoue, Stabilization of metallic supercooled liquid and bulk amorphous alloys, Acta Materialia 48, 279 (2000).
  • Wang et al. [2004] W.-H. Wang, C. Dong, and C. Shek, Bulk metallic glasses, Materials Science and Engineering: R: Reports 44, 45 (2004).
  • Ashby and Greer [2006] M. Ashby and A. L. Greer, Metallic glasses as structural materials, Scripta Materialia 54, 321 (2006).
  • Schroers [2010] J. Schroers, Processing of bulk metallic glass, Advanced Materials 22, 1566 (2010).
  • Johnson [1999] W. L. Johnson, Bulk glass-forming metallic alloys: Science and technology, MRS bulletin 24, 42 (1999).
  • Yu et al. [2013] H.-B. Yu, Y. Luo, and K. Samwer, Ultrastable metallic glass, Advanced Materials 25, 5904 (2013).
  • Lüttich et al. [2018] M. Lüttich, V. M. Giordano, S. Le Floch, E. Pineda, F. Zontone, Y. Luo, K. Samwer, and B. Ruta, Anti-aging in ultrastable metallic glasses, Physical review letters 120, 135504 (2018).
  • Luo et al. [2018] P. Luo, C. Cao, F. Zhu, Y. Lv, Y. Liu, P. Wen, H. Bai, G. Vaughan, M. Di Michiel, B. Ruta, et al., Ultrastable metallic glasses formed on cold substrates, Nature communications 9, 1389 (2018).
  • Sun et al. [2020] Q. Sun, D. M. Miskovic, K. Laws, H. Kong, X. Geng, and M. Ferry, Transition towards ultrastable metallic glasses in zr-based thin films, Applied Surface Science 533, 147453 (2020).
  • Zhao et al. [2022] Y. Zhao, B. Shang, B. Zhang, X. Tong, H. Ke, H. Bai, and W.-H. Wang, Ultrastable metallic glass by room temperature aging, Science Advances 8, eabn3623 (2022).
  • Sun et al. [2021] Q. Sun, D. M. Miskovic, and M. Ferry, Film thickness effect on formation of ultrastable metallic glasses, Materials Today Physics 18, 100370 (2021).
  • Swallen et al. [2007] S. F. Swallen, K. L. Kearns, M. K. Mapes, Y. S. Kim, R. J. McMahon, M. D. Ediger, T. Wu, L. Yu, and S. Satija, Organic glasses with exceptional thermodynamic and kinetic stability, Science 315, 353 (2007).
  • Kearns et al. [2009] K. L. Kearns, T. Still, G. Fytas, and M. Ediger, High-modulus organic glasses prepared by physical vapor deposition, Advanced Materials 1, 39 (2009).
  • Lyubimov et al. [2013] I. Lyubimov, M. D. Ediger, and J. J. de Pablo, Model vapor-deposited glasses: Growth front and composition effects, The Journal of Chemical Physics 139 (2013).
  • Rodriguez-Tinoco et al. [2022] C. Rodriguez-Tinoco, M. Gonzalez-Silveira, M. A. Ramos, and J. Rodriguez-Viejo, Ultrastable glasses: new perspectives for an old problem, La Rivista del Nuovo Cimento 45, 325 (2022).
  • Debenedetti and Stillinger [2001] P. G. Debenedetti and F. H. Stillinger, Supercooled liquids and the glass transition, Nature 410, 259 (2001).
  • Dyre [2006] J. C. Dyre, Colloquium: The glass transition and elastic models of glass-forming liquids, Reviews of modern physics 78, 953 (2006).
  • Binder and Kob [2011] K. Binder and W. Kob, Glassy materials and disordered solids: An introduction to their statistical mechanics (World scientific, 2011).
  • Ding et al. [2014] J. Ding, S. Patinet, M. L. Falk, Y. Cheng, and E. Ma, Soft spots and their structural signature in a metallic glass, Proceedings of the National Academy of Sciences 111, 14052 (2014).
  • Ashwin et al. [2013] J. Ashwin, E. Bouchbinder, and I. Procaccia, Cooling-rate dependence of the shear modulus of amorphous solids, Physical Review E 87, 042310 (2013).
  • Ryltsev et al. [2016] R. Ryltsev, B. Klumov, N. Chtchelkatchev, and K. Y. Shunyaev, Cooling rate dependence of simulated cu64. 5zr35. 5 metallic glass structure, The Journal of Chemical Physics 145 (2016).
  • Zhang et al. [2015] Y. Zhang, F. Zhang, C. Wang, M. Mendelev, M. Kramer, and K. Ho, Cooling rates dependence of medium-range order development in c u 64. 5 z r 35. 5 metallic glass, Physical Review B 91, 064105 (2015).
  • Vollmayr et al. [1996] K. Vollmayr, W. Kob, and K. Binder, How do the properties of a glass depend on the cooling rate? a computer simulation study of a lennard-jones system, The Journal of Chemical Physics 105, 4714 (1996).
  • Liu et al. [2007] Y. Liu, H. Bei, C. T. Liu, and E. P. George, Cooling-rate induced softening in a zr5050{}_{50}start_FLOATSUBSCRIPT 50 end_FLOATSUBSCRIPTcu5050{}_{50}start_FLOATSUBSCRIPT 50 end_FLOATSUBSCRIPT bulk metallic glass, Applied Physics Letters 90 (2007).
  • Yu et al. [2010] H. Yu, W. Wang, and H. Bai, An electronic structure perspective on glass-forming ability in metallic glasses, Applied Physics Letters 96 (2010).
  • Cheng et al. [2009] Y. Cheng, E. Ma, and H. Sheng, Atomic level structure in multicomponent bulk metallic glass, Physical Review Letters 102, 245501 (2009).
  • Ding and Cheng [2014] J. Ding and Y. Cheng, Charge transfer and atomic-level pressure in metallic glasses, Applied Physics Letters 104 (2014).
  • Pauly et al. [2009] S. Pauly, J. Das, N. Mattern, D. Kim, and J. Eckert, Phase formation and thermal stability in cu–zr–ti (al) metallic glasses, Intermetallics 17, 453 (2009).
  • Gutiérrez et al. [2015] R. Gutiérrez, S. Karmakar, Y. G. Pollack, and I. Procaccia, The static lengthscale characterizing the glass transition at lower temperatures, Europhysics Letters 111, 56009 (2015).
  • Berthier et al. [2016] L. Berthier, D. Coslovich, A. Ninarello, and M. Ozawa, Equilibrium sampling of hard spheres up to the jamming density and beyond, Physical review letters 116, 238002 (2016).
  • Ninarello et al. [2017] A. Ninarello, L. Berthier, and D. Coslovich, Models and algorithms for the next generation of glass transition studies, Physical Review X 7, 021039 (2017).
  • Grigera and Parisi [2001] T. S. Grigera and G. Parisi, Fast monte carlo algorithm for supercooled soft spheres, Physical Review E 63, 045102 (2001).
  • Berthier et al. [2019a] L. Berthier, E. Flenner, C. J. Fullerton, C. Scalliet, and M. Singh, Efficient swap algorithms for molecular dynamics simulations of equilibrium supercooled liquids, Journal of Statistical Mechanics: Theory and Experiment 2019, 064004 (2019a).
  • Berthier et al. [2017] L. Berthier, P. Charbonneau, D. Coslovich, A. Ninarello, M. Ozawa, and S. Yaida, Configurational entropy measurements in extremely supercooled liquids that break the glass ceiling, Proceedings of the National Academy of Sciences 114, 11356 (2017).
  • Ozawa et al. [2018] M. Ozawa, G. Parisi, and L. Berthier, Configurational entropy of polydisperse supercooled liquids, The Journal of Chemical Physics 149 (2018).
  • Berthier et al. [2019b] L. Berthier, M. Ozawa, and C. Scalliet, Configurational entropy of glass-forming liquids, The Journal of Chemical Physics 150 (2019b).
  • Parmar et al. [2020a] A. D. Parmar, M. Ozawa, and L. Berthier, Ultrastable metallic glasses in silico, Physical Review Letters 125, 085505 (2020a).
  • Sadigh et al. [2012] B. Sadigh, P. Erhart, A. Stukowski, A. Caro, E. Martinez, and L. Zepeda-Ruiz, Scalable parallel monte carlo algorithm for atomistic simulations of precipitation in alloys, Physical Review B 85, 184203 (2012).
  • Daw and Baskes [1984] M. S. Daw and M. I. Baskes, Embedded-atom method: Derivation and application to impurities, surfaces, and other defects in metals, Physical Review B 29, 6443 (1984).
  • Zhang et al. [2022a] Z. Zhang, J. Ding, and E. Ma, Shear transformations in metallic glasses without excessive and predefinable defects, Proceedings of the National Academy of Sciences 119, e2213941119 (2022a).
  • Plimpton et al. [2007] S. Plimpton, P. Crozier, and A. Thompson, Lammps-large-scale atomic/molecular massively parallel simulator, Sandia National Laboratories 18, 43 (2007).
  • Alvarez-Donado et al. [2019] R. Alvarez-Donado, S. Cajahuaringa, and A. Antonelli, Revisiting the fragile-to-strong crossover in metallic glass-forming liquids: Application to cux𝑥{}_{x}start_FLOATSUBSCRIPT italic_x end_FLOATSUBSCRIPT zrx𝑥{}_{x}start_FLOATSUBSCRIPT italic_x end_FLOATSUBSCRIPT al1002x1002𝑥{}_{100-2x}start_FLOATSUBSCRIPT 100 - 2 italic_x end_FLOATSUBSCRIPT alloy, Physical Review Materials 3, 085601 (2019).
  • de Koning et al. [1999] M. de Koning, A. Antonelli, and S. Yip, Optimized free-energy evaluation using a single reversible-scaling simulation, Physical Review Letters 83, 3973 (1999).
  • Paula Leite et al. [2016] R. Paula Leite, R. Freitas, R. Azevedo, and M. de Koning, The uhlenbeck-ford model: Exact virial coefficients and application as a reference system in fluid-phase free-energy calculations, The Journal of Chemical Physics 145 (2016).
  • Freitas et al. [2016] R. Freitas, M. Asta, and M. De Koning, Nonequilibrium free-energy calculation of solids using lammps, Computational Materials Science 112, 333 (2016).
  • Leite and de Koning [2019] R. P. Leite and M. de Koning, Nonequilibrium free-energy calculations of fluids using lammps, Computational Materials Science 159, 316 (2019).
  • Smith et al. [2017] H. L. Smith, C. W. Li, A. Hoff, G. R. Garrett, D. S. Kim, F. C. Yang, M. S. Lucas, T. Swan-Wood, J. Y. Lin, M. B. Stone, et al., Separating the configurational and vibrational entropy contributions in metallic glasses, Nature Physics 13, 900 (2017).
  • Dawson et al. [2011] K. Dawson, L. Zhu, L. A. Kopff, R. J. McMahon, L. Yu, and M. Ediger, Highly stable vapor-deposited glasses of four tris-naphthylbenzene isomers, The Journal of Physical Chemistry Letters 2, 2683 (2011).
  • Ramos et al. [2011] S. L. L. Ramos, M. Oguni, K. Ishii, and H. Nakayama, Character of devitrification, viewed from enthalpic paths, of the vapor-deposited ethylbenzene glasses, The Journal of Physical Chemistry B 115, 14327 (2011).
  • Maloney and Lemaitre [2006] C. E. Maloney and A. Lemaitre, Amorphous systems in athermal, quasistatic shear, Physical Review E 74, 016118 (2006).
  • Parmar et al. [2020b] A. D. Parmar, B. Guiselin, and L. Berthier, Stable glassy configurations of the kob–andersen model using swap monte carlo, The Journal of Chemical Physics 153, 134505 (2020b).
  • Bitzek et al. [2006] E. Bitzek, P. Koskinen, F. Gähler, M. Moseler, and P. Gumbsch, Structural relaxation made simple, Physical review letters 97, 170201 (2006).
  • [54] T. L. Cheung and S. C. H, Thermal and mechanical properties of cu–zr–al bulk metallic glasses, Journal of Alloys and Compou434-435 434–435, 71.
  • Zhang et al. [2022b] J. Zhang, Z. Zhou, Z. Zhang, M. Park, Q. Yu, Z. Li, J. Ma, A. Wang, H. Huang, M. Song, et al., Recent development of chemically complex metallic glasses: from accelerated compositional design, additive manufacturing to novel applications, Materials Futures 1, 012001 (2022b).
  • Bonfanti et al. [2023] S. Bonfanti, R. Guerra, R. Alvarez-Donado, P. Sobkowicz, S. Zapperi, and M. Alava, Quasi-localized modes in crystalline high entropy alloys, arXiv preprint arXiv:2303.09161  (2023).
  • Guiselin et al. [2022] B. Guiselin, C. Scalliet, and L. Berthier, Microscopic origin of excess wings in relaxation spectra of supercooled liquids, Nature Physics 18, 468 (2022).
  • Rickman and LeSar [2002] J. M. Rickman and R. LeSar, Free-energy calculations in materials research, Annual Review of Materials Research 32, 195 (2002).
  • Watanabe and Reinhardt [1990] M. Watanabe and W. P. Reinhardt, Phys. Rev. Lett. 65, 3301 (1990).
  • de Koning and Antonelli [1997] M. de Koning and A. Antonelli, Phys. Rev. B 55, 735 (1997).
  • de Koning et al. [2001] M. de Koning, A. Antonelli, and S. Yip, J. Chem. Phys. 115, 11025 (2001).
  • Miranda and Antonelli [2004] C. R. Miranda and A. Antonelli, J. Chem. Phys. 120, 11672 (2004).
  • Kirkwood [1935] J. G. Kirkwood, J. Chem. Phys. 3, 300 (1935).
  • Frenkel and Ladd [1984] D. Frenkel and A. J. Ladd, The Journal of chemical physics 81, 3188 (1984).
  • Paula Leite et al. [2017] R. Paula Leite, P. A. Santos-Flórez, and M. de Koning, Phys. Rev. E 96, 032115 (2017).