Abstract
Amorphous solids form an enormous and underutilized class of materials. In order to drive the discovery of new useful amorphous materials further we need to achieve a closer convergence between computational and experimental methods. In this review, we highlight some of the important gaps between computational simulations and experiments, discuss popular state-of-the-art computational techniques such as the Activation Relaxation Technique nouveau (ARTn) and Reverse Monte Carlo (RMC), and introduce more recent advances: machine learning interatomic potentials (MLIPs) and generative machine learning for simulations of amorphous matter (e.g., MAP). Examples are drawn from amorphous silicon and silica literature as well as from molecular glasses. Our outlook stresses the need for new computational methods to extend the time- and length-scales accessible through numerical simulations.
Keywords: Amorphous Materials, Computational methods, Activation Relaxation Technique nouveau, Machine Learning Interatomic Potentials, Deep Generative Models, Mesoscale Simulation, Inverse Design, Computational Material Science
Introduction
For over a century amorphous materials remained at the forefront of pure and applied research. From theoretical and computational perspective, at the forefront are the elusive physics of the glass transition,1 the quantification of hidden order in hyperuniform yet apparently random structures endowed with exotic properties,2 the interpretation of characterization experiments,3 the extension of computational simulations across multiple length- and time- scales for predictive modeling,4−7 and the development of machine learning approaches to help design useful amorphous materials. In this review, we present some of the challenges, the state-of-the-art computational approaches, and the opportunities that drive research farther into the amorphous chemical space.
Amorphous materials are typically defined as lacking long-range order, which means their atomic structure does not produce the well-defined diffraction patterns such as those seen in crystalline solids. This lack of a sharp, periodic arrangement results in indistinctive characteristics in many characterization experiments, such as broad diffuse scattering signals. However, this does not mean that amorphous structures are entirely without order. On a short-range scale, their atomic arrangements can display local order, with atoms adopting particular bonding configurations over a few atomic distances. This local order may vary depending on whether the material is in a liquid or solid state;8,9 the subset of amorphous structures in which the short-range order in the solid state is identical to that of the liquid phase is called glasses.
On the practical side, amorphous solids and glasses have numerous applications in biomedical engineering,10,11 sports equipment,12 energy conversion13,14 and even nuclear waste immobilization through vitrification.15 The number of potential glass compositions is estimated to be around 1052, with only about 105 compositions realized so far.16 The search for new compounds in this vast chemical space is enabled by sophisticated computational methods ranging from computational modeling of materials and their properties to informing search policies and hypothesis generation.
The main challenge in modeling of amorphous materials with atomistic resolution is the disconnect between lab-based and computer-based molecular structures which primarily arises due to the limited length- and time-scales accessible to atomistic simulation methods. On the technical level this is due to (1) the difficult statistical sampling in the rugged energy landscapes characteristic of amorphous matter, and (2) the limited physical validity of interatomic interaction models. In Part 1 of this review, we outline some of the problems that arise because of the mismatch between simulations and experiments and list the key factors that lead to the mismatch. In Part 2, we review the evolution of domain-relevant sampling methodologies over the past few decades, and finally, in Part 3, we discuss the most recent advances that propel the field forward by integrating machine learning into the simulation protocols. We conclude in Part 4 with a brief partial summary of the remaining challenges that make the field of simulation of amorphous materials particularly exciting today.
Part 1: Discord between Computational Simulations and Experiments
While massively parallel computing and, more recently, generative machine leaning models can, in principle, address the length-scale limitation of computational simulations, overcoming the time scale issue within the confines of molecular dynamics (MD) simulations— the workhorse of atomistic modeling—remains a significant challenge. Numerical methods can cover the first microsecond of reaction time, while experiments often take many orders of magnitude longer–seconds, minutes, hours. Consequently, various algorithms are employed to create computational models of amorphous materials, each exhibiting different structures and properties that often deviate significantly from those produced by experimental methods.
The isobaric enthalpy (H) versus temperature graph in the top panel of Figure 1(a) illustrates the difference in the stability of the resulting structures obtained in melt-and-quench protocols placing the simulated samples in a qualitatively higher enthalpic register. The inset illustrates this point using another measure of stability–the fictive temperature which is plotted versus cooling rate measured for silicate glasses in simulations and experiments (reproduced from ref.34) showing a large gap between the two.
Figure 1.
Summary of the existing simulation–experiment dichotomy in amorphous materials. (a) Mismatch between observable properties in simulated and real amorphous materials. (a) Top panel: Enthalpy H/Volume V as a function of temperature is schematically sketched (artist’s impression) for simulations (solid lines) and for experiments (dashed lines), illustrating isobaric cooling in a generic amorphous material as a function of temperature. Simulations that mimic experimental procedures typically employ cooling rates several order of magnitude faster (1014–109 K/s) than those possible in traditional melt quench experiments (102–100 K/s).17−20 The inset shows fictive temperature, as a measure of stability, plotted versus cooling rate measured for silicate glasses in simulations and experiments (Reprinted from ref (18), with the permission of AIP publishing). Bottom panel: Similarly, simulated vapor deposition rates (often >108 nm/s)21−26 outpace experimental rates which are usually below 100 nm/s.27−33 The figure shows the orientational order parameter, Sz, as a function of substrate temperature, observed in both simulations35 and experiments37 of TPD glasses (Reprinted from ref (35), with permissions of AIP publication). While simulations qualitatively capture the nonmonotonic behavior observed in experiments there are quantitative discrepancies between simulated and experimental trends. (b) Illustrations of common processing routes for amorphous materials. Quenching from melt: A molten material is rapidly cooled (quenched) below its melting point, preventing crystallization and resulting in an amorphous solid; Vapor deposition: Atoms or molecules in a gaseous state are deposited onto a cold substrate. The rapid solidification on the surface prevents the formation of crystalline structures. Spin coating: A solution is deposited on a spinning platform where its self-assembly is governed by the interplay between surface tension, centripetal forces, and evaporation process. Irradiation: A crystalline/ceramic material is exposed to a flux of high-energy particles (e.g., ion, neutrons) resulting in an induced structural transformation.
The bottom panel in Figure 1a shows the behavior of order parameter Sz that describes the average orientation of molecules within a thin film deposited via vapor deposition as a function of substrate temperature, observed in both simulations35−37 and experiments.37,38 At high substrate temperatures (Tsub > Tg, where Tg is the glass transition temperature), the molecules exhibit random orientations, resulting in Sz ≈ 0. As the substrate temperature decreases below Tg the molecules initially tend to align perpendicular to the substrate (Sz > 0). Upon further cooling, the molecular orientation transitions to a predominantly parallel alignment relative to the substrate (Sz < 0). While simulations qualitatively capture this nonmonotonic behavior observed in experiments, certain limitations exist. Since, deposition rate has been found to be an effective parameter in dictating the relaxation dynamics of vapor deposited glasses,39Tg estimated from simulations is generally higher than experimentally determined values.40 Such discrepancies arising from simpler models and limited time/length scales influence the quantitative agreement between simulated and experimental trends.
In Figure 1b we summarize schematically some of the typical processing routes starting from those that are reasonably approachable through simulations like melt-and-quench and vapor deposition to those that are notoriously difficult to model like spin-coating and irradiation. To illustrate the depth of the fundamental simulation-experiment discrepancy, consider for example that computational models of a-Si are routinely generated using melt-and-quench MD simulations, even though a-Si cannot be fabricated in the laboratory using a melt-and-quench process.
Specific Examples of Systems of Interest
Before we dive into the discussion of computational strategies, we first provide concrete context by introducing some of the commonly studied inorganic and organic amorphous solids, their applications, and the challenges associated with their preparation and simulation.
Inorganic Amorphous Solids
Amorphous silicon (a-Si) and amorphous silica (a-SiO2) are canonical examples of nonmetallic amorphous materials, with a-Si being a nonglassy amorphous solid and a-SiO2 being a glassy one. Over more than 60 years of extensive research, a broad consensus has emerged regarding many of their characteristics. Both materials are considered continuous random networks with few coordination defects, existing as metastable phases possessing a free energy higher than that of their crystalline counterparts. However, their exact structure and properties are highly dependent on their processing history. From a technological perspective, this variability in structure and properties is significant for established industries, such as hydrogenated a-Si photovoltaic panels fabricated by chemical vapor deposition (CVD),41 and critical for emerging industries, such as advanced photonics applications. Amorphous silicon, in particular, is being considered for use in photonic integrated circuits and advanced transistor devices because plasma-enhanced CVD a-Si deposition is compatible with other complementary metal–oxide–semiconductor fabrication steps, unlike traditional c-Si processing routes.42,43 For these applications, there is a direct relationship between processing conditions and electro-optical properties, making a-Si’s suitability for advanced applications highly dependent on its fabrication method.
Similarly, a-SiO2 exhibits variations when produced by different methods, such as fusing silica crystals versus formation through irradiation. Radiation-induced changes are particularly relevant for nuclear power plant aging and nuclear waste management. Irradiation can cause significant structural changes to silicate-based aggregates in concrete, leading to dimensional changes, alterations in chemical reactivity, and modifications of mechanical properties.44−48 The variability in processing routes also raises fundamental questions about how to define a reference “perfect” form of a-Si and a-SiO2. Different deposition conditions and thermal treatments can significantly affect the structure of a-Si.49−52 Even a-Si created by ion-implantation, which is often considered as a good reference because of its low porosity, can be arbitrarily relaxed by thermal treatment.53,54 Likewise, there is clear evidence that irradiation of fused (vitreous) silica causes substantial structural changes.55−57
Molecular Amorphous Solids
Beyond the traditional silicate-based compositions, when cooled from a molten state small organic molecules can transition into amorphous or glassy phases that are commonly known as molecular glasses or amorphous molecular materials.58 Molecular glasses have broad applications across various industries due to their unique properties. In pharmaceuticals, they enhance solubility and bioavailability of drugs.59−65 In electronics, they are utilized in technologies like OLEDs,66−72 organic photovoltaics73−75 and nonlinear optics76−81 offering flexibility and uniformity.
The preparation of molecular glasses primarily involves three techniques: Liquid-quenching methods, spin-coating, and physical vapor deposition (PVD). Liquid-quenching involves rapid cooling which prevents crystallization and results in a metastable glassy state.82−84 Spin coating is a technique used to deposit uniform layers of organic materials onto substrates by rapidly spinning them, ensuring precise control over film thickness and surface morphology.27,85−87 This process is widely employed in the fabrication of organic thin-film devices such as OLEDs,27,85,88−92 while PVD entails the evaporation or sublimation of a material in a vacuum chamber, where it condenses onto a substrate to form a glassy thin film.93 Interest in investigating the properties and applications of vapor-deposited molecular glasses stems from their superior qualities compared to conventional glasses formed through solution processing,27 notably their ultrastability.21,28−31,94 Vapor-deposited molecular glasses offer higher density,95−97 unique phase transitions,95,98,99 improved mechanical properties100−102 and anisotropy.27,31,35,37,71,87,91,103−124 Anisotropy in these glasses is evident in their optical birefringence,27,37,106,108,113,117,118 magnetic behavior,103 and structural characteristics.31,87,91,104,105,109,111,113,114
Understanding the interplay between structural anisotropy and glass stability is critical, as it is influenced by molecular structure and deposition conditions. Structural anisotropy in PVD glasses is largely attributed to preferred molecular orientation31,109,125 and molecular layering.105,111,115 At low deposition temperatures (Tdep), molecules tend to orient horizontally (parallel to the substrate).31 At intermediate Tdep, elongated molecules typically orient vertically (perpendicular to the substrate).113 Deposition near the glass transition temperature (Tg), at slow deposition rates, or with molecules possessing smaller aspect ratios, tends to result in isotropic packing.122,123 Both experimental and simulation studies indicate that this orientational anisotropy originates from the structure of the supercooled liquid near the surface35,37,113,123 and stable glasses are formed through surface-mediated equilibrium during PVD,28,126−131 where molecules near the surface have enhanced mobility allowing them to sample a large number of configurations in search of a more stable state. Thereby, the properties of these glasses can be controlled by adjusting deposition parameters.38,71,122,132
Computational Challenges
Understanding the relationship between processing conditions and structural properties is essential for optimizing the performance of glassy organic and inorganic films in various technological applications. While computational models based on Molecular Dynamics (MD) and Monte Carlo (MC) simulations offer valuable insights into the properties and behavior of molecular glasses,4,21−23,133 there remains a vast gap between these models and real-world laboratory conditions for glass preparation, see Table 1 for the summary of discrepancies in system size and accessible time scales in simulation and experiments.
Table 1. Comparative Analysis of Key Aspects between Simulations and Experiments in the Study of Molecular Glasses.
Aspect | Simulations | Experiments |
---|---|---|
System size | 102–106 molecules | ∼1023 molecules |
Time scale | 10–12–10–6 s | >10–105 s |
Deposition rate | 108–1011 nm/s | 10–2–10–1 nm/s |
Cooling rate | 109–1014 K/s | 100–102 K/s |
Film thickness | <10 nm | 102–104 nm |
Most notably, glasses simulations remain plagued by a series of challenges:
(i) System size. Due to high computational costs, a significant disparity exists between the system sizes accessible through simulations and those investigated experimentally. Glass simulations are typically limited to relatively small scales, involving systems that range from a few hundred to several thousand molecules/particles.133,134 In contrast, experimental studies can probe much larger systems, often extending to macroscopic dimensions encompassing 1023 molecules/particles or more. This difference highlights a key challenge: simulations, while providing detailed atomic-level insights, are constrained in capturing the full complexity and variability of real glassy materials, which can limit their applicability and accuracy in replicating experimental conditions.134 Other challenges include insufficient statistical sampling with too few molecules,135,136 enhanced thermodynamic fluctuations (which scale inversely with the square root of particle count),137,138 and errors from limited simulation box sizes that fail to capture large structural features.139,140 To balance accuracy and computational feasibility, the system size must be optimized to minimize finite-size effects while maintaining reasonable computational demands.
(ii) Limited accessible time scales. MD simulations are constrained to short time scales, typically up to a few microseconds,17 whereas experimental processes can span for hours or days.28,31,100 This disparity results in MD simulations employing ultrafast cooling rates (1014–109 K/s), far exceeding those in conventional experiments (102–10° K/s).17−20 Consequently, glasses prepared via MD tend to have higher fictive temperatures, making them less stable than experimentally synthesized glasses.20
For glasses prepared via melt and quench route, the rapid cooling leads to glasses that are less stable and more disordered, remaining in high-energy states. This significant difference in cooling rates between simulations and experiments creates a systematic gap, leading to less stable simulated glasses compared to their experimental counterparts.
On the other hand, for PVD glasses, deposition rates in simulations are also significantly higher (often higher than 108 nm/s)21−26 compared to experimental rates (typically less than 100 nm/s).27−33 This discrepancy means that simulated films tend to have different microstructures and properties compared to experimentally deposited films.23,35 These differences in deposition rates further exacerbate the divergence in film thickness (typically, in simulations, film thicknesses are generally less than 10 nm,21,24,25 whereas in experiments, they range from hundreds of nanometers29,32,33 to micrometers,31,39,76,79,100,101 see Table 1) and structural properties between simulations and experiments, as the rapid deposition does not allow for the same relaxation and ordering processes that occur in experimental conditions.39
Going beyond mimicking experimental protocols, and adding to the challenge, amorphous structures, with their intricate details influenced by preparation methods, have long posed fundamental questions regarding their most relaxed configuration and associated inherent features. These may or may not correspond to the structural properties of the optimal continuous random network, which is defined as the lowest strain perfectly coordinated large-scale network that can exist. While such structure is, by definition, elusive, it has oriented modeling efforts even when looking at specific materials. For example, early attempts to manually create amorphous silicon transitioned to sophisticated computational approaches, such as the bond-switching method by Wooten, Winer, and Weaire (WWW) for producing as close to perfectly coordinated optimal continuous random network compatible with silicon local bonding environment. Other significant techniques discussed in detail in Part 2 of this review, such as the Activation-Relaxation Technique nouveau (ARTn), Reverse Monte Carlo (RMC), swap Monte Carlo (SMC) have partially advanced our understanding of the relaxation mechanisms and energy landscapes of these materials measuring success both by matching experimental results and closeness to the best available CRNs.
(iii) Lack of accurate interatomic interaction potentials. Finally, the accuracy of interatomic interaction models remains limited. The electronic Density Functional Theory (DFT), a class of ab initio methods, is commonly regarded as the gold standard for modeling the potential energy surfaces (PESs) of amorphous materials. However, even at the DFT level, the use of different exchange-correlation functionals can result in significantly different PESs, leading to models with varying structures and properties. Moreover, exploring the DFT PES is computationally expensive, severely limiting the accessible length and time scales. As a result, classical interatomic interaction potentials are often employed, introducing further discrepancies between computational models and real-world amorphous materials.
Empirical force fields involve optimizing hundreds or thousands of parameters.141,142 Even simpler classical force fields require dozens of parameters, increasing with the number of elements involved.143 This high dimensionality complicates the identification of what would be the optimal structures based on the values in the cost function landscape of a continuous random network or the best match to experimental results.155 In addition, the cost function landscape for force field optimization is typically rough and filled with numerous local minima.144 As a result, traditional optimization methods, such as gradient descent, often get trapped in these local minima, making the process highly dependent on the initial starting point.145 This roughness necessitates numerous independent optimizations and relies heavily on intuition, making the parametrization process biased and laborious.143,146
In the past decade, machine-learning (ML) potentials trained on DFT data have helped narrow this gap, see Part 3 in this review for more details. However, challenges remain, particularly concerning interfaces, chemical changes, and charge transfer.
Part 2: Beyond Mimicking Experimental Protocols: Sampling and the Search for Optimal Structure
As mentioned before, while the structural details of amorphous configurations depend on the preparation procedures, fundamental questions arose very early on as to the nature and existence of the optimal continuous-random network (CRN) structure, its features, and its generality in representing real systems. Representing thermodynamic ensembles, such structures could be produced through nondynamical processes driven only by energy or structural deformation (bond length, bond angle and coordination, for example).
If for materials such as silica, a glass characterized by significant topological rigidity near the rigidity transition threshold,146 even short molecular dynamical procedures can lead to low-stress and low-defect structures.147 Finding low-energy structures in high-connectivity structures is a more significant challenge that amounts to overcoming barriers in a featured free energy landscape, see Figure 2a for an illustration. For a pedagogical review on rigidity and mechanical stability of amorphous solids see ref.148
Figure 2.
Illustration of computational methodologies used in sampling and potential energy modeling. (a) Featured free energy landscape of amorphous materials requires specialized enhanced sampling techniques such as ARTn, Monte Carlo, and, more recently, approaches based on generative artificial intelligence (AI). (b) Flow-chart of development of machine learning interatomic potentials (MLIPs) using ab initio methods, molecular dynamics (MD) simulations, and active learning (AL).
While the reader is referred to the recent review by Laurent J. Lewis149 for an extensive historical account of modeling amorphous silicon, we focus here on providing a high-level view of the various approaches used to generate high-quality disordered structures beyond the generic quench and melt. In the case of glasses, we similarly refer the reader to the review of Michoulaut and Bauchy, which describes how rigidity theory can be employed to study and generate glassy silicate models.150
Initial attempts to create such structures for amorphous silicon were done “by hand” by Polk151 and a few others in early 1970s. This was followed soon after by computer approaches. Among the early success is the ingenious bond-switching approach proposed by Wooten, Winer and Weaire (WWW)152 to create a disordered network starting from a perfect crystal while preventing the creation of coordination defects using a harmonic Keating potential.153
The WWW algorithm was used, in its original form,154 and revised versions that greatly accelerated the sampling allowing the generation of high-quality model of 4000155 to 100,000 atom models that have remained reference models until today.156 These methods were also used, iterating with relaxations using a modified Stillinger-Weber potential157 adapted to reproduce amorphous silicon structures,158 to generate near hyperuniform continuous-random network models.159 While the original WWW algorithm does not impose evenness on the loops connecting atoms, it is possible to modify the method to force even cycles and explore the effects of additional chemical ordering on the structure of binary amorphous semiconductors such as a-GaAs,160 showing the richness of this bond-switching approach to explore fundamental questions about the nature of CRNs. And while the general WWW approach produces the lowest energy structures without coordination defects, the bond-switching moves require crossing high energy barriers that are unphysical. Moreover, experimental evidence even for well-relaxed a-Si shows a significant concentration of low-coordination defects161 that need to be reproduced by modeling.
In general, defects in amorphous materials include the deviations from a perfect bonding configuration,162 dangling or floating bonds,163 and vacancies,164 local strain-related defects165 may modify physical and chemical properties of the material.166,167 In the past computational studies of defects explored the topological satisfaction of local structures using the atomic-level stresses164,168 identified locally favored structures,169 and examined local vibrational modes.170,171
Two general approaches were applied to a-Si to address this issue: Reverse Monte Carlo and the Activation-Relaxation Technique. Reverse Monte Carlo (RMC), first applied to a-Si in 1993,172 aimed at extracting local atomic structure from global experimental averages such as the radial distribution function, using the minimum number of additional constraints. Unlike MD simulations, which depend on interatomic potentials and attempt to replicate real-time dynamics, RMC aims to find atomic configurations that match experimental data without accounting for atomic interactions or thermodynamics.142 In RMC, atomic positions are randomly adjusted to minimize a cost function representing the difference between the model’s pair correlation function and the experimental data. This process follows an accept-reject protocol, similar to the Metropolis Monte Carlo algorithm, but with a key distinction: while Metropolis Monte Carlo uses potential energy as the cost function, RMC uses the discrepancy between the simulated and experimental signals173,174 thereby RMC avoids the high cooling rates and approximations associated with MD’s force fields by bypassing the need for melt-quenching. However, its reliance on experimental data limits its predictive power, as it can only replicate structures for conditions that have already been explored.130
A significant challenge in RMC is that it generates atomic configurations by inverting experimental data, which often results in nonunique solutions. Different atomic arrangements can produce similar or indistinguishable experimental fingerprints, such as g(r).143,144 This ambiguity is especially problematic for complex materials like glasses, where multiple valid yet potentially thermodynamically unstable structures may satisfy the same experimental constraints.130,143,145 Through multiple trials, it was discovered that the range of configurations able to reproduce macroscopically averaged experimental data included a large fraction of nonphysical configurations and that the inclusion of strict constraints of local configurations was necessary to generate models close to those obtained by quench and melt.175
Further improving on these methods, Drabold and collaborators introduced the “Force-Enhanced Atomic Refinement” (FEAR) method to improve RMC results by recursively optimizing the structure against local classical force field-based176 or quantum-mechanical DFT-based177 potentials and global experimental results.178 While the final configuration meets both constraints, often classical mechanics fails to provide adequate accuracy, and resorting to quantum-mechanical approaches limits the system sizes to only a few hundred atoms.179 This limits greatly the advantage of this method for generating relevant samples given the local configurational richness of disordered systems such as a-Si. RMC and related techniques are further discussed in Part 3 in the context of using experimental data to inform simulations.
The Activation-Relaxation Technique nouveau (ARTn), an open-ended method for finding local transition states surrounding a local minimum,180,181 was used to explore the energy landscape, identifying the relaxation mechanisms, and relax the structure of amorphous materials, including a-Si,182,183 a-GaAs184 and silica glass.185 As with molecular dynamics, the resulting low-energy structures correspond to low-energy points of the potential energy used. The validity of the result is therefore determined by the quality of the potential. ARTn, just as the WWW algorithm, does not describe real kinetics, but rather generates activated mechanisms that can be accepted or rejected using, for example, a Metropolis criterion. However, it allows for the generation of a much broader set of mechanisms186 that provide a better understanding the relaxation mechanisms and provides different pathways to low energy structures, an essential feature to assess the universal properties of optimal CRNs.
In the context of glass transition studies4 that aim to explore relaxation processes in increasingly viscous substances, Swap Monte Carlo (SMC) pioneered by Grigera and Parisi187 has become popular as it speeds up relaxation by introducing nonphysical moves that involve swapping physically distant particles. Enhanced efficiency is usually attributed to overcoming kinetic barriers typical in glass formers,188 but it also comes with some disadvantages: it has been shown that SMC leads to artificial crystallization in some systems.189,190 For instance, Brumer and Reichman found that while SMC was efficient for a two-dimensional hard disk system, 3D polydisperse systems were prone to phase separation/crystallization.189 Formation of such ordered phases is an unwanted artifact because the resulting structures no longer represent the glassy state of interest. A breakthrough was achieved when Ninarello et al. demonstrated that by carefully tuning the interaction potentials and polydispersity, a set of glass formers can be studied, with speed up thermalization up to 10 orders of magnitude without crystallization.191 SMC has since been successfully applied in numerous studies such as measuring the static length scales,192,193 and generating ultrastable glasses.194,195
Although, as discussed in ref.,191 SMC’s effectiveness is quite restricted, in recent years, adaptive methods augmented by machine learning protocols, such as reinforcement learning and normalizing flows have shown that proposal distributions in Metropolis-Hasting algorithm can enhance the sampling efficiency particularly in systems where the equilibration suffers from slow dynamics and metastability.196,197 Thanks to the ongoing intensive method development in the MC community, for example ref.198 we expect to see more novel designs of MC methods accompanied by “smart” moves in context of amorphous materials and glasses in near future.
Part 3: New Ideas: Machine Learning for Improved Accuracy, Sampling, and Reduced Mismatch with Experiments
Machine Learning Interatomic Potentials
Atomistic simulations, powered by machine learning interatomic interaction models, revolutionize materials science in unprecedented ways199,200 lifting many barriers in modeling. None of these barriers, from the spatiotemporal limitation of density function theory to the transferability limitations of semiempirical models pose a problem for machine learning interatomic potentials (MLIP). While several MLIP models have been developed since the first one by Bheler et al. in 2007201 they all rely on approximations of the potential energy surface (PES). This approximation assumes a medium without charge or polarization, allowing the total energy of a given system to be approximated by the sum of individual atomic energies. These individual energies strongly depend on the local atomic environment, which is captured by descriptors representing the surrounding atomic configurations.202
This concept, in principle, is based on atomic environment descriptors, regression methods, and quantum mechanical data.203,204 Except for the symmetry of the Hamiltonian, there is no direct parametrization based on the type of physical interactions. Instead, it relies on mathematical interpolation of the potential energy surface (PES), utilizing quantum mechanical data that constitutes a set of discrete points on the PES. Thus, the accuracy of MLIP depends on how well these discrete points cover the targeted PES, a problem that is often addressed by resorting to active learning protocols.
The flow-chart in Figure 2b outlines a typical MLIP training protocol. MLIPs differ based on the type of local atomic environment descriptors and regressors used. Current regression methods can be grouped into three main categories: artificial neural networks (NNs),201,205 kernel-based method,206 and linear regression.207,208 The first two categories have been employed in modeling amorphous silicon (a-Si) and amorphous silica (a-SiO2) using neural network potentials and Gaussian approximation potentials.209−213 Linear regression potentials include the Spectral Neighbor Analysis Potential (SNAP),207 atomic cluster expansion214 and the Moment Tensor Potential (MTP).208 These have recently been leveraged in order to create MLIPs able to jointly describe Si, SiO2 and their interfaces,213,215,216 in solid, liquid and amorphous states, with a very good level of fidelity.
Sampling
Approaches based on machine learning are enriching the search for both experimentally relevant and optimal structures. As mentioned above, progress for the former, was made with the development of DFT-quality machine-learned potentials that made it possible to generate well-relaxed large amorphous models with low defects using quench and melt approaches.206,217 Coupled with recent large-scale melt-and-quench work using the modified Stilliger-Weber potential,218 this work demonstrates that with the right potential and sufficient computational effort, it is possible to generate structure that, while presenting some coordination defects, are comparable in structure to the best continuous random networks (CRNs) generated using the Wooten, Winer and Weaire (WWW) approach.
Nevertheless, fundamental questions remain: the minimum strain structure that can be generated - how does it vary with size? After all the work needed to relax a structure to the same level as WWW increases faster than the number of atoms, as seen in ref.,155 for example; what of binary network, etc.? To answer these questions, Comin and Lewis developed a machine-learning approach to directly generate a-Si structures after learning from high-quality models.219 If the initial results are still far from optimal, they show that this task is possible with sufficient training data and the right ML model. Indeed, in recent years a variety of generative models have been explored, including GANs refs,220,221 autoencoders,219 normalizing flows.222 Following the work of Comin and Lewis, GAN models have been used to generate amorphous structures based on point cloud representations of molecular input.221
The fast-decaying structural correlations characteristic of amorphous materials suggest that autoregressive generation may be an effective strategy for sampling large scale amorphous configurations. In an autoregressive approach, the probability of transitions from one microstate to others is inferred from small-scale (order of correlation length) samples of the material. Then the larger sample is extrapolated from small samples, and it is generated one grid-point at a time conditional on previously generated molecular context. The cost of sampling is thereby limited to linear scaling with the number of populated grid points. Recent works in this area developed a modeling approach called the Morphological Autoregressive Protocol (MAP) based on the PixelCNN architecture223,224 with grid-based representation of molecular structures showing promising results on systems like amorphous graphene (2D), and liquid water (3D). Grid-based input representation offers the benefits of easy processing and, in the case of 2-dimensional films, the possibility of direct integration of experimental microscopy data into the modeling loop; but comes with high memory demands. As an alternative, point cloud representations offer scalability but require careful design to ensure symmetry invariance.225,226
Informing Simulations Using Experimental Data
The gap between simulation and experiment may be reduced by integrating experimentally measured parameters into simulations. This may be done by including experimental data as simulation parameters176,178,227 or as constraints.228−230 For instance, molecular dynamics simulations of amorphous silicon (a-Si) and amorphous silica (a-SiO2) are initialized using experimental structural data such as atomic coordinates and lattice parameters.228−231 These are usually determined experimentally via techniques such as X-ray crystallography232,233 and neutron diffraction.234,235 Furthermore, a cubic box corresponding to the experimental density of amorphous silica (2.20 g/cm3) was used to prepare the amorphous silica components of the MLIP databases.142,236,237
Experimental data obtained from diffraction, infrared (IR), and nuclear magnetic resonance (NMR) measurements is used constraints to minimize a cost function during the simulation of a-Si and a-SiO2 within the framework of Reverse Monte Carlo (RMC)238,239 and related Hybrid Reverse Monte Carlo (HRMC).240,241 For example, experimental structure factors, Si–Si–Si bond angles, and density data were used as constraints to simulate amorphous silicon with the standard Reverse Monte Carlo (RMC) method.242 In addition, realistic amorphous silica structures may be generated using the RMC method by applying experimental constraints such as Si–O bonds, intratetrahedral (O–Si–O) bond angles, and intertetrahedral (Si–O–Si) bond angles.243 The HRMC method encompasses, but is not limited to, Experimentally Constrained Molecular Relaxation (ECMR),244 Experimentally Constrained Structural Relaxation (ECSR),245 and Force-Enhanced Atomic Refinement (FEAR).176 Both a-Si and a-SiO2 were successfully simulated with high fidelity within these frameworks. For instance, in ECSR, the experimental reduced electron diffraction intensities and the experimental fluctuation electron microscopy FEM variance data were used.245 The FEAR method was successfully applied to both a-Si and a-SiO2 using the pair distribution function obtained from neutron diffraction and X-ray diffraction data as experimental constraint.176 These approaches may be applied to molecular glasses, but molecular glasses remain at earlier stages of computational exploration than amorphous silica and silicon.58
Part 4: Conclusions and Outlook
In this review, we examined the pivotal role of computational simulations in bridging the gap between theoretical predictions and experimental observations. We highlighted significant advancements in modeling amorphous materials, including a-Si, a-SiO2, and molecular glasses, while acknowledging the ongoing challenges related to matching the experimentally relevant system sizes and time scales.
Until recently computational simulations of amorphous materials reflected a choice of either accepting artifacts that arise due to inadequately small size of the simulated systems studied on time scales that are too short using methods that are too expensive to scale up or accepting the artifacts of the inaccuracies that come with reliance on force-fields/purely data-driven techniques. In this review we explored the progress in conformation sampling methods, from techniques that mimic experimental protocols such as the traditional quenching and melting and vapor deposition techniques to more general approaches such as Reverse Monte Carlo (RMC) and the Activation-Relaxation Technique nouveau (ARTn), to the novel machine learning-based approaches developed specifically for amorphous materials, including generative models and autoregressive techniques such as the Morphological Autoregressive Protocol (MAP). Finally, we have discussed a recent development that is of particular importance to the field of amorphous matters–the emergence of machine learning interatomic potentials (MLIPs) that bring DFT-level accuracy to atomistic simulations at a reduced cost relative to DFT, augmenting simultaneously the accuracy and sampling efficiency.
The advances discussed in this review bring us closer to the ultimate goal of informing experiments using computational simulations and guiding and accelerating the design of useful amorphous materials. New ideas, however, are still needed to completely close the gap in time- and length- scales between simulations and experiments. The exciting promise of this field is that by constructing new methods we may unlock an incomprehensively enormous set of new chemical species. And while it may not yet be widely appreciated, to us the future of materials appears to be delightfully amorphous.
Acknowledgments
A.M., E.A., and L.S. acknowledge funding from the National Research Council via the AI4Design program, NSERC Alliance Grants - Consortia Quantum “QuantaMole”, and IVADO. L.B. and K.Z. acknowledge support from NSERC and the Nuclear Waste Management Organization of Canada. N.M. acknowledges partial support from a NSERC Discovery Grant.
Author Contributions
The manuscript was written through contributions of all authors and all authors have given approval to the final version of the manuscript.
National Research Council via the AI4Design program, NSERC Alliance Grants - Consortia Quantum “QuantaMole” and NSERC Discovery Grant, IVADO, Nuclear Waste Management Organization of Canada.
The authors declare no competing financial interest.
References
- Berthier L.; Biroli G. Theoretical perspective on the glass transition and amorphous materials. Rev. Mod. Phys. 2011, 83 (2), 587–645. 10.1103/RevModPhys.83.587. [DOI] [Google Scholar]
- Torquato S. Hyperuniformity and its generalizations. Phys. Rev. E 2016, 94 (2–1), 022122. 10.1103/PhysRevE.94.022122. [DOI] [PubMed] [Google Scholar]
- Montoya-Castillo A.; Chen M. S.; Raj S. L.; Jung K. A.; Kjaer K. S.; Morawietz T.; Gaffney K. J.; van Driel T. B.; Markland T. E. Optically Induced Anisotropy in Time-Resolved Scattering: Imaging Molecular-Scale Structure and Dynamics in Disordered Media with Experiment and Theory. Phys. Rev. Lett. 2022, 129 (5), 056001. 10.1103/PhysRevLett.129.056001. [DOI] [PubMed] [Google Scholar]
- Berthier L.; Reichman D. R. Modern computational studies of the glass transition. Nat. Rev. Phys. 2023, 5 (2), 102–116. 10.1038/s42254-022-00548-x. [DOI] [Google Scholar]
- Schleder G. R.; Padilha A. C. M.; Acosta C. M.; Costa M.; Fazzio A. From DFT to machine learning: recent approaches to materials science–a review. J. Phys.: Materials 2019, 2 (3), 032001. 10.1088/2515-7639/ab084b. [DOI] [Google Scholar]
- Salahub D. R. Multiscale molecular modelling: from electronic structure to dynamics of nanosystems and beyond. Phys. Chem. Chem. Phys. 2022, 24 (16), 9051–9081. 10.1039/D1CP05928A. [DOI] [PubMed] [Google Scholar]
- Izvekov S.; Voth G. A. A multiscale coarse-graining method for biomolecular systems. J. Phys. Chem. B 2005, 109 (7), 2469–73. 10.1021/jp044629q. [DOI] [PubMed] [Google Scholar]
- Gupta P. K. Non-crystalline solids: glasses and amorphous solids. J. Non-Cryst. Solids 1996, 195 (1–2), 158–164. 10.1016/0022-3093(95)00502-1. [DOI] [Google Scholar]
- Filipponi A.; Di Cicco A. Short-range order in crystalline, amorphous, liquid, and supercooled germanium probed by x-ray-absorption spectroscopy. Phys. Rev. B 1995, 51 (18), 12322–12336. 10.1103/PhysRevB.51.12322. [DOI] [PubMed] [Google Scholar]
- Hua N.; Huang L.; Chen W.; He W.; Zhang T. Biocompatible Ni-free Zr-based bulk metallic glasses with high-Zr-content: compositional optimization for potential biomedical applications. Mater. Sci. Eng. C. Mater. Biol. Appl. 2014, 44, 400–10. 10.1016/j.msec.2014.08.049. [DOI] [PubMed] [Google Scholar]
- Kinser E. R.; et al. Nanopatterned Bulk Metallic Glass Biosensors. ACS Sens 2017, 2 (12), 1779–1787. 10.1021/acssensors.7b00455. [DOI] [PubMed] [Google Scholar]
- Khun N. W.; Yu H.; Chong Z. Z.; Tian P.; Tian Y.; Tor S. B.; Liu E. Mechanical and tribological properties of Zr-based bulk metallic glass for sports applications. Mater. Des. 2016, 92, 667–673. 10.1016/j.matdes.2015.12.050. [DOI] [Google Scholar]
- Tarigan H. J.; Kahler N.; Ramos N. S.; Kumar G.; Bernussi A. A. Low reflectance of nano-patterned Pt-Cu-Ni-P bulk metallic glass. Appl. Phys. Lett. 2015, 107 (2), 021903. 10.1063/1.4926873. [DOI] [Google Scholar]
- Uzun C.; Meduri C.; Kahler N.; de Peralta L. G.; McCollum J.M.; Pantoya M.; Kumar G.; Bernussi A.A. Photoinduced heat conversion enhancement of metallic glass nanowire arrays. J. Appl. Phys. 2019, 125 (1), 094306. 10.1063/1.5059423. [DOI] [Google Scholar]
- McCloy J. S.; Goel A. Glass-ceramics for nuclear-waste immobilization. MRS Bull. 2017, 42 (03), 233–240. 10.1557/mrs.2017.8. [DOI] [Google Scholar]
- Zanotto E. D.; Coutinho F. A. B. How many non-crystalline solids can be made from all the elements of the periodic table?. J. Non-Cryst. Solids 2004, 347 (1–3), 285–288. 10.1016/j.jnoncrysol.2004.07.081. [DOI] [Google Scholar]
- Lane J. M. Cooling rate and stress relaxation in silica melts and glasses via microsecond molecular dynamics. Phys. Rev. E Stat. Nonlin. Soft. Matter. Phys. 2015, 92 (1), 012320. 10.1103/PhysRevE.92.012320. [DOI] [PubMed] [Google Scholar]
- Li X.; et al. Cooling rate effects in sodium silicate glasses: Bridging the gap between molecular dynamics simulations and experiments. J. Chem. Phys. 2017, 147 (7), 074501. 10.1063/1.4998611. [DOI] [PubMed] [Google Scholar]
- Vollmayr K.; Kob W.; Binder K. Cooling-rate effects in amorphous silica: A computer-simulation study. Phys. Rev. B Condens. Matter. 1996, 54 (22), 15808–15827. 10.1103/PhysRevB.54.15808. [DOI] [PubMed] [Google Scholar]
- Vollmayr K.; Kob W.; Binder K. How do the properties of a glass depend on the cooling rate? A computer simulation study of a Lennard-Jones system. J. Chem. Phys. 1996, 105 (11), 4714–4728. 10.1063/1.472326. [DOI] [Google Scholar]
- Leoni F.; Martelli F.; Royall C.P.; Russo J. Structural Signatures of Ultrastability in a Deposited Glassformer. Phys. Rev. Lett. 2023, 130 (19), 198201. 10.1103/PhysRevLett.130.198201. [DOI] [PubMed] [Google Scholar]
- Yoo D.; Song H.; Youn Y.; Jeon S. H.; Cho Y.; Han S. A molecular dynamics study on the interface morphology of vapor-deposited amorphous organic thin films. Phys. Chem. Chem. Phys. 2019, 21 (3), 1484–1490. 10.1039/C8CP05294H. [DOI] [PubMed] [Google Scholar]
- Degitz C.; Konrad M.; Kaiser S.; Wenzel W. Simulating the growth of amorphous organic thin films. Org. Electron. 2022, 102, 106439. 10.1016/j.orgel.2022.106439. [DOI] [Google Scholar]
- Reid D. R.; Lyubimov I.; Ediger M. D.; de Pablo J. J. Age and structure of a model vapour-deposited glass. Nat. Commun. 2016, 7, 13062. 10.1038/ncomms13062. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Muccioli L.; D’Avino G.; Zannoni C. Simulation of vapor-phase deposition and growth of a pentacene thin film on C60 (001). Adv. Mater. 2011, 23 (39), 4532–4536. 10.1002/adma.201101652. [DOI] [PubMed] [Google Scholar]
- Grigoriev F. V.; Sulimov V.B. Atomistic Simulation of Physical Vapor Deposition of Optical Thin Films. Nanomaterials 2023, 13 (11), 1717. 10.3390/nano13111717. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shibata M.; Sakai Y.; Yokoyama D. Advantages and disadvantages of vacuum-deposited and spin-coated amorphous organic semiconductor films for organic light-emitting diodes. J. Mater. Chem. C 2015, 3 (42), 11178–11191. 10.1039/C5TC01911G. [DOI] [Google Scholar]
- Swallen S. F.; Kearns K. L.; Mapes M. K.; Kim Y. S.; McMahon R. J.; Ediger M. D.; Wu T.; Yu L.; Satija S. Organic Glasses with Exceptional Thermodynamic and Kinetic Stability. Science 2007, 315 (5810), 353–356. 10.1126/science.1135795. [DOI] [PubMed] [Google Scholar]
- Qiu Y.; Dalal S. S.; Ediger M. D. Vapor-deposited organic glasses exhibit enhanced stability against photodegradation. Soft Matter. 2018, 14 (15), 2827–2834. 10.1039/C8SM00183A. [DOI] [PubMed] [Google Scholar]
- Rodriguez-Tinoco C.; Gonzalez-Silveira M.; Ramos M. A.; Rodriguez-Viejo J. Ultrastable glasses: new perspectives for an old problem. Nuovo Cimento 2022, 45 (5), 325–406. 10.1007/s40766-022-00029-y. [DOI] [Google Scholar]
- Cheng S.; Lee Y.; Yu J.; Yu L.; Ediger M. D. Generic Behavior of Ultrastability and Anisotropic Molecular Packing in Codeposited Organic Semiconductor Glass Mixtures. Chem. Mater. 2024, 36 (7), 3205–3214. 10.1021/acs.chemmater.3c02935. [DOI] [Google Scholar]
- Sakai Y.; Shibata M.; Yokoyama D. Simple model-free estimation of orientation order parameters of vacuum-deposited and spin-coated amorphous films used in organic light-emitting diodes. Appl. Phys. Express 2015, 8 (9), 096601. 10.7567/APEX.8.096601. [DOI] [Google Scholar]
- Bishop C.; Bagchi K.; Toney M. F.; Ediger M. D. Vapor deposition rate modifies anisotropic glassy structure of an anthracene-based organic semiconductor. J. Chem. Phys. 2022, 156 (1), 014504. 10.1063/5.0074092. [DOI] [PubMed] [Google Scholar]
- Vaills Y.; Qu T.; Micoulaut M.; Chaimbault F.; Boolchand P. Direct evidence of rigidity loss and self-organization in silicate glasses. J. Phys.: Condens. Matter 2005, 17 (32), 4889–4896. 10.1088/0953-8984/17/32/003. [DOI] [Google Scholar]
- Lyubimov I.; Antony L.; Walters D. M.; Rodney D.; Ediger M. D.; de Pablo J. J. Orientational anisotropy in simulated vapor-deposited molecular glasses. J. Chem. Phys. 2015, 143 (9), 094502. 10.1063/1.4928523. [DOI] [PubMed] [Google Scholar]
- Jiang J.; Walters D. M.; Zhou D.; Ediger M. D. Substrate temperature controls molecular orientation in two-component vapor-deposited glasses. Soft Matter 2016, 12 (13), 3265–70. 10.1039/C6SM00262E. [DOI] [PubMed] [Google Scholar]
- Dalal S. S.; Walters D. M.; Lyubimov I.; de Pablo J. J.; Ediger M. D. Tunable molecular orientation and elevated thermal stability of vapor-deposited organic semiconductors. Proc. Natl. Acad. Sci. U. S. A. 2015, 112 (14), 4227–32. 10.1073/pnas.1421042112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bishop C.; Chen Z.; Toney M. F.; Bock H.; Yu L.; Ediger M. D. Using Deposition Rate and Substrate Temperature to Manipulate Liquid Crystal-Like Order in a Vapor-Deposited Hexagonal Columnar Glass. J. Phys. Chem. B 2021, 125 (10), 2761–2770. 10.1021/acs.jpcb.0c11564. [DOI] [PubMed] [Google Scholar]
- Zhang A.; Moore A. R.; Zhao H.; Govind S.; Wolf S. E.; Jin Y.; Walsh P. J.; Riggleman R. A.; Fakhraai Z. The role of intramolecular relaxations on the structure and stability of vapor-deposited glasses. J. Chem. Phys. 2022, 156 (24), 244703. 10.1063/5.0087600. [DOI] [PubMed] [Google Scholar]
- Mondal A. Molecular library of OLED host materials—Evaluating the multiscale simulation workflow. Chem. Phys. Rev. 2021, 2 (3), 031304. 10.1063/5.0049513. [DOI] [Google Scholar]
- I. Kabir M.; Ibarahim Z.; Sopian K.; Amin N. A Review on Progress of Amorphous and Microcrystalline Silicon Thin-Film Solar Cells. Recent Patents on Electrical Engineeringe 2011, 4 (1), 50–62. 10.2174/1874476111104010050. [DOI] [Google Scholar]
- Lipka T.; Moldenhauer L.; Müller J.; Trieu H. K. Photonic integrated circuit components based on amorphous silicon-on-insulator technology. Photonics Res. 2016, 4 (3), 126. 10.1364/PRJ.4.000126. [DOI] [Google Scholar]
- Della Corte F. G.; Rao S. Use of Amorphous Silicon for Active Photonic Devices. IEEE Trans. Electron Devices 2013, 60 (5), 1495–1505. 10.1109/TED.2013.2249069. [DOI] [Google Scholar]
- Luu V. N. Swelling of alpha-quartz induced by MeV ions irradiation: Critical dose and swelling mechanism. J. Nucl. Mater. 2020, 539, 152266. 10.1016/j.jnucmat.2020.152266. [DOI] [Google Scholar]
- Pignatelli I.; Kumar A.; Field K. G.; Wang B.; Yu Y.; Le Pape Y.; Bauchy M.; Sant G. Direct Experimental Evidence for Differing Reactivity Alterations of Minerals following Irradiation: The Case of Calcite and Quartz. Sci. Rep. 2016, 6, 20155. 10.1038/srep20155. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Krishnan N. M. A.; Le Pape Y.; Sant G.; Bauchy M. Effect of irradiation on silicate aggregates’ density and stiffness. J. Nucl. Mater. 2018, 512, 126–136. 10.1016/j.jnucmat.2018.10.009. [DOI] [Google Scholar]
- Silva C. M.; Rosseel T. M.; Kirkegaard M. C. Radiation-Induced Changes in Quartz, A Mineral Analog of Nuclear Power Plant Concrete Aggregates. Inorg. Chem. 2018, 57 (6), 3329–3338. 10.1021/acs.inorgchem.8b00096. [DOI] [PubMed] [Google Scholar]
- Wittels M. C. Structural behaviou of neutron irradiated quartz. Philos. Mag. 1957, 2 (24), 1445–1461. 10.1080/14786435708241190. [DOI] [Google Scholar]
- Thomas P. A.; Brodsky M. H.; Kaplan D.; Lepine D. Electron spin resonance of ultrahigh vacuum evaporated amorphous silicon:In situandex situstudies. Phys. Rev. B 1978, 18 (7), 3059–3073. 10.1103/PhysRevB.18.3059. [DOI] [Google Scholar]
- Fredrickson J. E.; Waddell C. N.; Spitzer W. G.; Hubler G. K. Effects of thermal annealing on the refractive index of amorphous silicon produced by ion implantation. Appl. Phys. Lett. 1982, 40 (2), 172–174. 10.1063/1.93032. [DOI] [Google Scholar]
- Sinke W.; Warabisako T.; Miyao M.; Tokuyama T.; Roorda S.; Saris F. W. Transient structural relaxation of amorphous silicon. J. Non-Cryst. Solids 1988, 99 (2–3), 308–323. 10.1016/0022-3093(88)90439-5. [DOI] [Google Scholar]
- Laaziri K.; Kycia S.; Roorda S.; Chicoine M.; Robertson J. L.; Wang J.; Moss S. C. High Resolution Radial Distribution Function of Pure Amorphous Silicon. Phys. Rev. Lett. 1999, 82 (17), 3460–3463. 10.1103/PhysRevLett.82.3460. [DOI] [Google Scholar]
- Wit L. D.; Roorda S.; Sinke W. C.; Saris F. W.; Berntsen A. J. M.; Van Der Weg W. F. Structural Relaxation of Amorphous Silicon Induced by High Temperature Annealing. MRS Proceedings 1990, 205, 3–8. 10.1557/PROC-205-3. [DOI] [Google Scholar]
- Mercure J. F.; Karmouch R.; Anahory Y.; Roorda S.; Schiettekatte F. Dependence of the structural relaxation of amorphous silicon on implantation temperature. Phys. Rev. B 2005, 71 (13), 134205. 10.1103/PhysRevB.71.134205. [DOI] [Google Scholar]
- Weeks R. A.; Nelson C. M. Irradiation Effects and Short-Range Order in Fused Silica and Quartz. J. Appl. Phys. 1960, 31 (9), 1555–1558. 10.1063/1.1735892. [DOI] [Google Scholar]
- Mota F.; Caturla M.J.; Perlado J.M.; Ibarra A.; León M.; Mollá J.. Radiation Damage Modeling of Fused Silica in Fusion Systems. 2006. Conference. Accessed 2024-November-24. https://www-pub.iaea.org/MTCD/Meetings/FEC2006/it_p1-30.pdf.
- Douillard L.; Duraud J. P. Swift heavy ion amorphization of quartz — a comparative study of the particle amorphization mechanism of quartz. Nucl. Instrum. Methods Phys. Res., Sect. B 1996, 107 (1–4), 212–217. 10.1016/0168-583X(95)01044-0. [DOI] [Google Scholar]
- Lebel O. S.; A Molecular Glasses: Emerging Materialsfor the Next Generation. Adv. Mater. 2020, 239–258. 10.1515/9783110537734-009. [DOI] [Google Scholar]
- Craig D. The relevance of the amorphous state to pharmaceutical dosage forms: glassy drugs and freeze dried systems. Int. J. Pharm. 1999, 179 (2), 179–207. 10.1016/S0378-5173(98)00338-X. [DOI] [PubMed] [Google Scholar]
- Willart J. F.; Descamps M. Solid state amorphization of pharmaceuticals. Mol. Pharmaceutics 2008, 5 (6), 905–920. 10.1021/mp800092t. [DOI] [PubMed] [Google Scholar]
- Babu N. J.; Nangia A. Solubility Advantage of Amorphous Drugs and Pharmaceutical Cocrystals. Cryst. Growth Des. 2011, 11 (7), 2662–2679. 10.1021/cg200492w. [DOI] [Google Scholar]
- Di Martino P.; Magnoni F.; Vargas Peregrina D.; Rosa Gigliobianco M.; Censi R.; Malaj L. Formation, Physicochemical Characterization, and Thermodynamic Stability of the Amorphous State of Drugs and Excipients. Curr. Pharm. Des. 2016, 22 (32), 4959–4974. 10.2174/1381612822666160726105658. [DOI] [PubMed] [Google Scholar]
- Kissi E. O.; Grohganz H.; Lobmann K.; Ruggiero M. T.; Zeitler J. A.; Rades T. Glass-Transition Temperature of the beta-Relaxation as the Major Predictive Parameter for Recrystallization of Neat Amorphous Drugs. J. Phys. Chem. B 2018, 122 (10), 2803–2808. 10.1021/acs.jpcb.7b10105. [DOI] [PubMed] [Google Scholar]
- Kawakami K. Ultraslow Cooling for the Stabilization of Pharmaceutical Glasses. J. Phys. Chem. B 2019, 123 (23), 4996–5003. 10.1021/acs.jpcb.9b02122. [DOI] [PubMed] [Google Scholar]
- Wang Y.; Wang Y.; Cheng J.; Chen H.; Xu J.; Liu Z.; Shi Q.; Zhang C. Recent Advances in the Application of Characterization Techniques for Studying Physical Stability of Amorphous Pharmaceutical Solids. Crystals 2021, 11 (12), 1440. 10.3390/cryst11121440. [DOI] [Google Scholar]
- Zhou X.; Blochwitz J.; Pfeiffer M.; Nollau A.; Fritz T.; Leo K. Enhanced Hole Injection into Amorphous Hole-Transport Layers of Organic Light-Emitting Diodes Using Controlled p-Type Doping. Adv. Funct. Mater. 2001, 11 (4), 310–314. . [DOI] [Google Scholar]
- Nomura M.; Fukukawa K.; Shibasaki Y.; Ueda M. New amorphous hole-transporting molecular materials: 1,1,1-Tris(4-(4-diarylaminobenzoyloxy)phenyl)ethane. Synth. Met. 2002, 132 (1), 9–13. 10.1016/S0379-6779(02)00216-3. [DOI] [Google Scholar]
- Nomura M.; Shibasaki Y.; Ueda M.; Tugita K.; Ichikawa M.; Taniguchi Y. New amorphous electron-transporting materials based on Tris-benzimidazoles for all wet-process OLED devices. Synth. Met. 2005, 151 (3), 261–268. 10.1016/j.synthmet.2005.05.010. [DOI] [Google Scholar]
- Moorthy J. N.; Venkatakrishnan P.; Natarajan P.; Lin Z.; Chow T. J. Nondoped pure-blue OLEDs based on amorphous phenylenevinylene-functionalized twisted bimesitylenes. J. Org. Chem. 2010, 75 (8), 2599–2609. 10.1021/jo1001565. [DOI] [PubMed] [Google Scholar]
- Kim J. Y.; Yasuda T.; Yang Y. S.; Adachi C. Bifunctional star-burst amorphous molecular materials for OLEDs: achieving highly efficient solid-state luminescence and carrier transport induced by spontaneous molecular orientation. Adv. Mater. 2013, 25 (19), 2666–2671. 10.1002/adma.201204902. [DOI] [PubMed] [Google Scholar]
- Bagchi K.; Ediger M. D. Controlling Structure and Properties of Vapor-Deposited Glasses of Organic Semiconductors: Recent Advances and Challenges. J. Phys. Chem. Lett. 2020, 11 (17), 6935–6945. 10.1021/acs.jpclett.0c01682. [DOI] [PubMed] [Google Scholar]
- Sun L.; Yoshida T.; Harada Y.; White M.S.; Suzuri Y. Amorphous dielectric metal-organic electron injection layer for efficient inverted organic light-emitting diodes. Org. Electron. 2023, 122, 106878. 10.1016/j.orgel.2023.106878. [DOI] [Google Scholar]
- Yokoyama D.; Qiang Wang Z.; Pu Y.-J.; Kobayashi K.; Kido J.; Hong Z. High-efficiency simple planar heterojunction organic thin-film photovoltaics with horizontally oriented amorphous donors. Sol. Energy Mater. Sol. Cells 2012, 98, 472–475. 10.1016/j.solmat.2011.10.014. [DOI] [Google Scholar]
- Hedley G. J.; Ruseckas A.; Samuel I. D. Light Harvesting for Organic Photovoltaics. Chem. Rev. 2017, 117 (2), 796–837. 10.1021/acs.chemrev.6b00215. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nunzi J. M.; Lebel O. Revisiting the Optimal Nano-Morphology: Towards Amorphous Organic Photovoltaics. Chem. Rec. 2019, 19 (6), 1028–1038. 10.1002/tcr.201800158. [DOI] [PubMed] [Google Scholar]
- Esembeson B.; Scimeca M. L.; Michinobu T.; Diederich F.; Biaggio I. A High-Optical Quality Supramolecular Assembly for Third-Order Integrated Nonlinear Optics. Adv. Mater. 2008, 20 (23), 4584–4587. 10.1002/adma.200801552. [DOI] [Google Scholar]
- Jang S. H.; Jen A. K. Electro-optic (E-O) molecular glasses. Chem.—Asian J. 2009, 4 (1), 20–31. 10.1002/asia.200800179. [DOI] [PubMed] [Google Scholar]
- Traskovskis K.; Mihailovs I.; Tokmakovs A.; Jurgis A.; Kokars V.; Rutkis M. Triphenyl moieties as building blocks for obtaining molecular glasses with nonlinear optical activity. J. Mater. Chem. 2012, 22 (22), 11268. 10.1039/c2jm30861d. [DOI] [Google Scholar]
- Seniutinas G.; Tomašiu̅nas R.; Czaplicki R.; Sahraoui B.; Daškevičienė M.; Getautis V.; Balevičius Z. Arylmethylene-1,3-indandione based molecular glasses: Third order optical non-linearity. Dyes Pigm. 2012, 95 (1), 33–40. 10.1016/j.dyepig.2012.03.011. [DOI] [Google Scholar]
- Gu C.; Huang Z.; Rahman A.; Chen X.; Zeng Z.; Liang Z.; Shi L.; Liu F.; Wang J. Supramolecular self-assembled nonlinear optical molecular glasses with enhanced electro-optic activity and alignment stability. Dyes Pigm. 2022, 202, 110283. 10.1016/j.dyepig.2022.110283. [DOI] [Google Scholar]
- Biaggio I. The Appeal of Small Molecules for Practical Nonlinear Optics. Chemistry 2022, 28 (6), e202103168 10.1002/chem.202103168. [DOI] [PubMed] [Google Scholar]
- Angell C. A. Formation of Glasses from Liquids and Biopolymers. Science 1995, 267 (5206), 1924–1935. 10.1126/science.267.5206.1924. [DOI] [PubMed] [Google Scholar]
- Angell C. A.; Ngai K. L.; McKenna G. B.; McMillan P. F.; Martin S. W. Relaxation in glassforming liquids and amorphous solids. J. Appl. Phys. 2000, 88 (6), 3113–3157. 10.1063/1.1286035. [DOI] [Google Scholar]
- Cavagna A. Supercooled liquids for pedestrians. Phys. Rep. 2009, 476 (4–6), 51–124. 10.1016/j.physrep.2009.03.003. [DOI] [Google Scholar]
- Lee T. W.; Noh T.; Shin H. W.; Kwon O.; Park J. J.; Choi B. K.; Kim M. S.; Shin D. W.; Kim Y. R. Characteristics of Solution-Processed Small-Molecule Organic Films and Light-Emitting Diodes Compared with their Vacuum-Deposited Counterparts. Adv. Funct. Mater. 2009, 19 (10), 1625–1630. 10.1002/adfm.200801045. [DOI] [Google Scholar]
- Niu Q.; Zhang Q.; Xu W.; Jiang Y.; Xia R.; Bradley D. D. C.; Li D.; Wen X. Solution-processed anthracene-based molecular glasses as stable blue-light-emission laser gain media. Org. Electron. 2015, 18, 95–100. 10.1016/j.orgel.2015.01.012. [DOI] [Google Scholar]
- Gujral A.; Yu L.; Ediger M. D. Anisotropic organic glasses. Curr. Opin. Solid State Mater. Sci. 2018, 22 (2), 49–57. 10.1016/j.cossms.2017.11.001. [DOI] [Google Scholar]
- Mao G.; Wu Z.; He Q.; Jiao B.; Xu G.; Hou X.; Chen Z.; Gong Q. Considerable improvement in the stability of solution processed small molecule OLED by annealing. Appl. Surf. Sci. 2011, 257 (17), 7394–7398. 10.1016/j.apsusc.2011.02.046. [DOI] [Google Scholar]
- Feng S.; Duan L.; Hou L.; Qiao J.; Zhang D.; Dong G.; Wang L.; Qiu Y. A Comparison Study of the Organic Small Molecular Thin Films Prepared by Solution Process and Vacuum Deposition: Roughness, Hydrophilicity, Absorption, Photoluminescence, Density, Mobility, and Electroluminescence. J. Phys. Chem. C 2011, 115 (29), 14278–14284. 10.1021/jp203674p. [DOI] [Google Scholar]
- Wang D.; Wu Z.; Lei X.; Zhang W.; Jiao B.; Wang D.; Hou X. Electroluminescence of solution-processed organic light-emitting diodes based on fluorescent small molecules and polymer as hole-transporting layer. Phys. Status Solidi 2013, 210 (12), 2556–2560. 10.1002/pssa.201329205. [DOI] [Google Scholar]
- Xing X.; Zhong L.; Zhang L.; Chen Z.; Qu B.; Chen E.; Xiao L.; Gong Q. Essential Differences of Organic Films at the Molecular Level via Vacuum Deposition and Solution Processes for Organic Light-Emitting Diodes. J. Phys. Chem. C 2013, 117 (48), 25405–25408. 10.1021/jp410547w. [DOI] [Google Scholar]
- Prakash S.; Sims M.; Wyrsta I.; Parker I. D.; Kondakov D.; Gao W. 49.1: Invited Paper: Solution-Processed OLEDs: Unique Challenges and Advantages. SID Symp. Dig. Technol. Pap. 2013, 44 (1), 678–681. 10.1002/j.2168-0159.2013.tb06303.x. [DOI] [Google Scholar]
- Pfeiffer F.; Felix N. M.; Neuber C.; Ober C. K.; Schmidt H. W. Physical Vapor Deposition of Molecular Glass Photoresists: A New Route to Chemically Amplified Patterning. Adv. Funct. Mater. 2007, 17 (14), 2336–2342. 10.1002/adfm.200600717. [DOI] [Google Scholar]
- Singh S.; de Pablo J. J. A molecular view of vapor deposited glasses. J. Chem. Phys. 2011, 134 (19), 194903. 10.1063/1.3586805. [DOI] [PubMed] [Google Scholar]
- Beasley M. S.; Bishop C.; Kasting B. J.; Ediger M. D. Vapor-Deposited Ethylbenzene Glasses Approach ″Ideal Glass″ Density. J. Phys. Chem. Lett. 2019, 10 (14), 4069–4075. 10.1021/acs.jpclett.9b01508. [DOI] [PubMed] [Google Scholar]
- Jin Y.; Zhang A.; Wolf S.E.; Govind S.; Moore A.R.; Zhernenkov M.; Freychet G.; Arabi Shamsabadi A.; Fakhraai Z. Glasses denser than the supercooled liquid. Proc. Natl. Acad. Sci. U. S. A. 2021, 118 (31), e2100738118. 10.1073/pnas.2100738118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Luo P.; et al. High-density stable glasses formed on soft substrates. Nat. Mater. 2024, 23 (5), 688–694. 10.1038/s41563-024-01828-w. [DOI] [PubMed] [Google Scholar]
- Kearns K. L.; Swallen S. F.; Ediger M. D.; Wu T.; Sun Y.; Yu L. Hiking down the energy landscape: progress toward the Kauzmann temperature via vapor deposition. J. Phys. Chem. B 2008, 112 (16), 4934–4942. 10.1021/jp7113384. [DOI] [PubMed] [Google Scholar]
- Ishii K.; Yokoyama Y.; Moriyama R.; Nakayama H. Liquid–Liquid Relaxation in the Supercooled Liquid State of Ethylbenzene: Thermal Studies Using a Prototype DTA Sensor for the Study of Vapor-deposited Samples. Chem. Lett. 2010, 39 (9), 958–960. 10.1246/cl.2010.958. [DOI] [Google Scholar]
- Kearns K. L.; Still T.; Fytas G.; Ediger M. D. High-modulus organic glasses prepared by physical vapor deposition. Adv. Mater. 2010, 22 (1), 39–42. 10.1002/adma.200901673. [DOI] [PubMed] [Google Scholar]
- Fakhraai Z.; Still T.; Fytas G.; Ediger M. D. Structural Variations of an Organic Glassformer Vapor-Deposited onto a Temperature Gradient Stage. J. Phys. Chem. Lett. 2011, 2 (5), 423–427. 10.1021/jz101723d. [DOI] [Google Scholar]
- Wolf S. E.; Fulco S.; Zhang A.; Zhao H.; Walsh P. J.; Turner K. T.; Fakhraai Z. Role of Molecular Layering in the Enhanced Mechanical Properties of Stable Glasses. J. Phys. Chem. Lett. 2022, 13 (15), 3360–3368. 10.1021/acs.jpclett.2c00232. [DOI] [PubMed] [Google Scholar]
- Hellman F. Surface-induced ordering: A model for vapor-deposition growth of amorphous materials. Appl. Phys. Lett. 1994, 64 (15), 1947–1949. 10.1063/1.111751. [DOI] [Google Scholar]
- Yokoyama D.; Setoguchi Y.; Sakaguchi A.; Suzuki M.; Adachi C. Orientation Control of Linear-Shaped Molecules in Vacuum-Deposited Organic Amorphous Films and Its Effect on Carrier Mobilities. Adv. Funct. Mater. 2010, 20 (3), 386–391. 10.1002/adfm.200901684. [DOI] [Google Scholar]
- Dawson K. J.; Zhu L.; Yu L.; Ediger M. D. Anisotropic structure and transformation kinetics of vapor-deposited indomethacin glasses. J. Phys. Chem. B 2011, 115 (3), 455–463. 10.1021/jp1092916. [DOI] [PubMed] [Google Scholar]
- Yokoyama D. Molecular orientation in small-molecule organic light-emitting diodes. J. Mater. Chem. 2011, 21 (48), 19187. 10.1039/c1jm13417e. [DOI] [Google Scholar]
- Dawson K.; Kopff L. A.; Zhu L.; McMahon R. J.; Yu L.; Richert R.; Ediger M. D. Molecular packing in highly stable glasses of vapor-deposited tris-naphthylbenzene isomers. J. Chem. Phys. 2012, 136 (9), 094505. 10.1063/1.3686801. [DOI] [PubMed] [Google Scholar]
- Dalal S. S.; Ediger M. D. Molecular Orientation in Stable Glasses of Indomethacin. J. Phys. Chem. Lett. 2012, 3 (10), 1229–1233. 10.1021/jz3003266. [DOI] [PubMed] [Google Scholar]
- Gujral A.; O’Hara K. A.; Toney M. F.; Chabinyc M. L.; Ediger M. D. Structural Characterization of Vapor-Deposited Glasses of an Organic Hole Transport Material with X-ray Scattering. Chem. Mater. 2015, 27 (9), 3341–3348. 10.1021/acs.chemmater.5b00583. [DOI] [Google Scholar]
- Gomez J.; Jiang J.; Gujral A.; Huang C.; Yu L.; Ediger M. D. Vapor deposition of a smectic liquid crystal: highly anisotropic, homogeneous glasses with tunable molecular orientation. Soft. Matter. 2016, 12 (11), 2942–2947. 10.1039/C5SM02944A. [DOI] [PubMed] [Google Scholar]
- Liu T.; et al. Birefringent Stable Glass with Predominantly Isotropic Molecular Orientation. Phys. Rev. Lett. 2017, 119 (9), 095502. 10.1103/PhysRevLett.119.095502. [DOI] [PubMed] [Google Scholar]
- Laventure A.; Gujral A.; Lebel O.; Pellerin C.; Ediger M. D. Influence of Hydrogen Bonding on the Kinetic Stability of Vapor-Deposited Glasses of Triazine Derivatives. J. Phys. Chem. B 2017, 121 (10), 2350–2358. 10.1021/acs.jpcb.6b12676. [DOI] [PubMed] [Google Scholar]
- Walters D. M.; Antony L.; de Pablo J. J.; Ediger M. D. Influence of Molecular Shape on the Thermal Stability and Molecular Orientation of Vapor-Deposited Organic Semiconductors. J. Phys. Chem. Lett. 2017, 8 (14), 3380–3386. 10.1021/acs.jpclett.7b01097. [DOI] [PubMed] [Google Scholar]
- Sohn S.; Park K. H.; Kwon S. K.; Lee H. K.; Ahn H.; Jung S.; Kim Y. H. Preferential Orientation of Tetrahedral Silicon-Based Hosts in Phosphorescent Organic Light-Emitting Diodes. ACS Omega 2018, 3 (8), 9989–9996. 10.1021/acsomega.8b01358. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bagchi K.; Jackson N. E.; Gujral A.; Huang C.; Toney M. F.; Yu L.; de Pablo J. J.; Ediger M. D. Origin of Anisotropic Molecular Packing in Vapor-Deposited Alq3 Glasses. J. Phys. Chem. Lett. 2019, 10 (2), 164–170. 10.1021/acs.jpclett.8b03582. [DOI] [PubMed] [Google Scholar]
- Ediger M. D.; de Pablo J.; Yu L. Anisotropic Vapor-Deposited Glasses: Hybrid Organic Solids. Acc. Chem. Res. 2019, 52 (2), 407–414. 10.1021/acs.accounts.8b00513. [DOI] [PubMed] [Google Scholar]
- Oh-e M.; Ogata H.; Araoka F. Randomization and Constraint of Molecular Alignment and Orientation: Temperature-Dependent Anisotropy and Phase Transition in Vapor-Deposited Thin Films of an Organic Cross-Shaped Molecule. ACS Omega 2019, 4 (1), 39–47. 10.1021/acsomega.8b02560. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mikaeili A.; Matsushima T.; Esaki Y.; Yazdani S. A.; Adachi C.; Mohajerani E. The origin of changes in electrical properties of organic films fabricated at various vacuum-deposition rates. Opt. Mater. 2019, 91, 93–100. 10.1016/j.optmat.2019.03.012. [DOI] [Google Scholar]
- Cang Y.; Wang Z.; Bishop C.; Yu L.; Ediger M.D.; Fytas G. Extreme Elasticity Anisotropy in Molecular Glasses. Adv. Funct. Mater. 2020, 30 (23), 2001481. 10.1002/adfm.202001481. [DOI] [Google Scholar]
- Thelen J. L.; et al. Molecular Orientation Depth Profiles in Organic Glasses Using Polarized Resonant Soft X-ray Reflectivity. Chem. Mater. 2020, 32 (15), 6295–6309. 10.1021/acs.chemmater.0c02044. [DOI] [Google Scholar]
- Tanaka M.; Noda H.; Nakanotani H.; Adachi C. Molecular orientation of disk-shaped small molecules exhibiting thermally activated delayed fluorescence in host–guest films. Appl. Phys. Lett. 2020, 116 (2), 023302. 10.1063/1.5140210. [DOI] [Google Scholar]
- Bishop C.; Li Y.; Toney M. F.; Yu L.; Ediger M. D. Molecular Orientation for Vapor-Deposited Organic Glasses Follows Rate-Temperature Superposition: The Case of Posaconazole. J. Phys. Chem. B 2020, 124 (12), 2505–2513. 10.1021/acs.jpcb.0c00625. [DOI] [PubMed] [Google Scholar]
- Ferron T. J.; Thelen J. L.; Bagchi K.; Deng C.; Gann E.; de Pablo J. J.; Ediger M. D.; Sunday D. F.; DeLongchamp D. M. Characterization of the Interfacial Orientation and Molecular Conformation in a Glass-Forming Organic Semiconductor. ACS Appl. Mater. Interfaces 2022, 14 (2), 3455–3466. 10.1021/acsami.1c19948. [DOI] [PubMed] [Google Scholar]
- Ferron T. J.; Fiori M. E.; Ediger M. D.; DeLongchamp D. M.; Sunday D. F. Composition Dictates Molecular Orientation at the Heterointerfaces of Vapor-Deposited Glasses. JACS Au 2023, 3 (7), 1931–1938. 10.1021/jacsau.3c00168. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bishop C.; Thelen J. L.; Gann E.; Toney M. F.; Yu L.; DeLongchamp D. M.; Ediger M. D. Vapor deposition of a nonmesogen prepares highly structured organic glasses. Proc. Natl. Acad. Sci. U. S. A. 2019, 116 (43), 21421–21426. 10.1073/pnas.1908445116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stevenson J. D.; Wolynes P. G. On the surface of glasses. J. Chem. Phys. 2008, 129 (23), 234514. 10.1063/1.3041651. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yu L. Surface mobility of molecular glasses and its importance in physical stability. Adv. Drug. Delivery Rev. 2016, 100, 3–9. 10.1016/j.addr.2016.01.005. [DOI] [PubMed] [Google Scholar]
- Samanta S.; et al. Exploring the Importance of Surface Diffusion in Stability of Vapor-Deposited Organic Glasses. J. Phys. Chem. B 2019, 123 (18), 4108–4117. 10.1021/acs.jpcb.9b01012. [DOI] [PubMed] [Google Scholar]
- Chen Y.; Chen Z.; Tylinski M.; Ediger M. D.; Yu L. Effect of molecular size and hydrogen bonding on three surface-facilitated processes in molecular glasses: Surface diffusion, surface crystal growth, and formation of stable glasses by vapor deposition. J. Chem. Phys. 2019, 150 (2), 024502. 10.1063/1.5079441. [DOI] [PubMed] [Google Scholar]
- Fiori M. E.; Bagchi K.; Toney M.F.; Ediger M.D. Surface equilibration mechanism controls the molecular packing of glassy molecular semiconductors at organic interfaces. Proc. Natl. Acad. Sci. U. S. A. 2021, 118 (42), e2111988118. 10.1073/pnas.2111988118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Luo P.; Fakhraai Z. Surface-Mediated Formation of Stable Glasses. Annu. Rev. Phys. Chem. 2023, 74, 361–389. 10.1146/annurev-physchem-042018-052708. [DOI] [PubMed] [Google Scholar]
- Mayr C.; Brütting W. Control of Molecular Dye Orientation in Organic Luminescent Films by the Glass Transition Temperature of the Host Material. Chem. Mater. 2015, 27 (8), 2759–2762. 10.1021/acs.chemmater.5b00062. [DOI] [Google Scholar]
- Liu H.; Zhao Z.; Zhou Q.; Chen R.; Yang K.; Wang Z.; Tang L.; Bauchy M. Challenges and opportunities in atomistic simulations of glasses: a review. C.R. GEOSCI. 2022, 354 (S1), 35–77. 10.5802/crgeos.116. [DOI] [Google Scholar]
- Du J., Challenges in Molecular Dynamics Simulations of Multicomponent Oxide Glasses In Molecular Dynamics Simulations of Disordered Materials: From Network Glasses to Phase-Change Memory Alloys; Massobrio C. et al, Eds.; 2015, Springer International Publishing: Cham. p 157–180. [Google Scholar]
- Horbach J.; Kob W.; Binder K.; Angell C. A. Finite size effects in simulations of glass dynamics. Phys. Rev. E Stat. Phys. Plasmas Fluids Relat. Interdiscip. Topics 1996, 54 (6), R5897–R5900. 10.1103/PhysRevE.54.R5897. [DOI] [PubMed] [Google Scholar]
- Berthier L.; Biroli G.; Coslovich D.; Kob W.; Toninelli C. Finite-size effects in the dynamics of glass-forming liquids. Phys. Rev. E. Stat. Nonlin. Soft Matter. Phys. 2012, 86 (3 Pt 1), 031502. 10.1103/PhysRevE.86.031502. [DOI] [PubMed] [Google Scholar]
- Du J.Molecular dynamics simulations of oxide glasses. In Springer Handbook of Glass; Musgraves J. D., Hu J.; Calvez L., Eds.; Springer International Publishing, Cham, 2019; pp 1131–1155. [Google Scholar]
- Leach A. R.Molecular Modelling: Principles and Applications; Prentice Hall, New York. 2001. [Google Scholar]
- Ganster P.; Benoit M.; Kob W.; Delaye J. M. Structural properties of a calcium aluminosilicate glass from molecular-dynamics simulations: a finite size effects study. J. Chem. Phys. 2004, 120 (21), 10172–10181. 10.1063/1.1724815. [DOI] [PubMed] [Google Scholar]
- Nakano A.; Kalia R. K.; Vashishta P. First sharp diffraction peak and intermediate-range order in amorphous silica: finite-size effects in molecular dynamics simulations. J. Non-Cryst. Solids 1994, 171 (2), 157–163. 10.1016/0022-3093(94)90351-4. [DOI] [Google Scholar]
- van Duin A. C. T.; Strachan A.; Stewman S.; Zhang Q.; Xu X.; Goddard W. A. ReaxFFSiO Reactive Force Field for Silicon and Silicon Oxide Systems. J. Phys. Chem. A 2003, 107 (19), 3803–3811. 10.1021/jp0276303. [DOI] [Google Scholar]
- Liu H.; Fu Z.; Li Y.; Sabri N. F. A.; Bauchy M. Balance between accuracy and simplicity in empirical forcefields for glass modeling: Insights from machine learning. J. Non-Cryst. Solids 2019, 515, 133–142. 10.1016/j.jnoncrysol.2019.04.020. [DOI] [Google Scholar]
- Massobrio C., Du J.; Bernasconi M.; Salmon P.S.. Molecular Dynamics Simulations of Disordered Materials: From Network Glasses to Phase-Change Memory Alloys; Springer International Publishing, 2015. [Google Scholar]
- Shewchuk J. R.An Introduction to the Conjugate Gradient Method Without the Agonizing Pain, 1994. Carnegie Mellon University, United States. https://www.cs.cmu.edu/%7Equake-papers/painless-conjugate-gradient.pdf. [Google Scholar]
- Liu H.; Fu Z.; Li Y.; Sabri N. F. A.; Bauchy M. Parameterization of empirical forcefields for glassy silica using machine learning. MRS Commun. 2019, 9 (2), 593–599. 10.1557/mrc.2019.47. [DOI] [Google Scholar]
- Thorpe M. F. Continuous deformations in random networks. J. Non Cryst. Solids 1983, 57 (3), 355–370. 10.1016/0022-3093(83)90424-6. [DOI] [Google Scholar]
- Horbach J.; Kob W. Static and dynamic properties of a viscous silica melt. Phys. Rev. B 1999, 60 (5), 3169–3181. 10.1103/PhysRevB.60.3169. [DOI] [Google Scholar]
- Zaccone A.Theory of Disordered Solids; Lecture Notes in Physics; Springer: Cham, 2023. [Google Scholar]
- Lewis L. J. Fifty years of amorphous silicon models: the end of the story?. J. Non Cryst. Solids 2022, 580, 121383. 10.1016/j.jnoncrysol.2021.121383. [DOI] [Google Scholar]
- Micoulaut M.; Bauchy M. Topology and Rigidity of Silicate Melts and Glasses. Rev. Mineral. Geochem. 2022, 87 (1), 163–191. 10.2138/rmg.2022.87.04. [DOI] [Google Scholar]
- Polk D. E. Structural model for amorphous silicon and germanium. J. Non Cryst. Solids 1971, 5 (5), 365–376. 10.1016/0022-3093(71)90038-X. [DOI] [Google Scholar]
- Wooten F.; Winer K.; Weaire D. Computer generation of structural models of amorphous Si and Ge. Phys. Rev. Lett. 1985, 54 (13), 1392–1395. 10.1103/PhysRevLett.54.1392. [DOI] [PubMed] [Google Scholar]
- Keating P. N. Effect of Invariance Requirements on the Elastic Strain Energy of Crystals with Application to the Diamond Structure. Phys. Rev. 1966, 145 (2), 637–645. 10.1103/PhysRev.145.637. [DOI] [Google Scholar]
- Djordjevic B. R.; Thorpe M. F.; Wooten F. Computer model of tetrahedral amorphous diamond. Phys. Rev. B Condens. Matter. 1995, 52 (8), 5685–5689. 10.1103/PhysRevB.52.5685. [DOI] [PubMed] [Google Scholar]
- Barkema G. T.; Mousseau N. High-quality continuous random networks. Phys. Rev. B 2000, 62 (8), 4985–4990. 10.1103/PhysRevB.62.4985. [DOI] [Google Scholar]
- Vink R. L. C.; Barkema G. T.; Stijnman M. A.; Bisseling R. H. Device-size atomistic models of amorphous silicon. Phys. Rev. B 2001, 64 (24), 245214. 10.1103/PhysRevB.64.245214. [DOI] [Google Scholar]
- Stillinger F. H.; Weber T. A. Computer simulation of local order in condensed phases of silicon. Phys. Rev. B Condens. Matter. 1985, 31 (8), 5262–5271. 10.1103/PhysRevB.31.5262. [DOI] [PubMed] [Google Scholar]
- Vink R. L. C.; Barkema G. T.; van der Weg W. F.; Mousseau N. Fitting the Stillinger–Weber potential to amorphous silicon. J. Non Cryst. Solids 2001, 282 (2–3), 248–255. 10.1016/S0022-3093(01)00342-8. [DOI] [Google Scholar]
- Hejna M.; Steinhardt P. J.; Torquato S. Nearly hyperuniform network models of amorphous silicon. Phys. Rev. B 2013, 87 (24), 245204. 10.1103/PhysRevB.87.245204. [DOI] [Google Scholar]
- Mousseau N.; Barkema G. T. Binary continuous random networks. J. Phys.: Condens. Matter 2004, 16 (44), S5183–S5190. 10.1088/0953-8984/16/44/015. [DOI] [Google Scholar]
- Laaziri K.; Kycia S.; Roorda S.; Chicoine M.; Robertson J. L.; Wang J.; Moss S. C. High-energy x-ray diffraction study of pure amorphous silicon. Phys. Rev. B 1999, 60 (19), 13520–13533. 10.1103/PhysRevB.60.13520. [DOI] [Google Scholar]
- Ovshinsky S. R. Localized States in the Gap of Amorphous Semiconductors. Phys. Rev. Lett. 1976, 36 (24), 1469–1472. 10.1103/PhysRevLett.36.1469. [DOI] [Google Scholar]
- Pandey K. C. Newπ-Bonded Chain Model for Si(111)-(2 × 1) Surface. Phys. Rev. Lett. 1981, 47 (26), 1913–1917. 10.1103/PhysRevLett.47.1913. [DOI] [Google Scholar]
- Srolovitz D.; Maeda K.; Vitek V.; Egami T. Structural defects in amorphous solids Statistical analysis of a computer model. Philos. Mag. A 1981, 44 (4), 847–866. 10.1080/01418618108239553. [DOI] [Google Scholar]
- Adler D. Normal structural bonding and defects in covalent amorphous solids. J. Solid State Chem. 1982, 45 (1), 40–50. 10.1016/0022-4596(82)90289-4. [DOI] [Google Scholar]
- Rucavado E.; et al. Enhancing the optoelectronic properties of amorphous zinc tin oxide by subgap defect passivation: A theoretical and experimental demonstration. Phys. Rev. B 2017, 95 (24), 245204. 10.1103/PhysRevB.95.245204. [DOI] [Google Scholar]
- Newnham R. E.; Sundar V.; Yimnirun R.; Su J.; Zhang Q. M. Electrostriction: Nonlinear Electromechanical Coupling in Solid Dielectrics. J. Phys. Chem. B 1997, 101 (48), 10141–10150. 10.1021/jp971522c. [DOI] [Google Scholar]
- Egami T.; Maeda K.; Vitek V. Structural defects in amorphous solids A computer simulation study. Philos. Mag. A 1980, 41 (6), 883–901. 10.1080/01418618008243894. [DOI] [Google Scholar]
- Paret J.; Jack R. L.; Coslovich D. Assessing the structural heterogeneity of supercooled liquids through community inference. J. Chem. Phys. 2020, 152 (14), 144502. 10.1063/5.0004732. [DOI] [PubMed] [Google Scholar]
- Tanguy A.; Mantisi B.; Tsamados M. Vibrational modes as a predictor for plasticity in a model glass. EPL (Europhysics Letters) 2010, 90 (1), 16004. 10.1209/0295-5075/90/16004. [DOI] [Google Scholar]
- Manning M. L.; Liu A. J. Vibrational modes identify soft spots in a sheared disordered packing. Phys. Rev. Lett. 2011, 107 (10), 108302. 10.1103/PhysRevLett.107.108302. [DOI] [PubMed] [Google Scholar]
- Pusztai L.; Kugler S. Reverse Monte Carlo simulation: the structure of amorphous silicon. J. Non-Cryst. Solids 1993, 164–166, 147–150. 10.1016/0022-3093(93)90513-W. [DOI] [Google Scholar]
- McGreevy R. L.; Pusztai L. Reverse Monte Carlo Simulation: A New Technique for the Determination of Disordered Structures. Mol. Simul. 1988, 1 (6), 359–367. 10.1080/08927028808080958. [DOI] [Google Scholar]
- McGreevy R. L.; Howe M. A. RMC: Modeling Disordered Structures. Annu. Rev. Mater. Sci. 1992, 22 (1), 217–242. 10.1146/annurev.ms.22.080192.001245. [DOI] [Google Scholar]
- Biswas P.; Atta-Fynn R.; Drabold D. A. Reverse Monte Carlo modeling of amorphous silicon. Phys. Rev. B 2004, 69 (19), 195207. 10.1103/PhysRevB.69.195207. [DOI] [Google Scholar]
- Pandey A.; Biswas P.; Drabold D.A. Force-enhanced atomic refinement: Structural modeling with interatomic forces in a reverse Monte Carlo approach applied to amorphous Si and SiO2. Phys. Rev. B 2015, 92 (15), 155205. 10.1103/PhysRevB.92.155205. [DOI] [Google Scholar]
- Pandey A.; Biswas P.; Drabold D. A. Inversion of diffraction data for amorphous materials. Sci. Rep. 2016, 6 (1), 33731. 10.1038/srep33731. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pandey A.; Biswas P.; Bhattarai B.; Drabold D. A. Realistic inversion of diffraction data for an amorphous solid: The case of amorphous silicon. Phys. Rev. B 2016, 94 (23), 235208. 10.1103/PhysRevB.94.235208. [DOI] [Google Scholar]
- Igram D.; Bhattarai B.; Biswas P.; Drabold D. A. Large and realistic models of amorphous silicon. J. Non Cryst. Solids 2018, 492, 27–32. 10.1016/j.jnoncrysol.2018.04.011. [DOI] [Google Scholar]
- Barkema G. T.; Mousseau N. Event-Based Relaxation of Continuous Disordered Systems. Phys. Rev. Lett. 1996, 77 (21), 4358–4361. 10.1103/PhysRevLett.77.4358. [DOI] [PubMed] [Google Scholar]
- Malek R.; Mousseau N. Dynamics of lennard-jones clusters: A characterization of the activation-relaxation technique. Phys. Rev. E Stat. Phys. Plasmas Fluids Relat. Interdiscip. Topics 2000, 62 (6 Pt A), 7723–7728. 10.1103/PhysRevE.62.7723. [DOI] [PubMed] [Google Scholar]
- Mousseau N.; Barkema G. T. Traveling through potential energy landscapes of disordered materials: The activation-relaxation technique. Phys. Rev. E 1998, 57 (2), 2419–2424. 10.1103/PhysRevE.57.2419. [DOI] [Google Scholar]
- Kallel H.; Mousseau N.; Schiettekatte F. Evolution of the potential-energy surface of amorphous silicon. Phys. Rev. Lett. 2010, 105 (4), 045503. 10.1103/PhysRevLett.105.045503. [DOI] [PubMed] [Google Scholar]
- Mousseau N.; Lewis L. J. Topology of Amorphous Tetrahedral Semiconductors on Intermediate Length Scales. Phys. Rev. Lett. 1997, 78 (8), 1484–1487. 10.1103/PhysRevLett.78.1484. [DOI] [Google Scholar]
- Mousseau N.; Barkema G. T.; de Leeuw S. W. Elementary mechanisms governing the dynamics of silica. J. Chem. Phys. 2000, 112 (2), 960–964. 10.1063/1.480621. [DOI] [Google Scholar]
- Mousseau N.; Barkema G. T. Activated mechanisms in amorphous silicon: An activation-relaxation-technique study. Phys. Rev. B 2000, 61 (3), 1898–1906. 10.1103/PhysRevB.61.1898. [DOI] [Google Scholar]
- Grigera T. S.; Parisi G. Fast Monte Carlo algorithm for supercooled soft spheres. Phys. Rev. E Stat. Nonlin. Soft. Matter. Phys. 2001, 63 (4 Pt 2), 045102. 10.1103/PhysRevE.63.045102. [DOI] [PubMed] [Google Scholar]
- Berthier L.; Biroli G.; Bouchaud J.-P.; Tarjus G. Can the glass transition be explained without a growing static length scale?. J. Chem. Phys. 2019, 150 (9), 094501. 10.1063/1.5086509. [DOI] [PubMed] [Google Scholar]
- Brumer Y.; Reichman D. R. Numerical Investigation of the Entropy Crisis in Model Glass Formers. J. Phys. Chem. B 2004, 108 (21), 6832–6837. 10.1021/jp037617y. [DOI] [Google Scholar]
- Fernández L. A.; Martín-Mayor V.; Verrocchio P. Optimized Monte Carlo method for glasses. Philos. Mag. 2007, 87 (3–5), 581–586. 10.1080/14786430600919302. [DOI] [Google Scholar]
- Ninarello A.; Berthier L.; Coslovich D. Models and Algorithms for the Next Generation of Glass Transition Studies. Phys. Rev. X 2017, 7 (2), 021039. 10.1103/PhysRevX.7.021039. [DOI] [Google Scholar]
- Scalliet C.; Guiselin B.; Berthier L. Thirty Milliseconds in the Life of a Supercooled Liquid. Phys. Rev. X 2022, 12 (4), 041028. 10.1103/PhysRevX.12.041028. [DOI] [Google Scholar]
- Gutiérrez R.; Karmakar S.; Pollack Y. G.; Procaccia I. The static lengthscale characterizing the glass transition at lower temperatures. EPL 2015, 111 (5), 56009. 10.1209/0295-5075/111/56009. [DOI] [Google Scholar]
- Parmar A. D. S.; Ozawa M.; Berthier L. Ultrastable Metallic Glasses In Silico. Phys. Rev. Lett. 2020, 125 (8), 085505. 10.1103/PhysRevLett.125.085505. [DOI] [PubMed] [Google Scholar]
- Fullerton C. J.; Berthier L. Density controls the kinetic stability of ultrastable glasses. EPL 2017, 119 (3), 36003. 10.1209/0295-5075/119/36003. [DOI] [Google Scholar]
- Bojesen T. A. Policy-guided Monte Carlo: Reinforcement-learning Markov chain dynamics. Phys. Rev. E 2018, 98 (6), 063303. 10.1103/PhysRevE.98.063303. [DOI] [Google Scholar]
- Gabrie M.; Rotskoff G. M.; Vanden-Eijnden E. Adaptive Monte Carlo augmented with normalizing flows. Proc. Natl. Acad. Sci. U. S. A. 2022, 119 (10), e2109420119 10.1073/pnas.2109420119. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Galliano L.; Rende R.; Coslovich D. Policy-guided Monte Carlo on general state spaces: Application to glass-forming mixtures. J. Chem. Phys. 2024, 161 (6), 064503. 10.1063/5.0221221. [DOI] [PubMed] [Google Scholar]
- Deringer V. L.; Caro M. A.; Csanyi G. Machine Learning Interatomic Potentials as Emerging Tools for Materials Science. Adv. Mater. 2019, 31 (46), e1902765 10.1002/adma.201902765. [DOI] [PubMed] [Google Scholar]
- Caro M. A. Machine learning based modeling of disordered elemental semiconductors: understanding the atomic structure of a-Si and a-C. Semicond. Sci. Technol. 2023, 38 (4), 043001. 10.1088/1361-6641/acba3d. [DOI] [Google Scholar]
- Behler J.; Parrinello M. Generalized neural-network representation of high-dimensional potential-energy surfaces. Phys. Rev. Lett. 2007, 98 (14), 146401. 10.1103/PhysRevLett.98.146401. [DOI] [PubMed] [Google Scholar]
- Zuo Y.; et al. Performance and Cost Assessment of Machine Learning Interatomic Potentials. J. Phys. Chem. A 2020, 124 (4), 731–745. 10.1021/acs.jpca.9b08723. [DOI] [PubMed] [Google Scholar]
- Mishin Y. Machine-learning interatomic potentials for materials science. Acta Mater. 2021, 214, 116980. 10.1016/j.actamat.2021.116980. [DOI] [Google Scholar]
- Wang G.; Wang C.; Zhang X.; Li Z.; Zhou J.; Sun Z. Machine learning interatomic potential: Bridge the gap between small-scale models and realistic device-scale simulations. iScience 2024, 27 (5), 109673. 10.1016/j.isci.2024.109673. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Khorshidi A.; Peterson A. A. Amp: A modular approach to machine learning in atomistic simulations. Comput. Phys. Commun. 2016, 207, 310–324. 10.1016/j.cpc.2016.05.010. [DOI] [Google Scholar]
- Bartók A. P.; Kermode J.; Bernstein N.; Csányi G. Machine Learning a General-Purpose Interatomic Potential for Silicon. Phys. Rev. X 2018, 8 (4), 195207. 10.1103/PhysRevX.8.041048. [DOI] [Google Scholar]
- Thompson A. P.; Swiler L. P.; Trott C. R.; Foiles S. M.; Tucker G. J. Spectral neighbor analysis method for automated generation of quantum-accurate interatomic potentials. J. Comput. Phys. 2015, 285, 316–330. 10.1016/j.jcp.2014.12.018. [DOI] [Google Scholar]
- Shapeev A. V. Moment Tensor Potentials: A Class of Systematically Improvable Interatomic Potentials. Multiscale Model. Sim. 2016, 14 (3), 1153–1173. 10.1137/15M1054183. [DOI] [Google Scholar]
- Deringer V. L.; Bernstein N.; Bartok A. P.; Cliffe M. J.; Kerber R. N.; Marbella L. E.; Grey C. P.; Elliott S. R.; Csanyi G. Realistic Atomistic Structure of Amorphous Silicon from Machine-Learning-Driven Molecular Dynamics. J. Phys. Chem. Lett. 2018, 9 (11), 2879–2885. 10.1021/acs.jpclett.8b00902. [DOI] [PubMed] [Google Scholar]
- Bartók A. P.; Kermode J.; Bernstein N.; Csányi G. Machine Learning a General-Purpose Interatomic Potential for Silicon. Phys. Rev. X 2018, 8 (4), 041048. 10.1103/PhysRevX.8.041048. [DOI] [Google Scholar]
- Bernstein N.; Bhattarai B.; Csanyi G.; Drabold D. A.; Elliott S. R.; Deringer V. L. Quantifying Chemical Structure and Machine-Learned Atomic Energies in Amorphous and Liquid Silicon. Angew. Chem., Int. Ed. Engl. 2019, 58 (21), 7057–7061. 10.1002/anie.201902625. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Erhard L. C.; Rohrer J.; Albe K.; Deringer V. L. A machine-learned interatomic potential for silica and its relation to empirical models. npj Comput. Mater. 2022, 8 (1), 90. 10.1038/s41524-022-00768-w. [DOI] [Google Scholar]
- Erhard L. C.; Rohrer J.; Albe K.; Deringer V. L. Modelling atomic and nanoscale structure in the silicon–oxygen system through active machine learning. Nat. Commun. 2024, 15 (1), 1927. 10.1038/s41467-024-45840-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Drautz R. Atomic cluster expansion for accurate and transferable interatomic potentials. Phys. Rev. B 2019, 99 (1), 249901. 10.1103/PhysRevB.99.014104. [DOI] [Google Scholar]
- Zongo K.; Sun H.; Ouellet-Plamondon C.; Béland L. K. A unified moment tensor potential for silicon, oxygen, and silica. npj Comput. Mater. 2024, 10 (1), 218. 10.1038/s41524-024-01390-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zongo K.; Béland L. K.; Ouellet-Plamondon C. First-principles database for fitting a machine-learning silicon interatomic force field. MRS Adv. 2022, 7 (2–3), 39–47. 10.1557/s43580-022-00228-z. [DOI] [Google Scholar]
- Morrow J. D.; Ugwumadu C.; Drabold D. A.; Elliott S. R.; Goodwin A. L.; Deringer V. L. Understanding Defects in Amorphous Silicon with Million-Atom Simulations and Machine Learning. Angew. Chem., Int. Ed. Engl. 2024, 63 (22), e202403842 10.1002/anie.202403842. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Atta-Fynn R.; Biswas P. Nearly defect-free dynamical models of disordered solids: The case of amorphous silicon. J. Chem. Phys. 2018, 148 (20), 204503. 10.1063/1.5021813. [DOI] [PubMed] [Google Scholar]
- Comin M.; Lewis L. J. Deep-learning approach to the structure of amorphous silicon. Phys. Rev. B 2019, 100 (9), 094107. 10.1103/PhysRevB.100.094107. [DOI] [Google Scholar]
- Mills K.; Casert C.; Tamblyn I. Adversarial Generation of Mesoscale Surfaces from Small-Scale Chemical Motifs. J. Phys. Chem. C 2020, 124 (42), 23158–23163. 10.1021/acs.jpcc.0c06673. [DOI] [Google Scholar]
- Zhou Z.; Shang Y.; Liu X.; Yang Y. A generative deep learning framework for inverse design of compositionally complex bulk metallic glasses. npj Comput. Mater. 2023, 9 (1), 15. 10.1038/s41524-023-00968-y. [DOI] [Google Scholar]
- Invernizzi M.; Kramer A.; Clementi C.; Noe F. Skipping the Replica Exchange Ladder with Normalizing Flows. J. Phys. Chem. Lett. 2022, 13 (50), 11643–11649. 10.1021/acs.jpclett.2c03327. [DOI] [PubMed] [Google Scholar]
- Kilgour M.; Gastellu N.; Hui D. Y. T.; Bengio Y.; Simine L. Generating Multiscale Amorphous Molecular Structures Using Deep Learning: A Study in 2D. J. Phys. Chem. Lett. 2020, 11 (20), 8532–8537. 10.1021/acs.jpclett.0c02535. [DOI] [PubMed] [Google Scholar]
- Madanchi A.; Kilgour M.; Zysk F.; Kühne T. D.; Simine L. Simulations of Disordered Matter in 3D with the Morphological Autoregressive Protocol (MAP) and Convolutional Neural Networks. J. Chem. Phys. 2024, 024101. 10.1063/5.0174615. [DOI] [PubMed] [Google Scholar]
- Wang H., Jiang Z.; Yi L.; Mo K.; Su H.; Guibas L.. Rethinking Sampling in 3D Point Cloud Generative Adversarial Networks. arXiv, 12 Jun 2020. 10.48550/arXiv.2006.07029 (accessed 2024-November-24). [DOI]
- Charles R. Q., Su H., Kaichun M., Guibas L. J.. PointNet: Deep Learning on Point Sets for 3D Classification and Segmentation. In IEEE Conference on Computer Vision and Pattern Recognition (CVPR). 2017 Conference. 10.1109/CVPR.2017.16. [DOI]
- Zhou Q.; Du T.; Guo L.; Smedskjaer M.M.; Bauchy M. New insights into the structure of sodium silicate glasses by force-enhanced atomic refinement. J. Non-Cryst. Solids 2020, 536, 120006. 10.1016/j.jnoncrysol.2020.120006. [DOI] [Google Scholar]
- Santos I.; Aboy M.; Marqués L. A.; López P.; Pelaz L. Generation of amorphous Si structurally compatible with experimental samples through the quenching process: A systematic molecular dynamics simulation study. J. Non-Cryst. Solids 2019, 503–504, 20–27. 10.1016/j.jnoncrysol.2018.09.024. [DOI] [Google Scholar]
- Cooper N. C.; Goringe C. M.; McKenzie D. R. Density functional theory modelling of amorphous silicon. Comput. Mater. Sci. 2000, 17 (1), 1–6. 10.1016/S0927-0256(99)00037-3. [DOI] [Google Scholar]
- Marques L. A.; Pelaz L.; Santos I.; Bail6n L.; Barbolla J.. Study of the amorphous phase of silicon using molecular dynamics simulation techniques. In Conference on Electron Devices. 2005. Conference. 10.1109/SCED.2005.1504416. [DOI]
- Sundararaman S.; Huang L.; Ispas S.; Kob W. New optimization scheme to obtain interaction potentials for oxide glasses. J. Chem. Phys. 2018, 148 (19), 194504. 10.1063/1.5023707. [DOI] [PubMed] [Google Scholar]
- Giacovazzo C.Fundamentals of crystallography; Oxford university press, 2002; Vol. 7. [Google Scholar]
- Pecharsky V. K.; Zavalij P. Y.. Fundamentals of Powder Diffraction and Structural Characterization of Materials; Springer: New York, 2009. [Google Scholar]
- Zhang N.; Yokota H.; Glazer A. M.; Thomas P. A. Neutron powder diffraction refinement of PbZr(1 - x)Ti(x)O3. Acta. Crystallogr. B 2011, 67 (Pt 5), 386–398. 10.1107/S0108768111028631. [DOI] [PubMed] [Google Scholar]
- Kisi E. H.; Howard C.J.. Applications of Neutron Powder Diffraction; Oxford University Press, 2008. [Google Scholar]
- Li W.; Ando Y. Comparison of different machine learning models for the prediction of forces in copper and silicon dioxide. Phys. Chem. Chem. Phys. 2018, 20 (47), 30006–30020. 10.1039/C8CP04508A. [DOI] [PubMed] [Google Scholar]
- Balyakin I. A.; Rempel S. V.; Ryltsev R. E.; Rempel A. A. Deep machine learning interatomic potential for liquid silica. Phys. Rev. E 2020, 102 (5–1), 052125. 10.1103/PhysRevE.102.052125. [DOI] [PubMed] [Google Scholar]
- Gereben O.; Pusztai L. Structure of amorphous semiconductors: Reverse Monte Carlo studies on a-C, a-Si, and a-Ge. Phys. Rev. B. Condens. Matter. 1994, 50 (19), 14136–14143. 10.1103/PhysRevB.50.14136. [DOI] [PubMed] [Google Scholar]
- Biswas P.; Tafen D. N.; Atta-Fynn R.; Drabold D. The inclusion of experimental information in first principles modelling of materials. J. Phys.: Condens. Matter 2004, 16 (44), S5173–S5182. 10.1088/0953-8984/16/44/014. [DOI] [Google Scholar]
- Opletal G.; Petersen T. C.; O’Malley B.; Snook I. K.; McCulloch D. G.; Yarovsky I. HRMC: Hybrid Reverse Monte Carlo method with silicon and carbon potentials. Comput. Phys. Commun. 2008, 178 (10), 777–787. 10.1016/j.cpc.2007.12.007. [DOI] [Google Scholar]
- Opletal G.; Petersen T. C.; McCulloch D. G.; Snook I. K.; Yarovsky I. The structure of disordered carbon solids studied using a hybrid reverse Monte Carlo algorithm. J. Phys.: Condens. Matter 2005, 17 (17), 2605–2616. 10.1088/0953-8984/17/17/008. [DOI] [Google Scholar]
- Singh R.; ZetterstrÖM P. E. R.; Mohan I.; Ahluwalia P. K. Structural Properties of Amorphous Silicon Models Generated with Reverse Monte-Carlo Method. International Journal of Modern Physics B 2006, 20 (07), 779–790. 10.1142/S0217979206033619. [DOI] [Google Scholar]
- Tucker M. G.; Keen D. A.; Dove M. T.; Trachenko K. Refinement of the Si–O–Si bond angle distribution in vitreous silica. J. Phys.: Condens. Matter 2005, 17 (5), S67–S75. 10.1088/0953-8984/17/5/008. [DOI] [Google Scholar]
- Biswas P.; Tafen D. N.; Drabold D. A. Experimentally constrained molecular relaxation: The case of glassyGeSe2. Phys. Rev. B 2005, 71 (5), 054204. 10.1103/PhysRevB.71.054204. [DOI] [Google Scholar]
- Treacy M. M.; Borisenko K. B. The local structure of amorphous silicon. Science 2012, 335 (6071), 950–953. 10.1126/science.1214780. [DOI] [PubMed] [Google Scholar]