Citation: Martin O. Steinhauser. Multiscale modeling, coarse-graining and shock wave computer simulations in materials science[J]. AIMS Materials Science, 2017, 4(6): 1319-1357. doi: 10.3934/matersci.2017.6.1319
[1] | Koufos Evan, Muralidharan Bharatram, Dutt Meenakshi . Computational Design of Multi-component Bio-Inspired Bilayer Membranes. AIMS Materials Science, 2014, 1(2): 103-120. doi: 10.3934/matersci.2014.2.103 |
[2] | Michael Sebastiano, Xiaolei Chu, Fikret Aydin, Leebyn Chong, Meenakshi Dutt . Interactions of Bio-Inspired Membranes with Peptides and Peptide-Mimetic Nanoparticles. AIMS Materials Science, 2015, 2(3): 303-318. doi: 10.3934/matersci.2015.3.303 |
[3] | Mica Grujicic, Jennifer Snipes, S. Ramaswami . Meso-scale computational investigation of polyurea microstructure and its role in shockwave attenuation/dispersion. AIMS Materials Science, 2015, 2(3): 163-188. doi: 10.3934/matersci.2015.3.163 |
[4] | Grujicic Mica, Ramaswami S., S. Snipes J., Yavari R. . Multi-Scale Computation-Based Design of Nano-Segregated Polyurea for Maximum Shockwave-Mitigation Performance. AIMS Materials Science, 2014, 1(1): 15-27. doi: 10.3934/matersci.2014.1.15 |
[5] | Mica Grujicic, Jennifer S. Snipes, S. Ramaswami . Polyurea/Fused-silica interfacial decohesion induced by impinging tensile stress-waves. AIMS Materials Science, 2016, 3(2): 486-507. doi: 10.3934/matersci.2016.2.486 |
[6] | C. Aris Chatzidimitriou-Dreismann . Quantumness of correlations in nanomaterials—experimental evidence and unconventional effects. AIMS Materials Science, 2022, 9(3): 382-405. doi: 10.3934/matersci.2022023 |
[7] | Shaohua Chen, Yaopengxiao Xu, Yang Jiao . Modeling solid-state sintering with externally applied pressure: a geometric force approach. AIMS Materials Science, 2017, 4(1): 75-88. doi: 10.3934/matersci.2017.1.75 |
[8] | Jing Chen, Ben J. Hanson, Melissa A. Pasquinelli . Molecular Dynamics Simulations for Predicting Surface Wetting. AIMS Materials Science, 2014, 1(2): 121-131. doi: 10.3934/matersci.2014.2.121 |
[9] | Julian Konrad, Dirk Zahn . Assessing the mechanical properties of molecular materials from atomic simulation. AIMS Materials Science, 2021, 8(6): 867-880. doi: 10.3934/matersci.2021053 |
[10] | Elena Kossovich . Theoretical study of chitosan-graphene and other chitosan-based nanocomposites stability. AIMS Materials Science, 2017, 4(2): 317-327. doi: 10.3934/matersci.2017.2.317 |
A particularly fascinating problem in science is the description of phenomena involving multiple spatial or temporal scales which together determine the behavior of a system. Many problems in materials science are connected with this structure-property paradigm. The atomic interactions at the microscopic level on the scale of nanometers and femtoseconds fully determine material behavior at the macroscopic scale (on the order of centimeters and milliseconds and beyond). The idea of performing material simulations across several characteristic length and time scales has therefore obvious appeal not only for fundamental research but also as a tool for technological innovation, see e.g., the discussions in [1,2,3].
For example, in engineering, understanding the key mechanisms of failure in materials on various length-and time scales is a prerequisite for the design of new materials with desired superior properties such as high strength or toughness. As "material" we understand all condensed forms of matter, i.e., a state of matter, where the electrons are not separated from the atoms (a state which is called "plasma") and thus either form hard matter, i.e., crystalline solids, or soft matter, i.e., biological macromolecules and cells. Most of biological cells are about 1–100 microns in size, and all cells use a plasma membrane to separate and protect their chemical components from the outside environment. The plasma membrane acts as a barrier that separates the interior of the cell from its exterior, preventing the contents of the cell from mixing with the surrounding medium. It also allows certain molecules to pass inward—across the membrane through channel—and waste products to pass out. A plasma membrane is very flexible and deforms without tearing when a cell grows and changes shape.
The interior of the cell is made up of a liquid phase (cytosol), a nucleus, the cytoskeleton consisting of many networks of microtubules, actin and intermediate filaments, organelles of different sizes and shapes, and other proteins. The resistance of single cells to elastic deformation, as quantified by a huge range of the effective elastic modulus as displayed in Figure 1, is orders of magnitude smaller than that of typical engineering materials such as metals and ceramics. This is the reason why cells and polymers are "soft" and governed in their behavior mostly by the interplay of entropy and energy. If a cell's plasma membrane is penetrated, it neither simply collapses, nor does it remain torn; instead, it has the capability of resealing as has been shown in several single cell experiments [4]. This transient permeability of a cell's plasma membrane can also be achieved by exposing cells to shock waves, as I have demonstrated in a recent publication where a new experimental set-up for exposing U87 glioblastoma tumor cells to laser-induced shock waves was introdced [5]. A shock wave treatment of cells might make possible new interesting biological and medical applications, e.g., in gene therapy, where the permeabilization of the membrane aims at delivering therapeutic molecules, genes or other genetic material into the cells' cytoplasm [6,7,8,9,10,11,12,13,14,15,16].
This review article is organized as follows: In Section 2, I provide a review of the basics of shock wave physics. In Section 3, I discuss different aspects of multiscale modeling of materials. Section 4 introduces the concept of coarse-graining which is often (but not solely) used for modeling soft matter systems on the mesoscopic scale. Then, in Section 5, I discuss several recent applications that illustrate how different modeling approaches may be usefully combined in detailed computational studies. Finally, in Section 6, a few general conclusions are drawn and several considerations are offered.
The theoretical foundations of shock wave physics originated from 17th-century studies of classical acoustics and 18th-century aeroballistics. With the emergence of high-speed photography in late 19th century and further progress in experimental visualization techniques in 20th century, the study of shock waves in different states of matter has gradually emerged from a very small and unnoticed branch of physics to a major complex and interdisciplinary science [18,19].
Shock waves are in essence discontinuous, entropy-increasing, rapid mechanical phenomena that have been observed and studied in the laboratory and in nature, in microscopic as well as in macroscopic dimensions and in all states of matter: gaseous [20,21], liquid [22,23,24], solid [25,26,27], plasma [28,29], and even in Bose-Einstein condensates [30,31,32]. Terrestrial examples [33] of naturally occurring discontinuities are high-energy events such as volcanic explosions, thunder, meteorite impacts, sea-and earthquakes or tsunamis, while in outer space [34,35] they encompass plasma shock waves induced by solar wind, supernovae explosions, implosions of white dwarfs, comet and asteroid impacts, and stellar or galactic jets. This variety of phenomena and in particular the possible applications of shock waves in different areas such as biology, chemistry, physics, medicine and even engineering renders the scientific study of shock waves a rather fascinating subject.
A shock wave is a mechanical wave characterized by an area or sheet of discontinuity in which, within a narrow region, thermodynamic quantities such as pressure p, density ρ, particle velocity →v or temperature T change abruptly. While a theory of shock waves with an infinitesimal jump condition can be developed mathematically in the framework of approximative continuum models of condensed matter, i.e., based on the hydrodynamic equations of ideal fluids or gases, in physical applications, shock wave theory is somewhat limited, because the width of a genuine shock wave is always of finite size and because of the actually discrete atomic nature of condensed matter. It turns out that shock waves are in essence small regions were non-adiabatic, i.e., irreversible energy dissipation occurs and for which the shock wave is a mathematical idealization. The width of the shock wave, i.e., the size of the dissipative region, establishes itself according to the conservation laws of continuum theory: the conservation of mass, momentum and energy. In very strong, high-amplitude shock waves, their width becomes so small that they are practically indistinguishable from the mathematical idealization of an infinitesimally small perturbation with jump conditions for thermodynamic variables, see Figure 2.
In solids, physical shock waves are mechanical waves of finite amplitudes and arise when condensed matter is subjected to a rapid compression. Shock waves can be defined by several major distinctive properties as follows:
• the formation of a steep wave front with a sudden, irreversible change of thermodynamic quantities;
• a pressure-dependent, supersonic velocity of propagation;
• the creation of entropy;
• non-linear superposition properties (for reflection and interaction).
Sometimes another criterion for a shock wave is listed in shock wave literature which is an extremely short rise time of the pressure within tens of nanoseconds. However, this is not a defining criterion of shock waves, because no generally agreed upon definition exists of what "extremely short" really is supposed to mean—it is a term of rather subjective nature. A nanosecond rise time certainly applies for lithotripter shock waves which are used in medicine as a non-invasive technique to comminute calculi [36] and for other therapeutic purposes [37]; however, in the geologic realm for example, meteorite impacts, explosive volcanic eruptions and earthquakes can provoke drastic and irreversible changes within seconds which is nine orders of magnitude slower.
One thing that distinguishes a shock wave from an ordinary sound wave is that the initial disturbance in the medium that causes a shock wave is always traveling at a velocity greater than the phase velocity of sound in the medium. The electromagnetic analog of this is called Cherenkov radiation [38], which is caused by a charged particle traveling through a dielectric medium at a velocity faster than the speed of light in that medium. Because a shock wave moves faster than the speed of sound, the medium ahead of the shock front cannot respond until the shock strikes, and so the shock wave falling upon the particles of matter initially at rest, is a supersonic phenomenon. In a steady flow with constant velocity →v the speed at which a perturbation travels relative to a reference frame in which the laboratory is at rest, is composed of two parts, see Figure 3. On the one hand the perturbation travels with velocity →vs=→c relative to the fluid into a direction →n. On the other hand, the source of disturbance moves at the same time with the velocity →v of the fluid flow. Figure 4 exhibits one of the first ever published schlieren picture of the bow shock wave generated by a projectile moving at supersonic speed. The two vertical lines visible in Figure 4 are wires with a spark gap inserted for triggering the spark light source. The projectile had a speed of v0≈530m/s which resulted in a Mach cone angle of α≈40∘ (cf. Figure 5b).
For simplicity, we assume here constant flow velocity →v and consider a perturbation in the fluid at a fixed point O. This perturbation propagates with velocity →v+c→n from O. We can see in Figure 5 that the velocity depends on the direction of →n. Vector →n is a radial unit vector starting from point O with respect to the moving fluid flow. The possible values of the velocity lie on the surface of a sphere according to Figure 5. If vs<c, the vectors →v+c→n can attain any direction in space, cf. Figure 5a. Hence, a flow that propagates with subsonic speed will affect the whole domain of the fluid. For vs>c, the velocities →v+c→n lie within a cone the tip of which is located in point O, cf. Figure 5b. This cone is the envelope of the sphere, so we find for the total cone angle, the so-called Mach angle α the following relation:
sinα=cv. | (1) |
The ratio
Ma=vc | (2) |
is called Mach number. For Ma>1, the flow is supersonic and for Ma<1 it is subsonic, respectively.
Hence, we find that a perturbation travels only with supersonic speed in the direction of flow within the boundaries of a cone with angle 2α. Angle α gets smaller with increasing flow speed v, so the perturbation in a flow has no effect in the region outside that cone. The boundary layer of the region which can be influenced by a perturbation is called Mach's cone, characteristic lines or simply characteristics. The parts c) and d) in Figure 5 exhibit the situation in a fluid at rest through which a perturbation moves with velocity →v.
Another difference between a shock wave and an ordinary sound wave is that the entropy is increased in a shock wav—hence, in contrast to ordinary sound waves, shock waves constitute an irreversible process. In engineering applications, shock waves produced in air by an explosion and radiating outward from its center are termed blast waves, because they cause a strong wind, while the term shock wave is preferred for such waves occurring in water or the ground, because here the effect is like that of a sudden impact. Impact experiments are often performed to study the fracture and failure behavior of solids or in an attempt to determine a shock equation of state, describing a material under extreme conditions of high pressure and/or temperature. Along with phenomenological constitutive equations, this allows for numerically analyzing a material in a shocked state, e.g., in a finite element analysis, cf. Figure 6.
Upon hydrostatic compression, a solid shrinks in volume and the pressure increases [41]. The hydrostatic curve is smooth as a function of stress and volume. If the solid is compressed uniaxially, stress and pressure also increase until the maximal shear stress of the material is reached. Then it will yield and change from the uniaxially compressed state to hydrostatic compression. Many solid materials under moderate shock loading exhibit this two-wave structure where the first one is essentially an elastic wave followed by a slower plastic deformation wave front, which can be seen in Figure 6, too. In the p−V-diagram this transition from elastic to plastic behavior is visible through a cusp, called the Hugoniot elastic limit (HEL), and there are a considerable amount of experimental data which illustrate these elastic–plastic–shock phenomena (EPSP) with an elastic precursor wave [42].
Very often, impact experiments are used as test cases for the predictions obtained from numerical multiscale models—such as the ones presented in Figure 6—which simulate a material specimen that can be macroscopic in size and take several micro-structural features of the specific solid into account. Another important use of impact experiments in the laboratory is to study the crack initiation and propagation until ultimate failure occurs in a set-up that allows for very reproduciblex shock wave generation. One of these set-ups is the so-called edge-on-impact test, see Figure 7, where an impactor strikes the edge of a material specimen at high speed, initiating a well-defined plane shock wave through the material that ultimately leads to fracture and destruction.
Another important application of research in shock wave physics is the investigation of hypervelocity impact phenomena (with relative speeds of objects beyond ∼2 km/s) which occur in impact events in space. Here, scientists try to understand via novel computer simulations and impact experiments on a laboratory scale the physics behind the formation of space debris clouds in the low Earth orbit (LEO) [44]. Such clouds form when small debris objects from former space craft and satellite missions impact a satellite in its orbit at hypervelocity speed. These types of investigations also contribute to the assessment of the socio-economic consequences that are involved in the risks in space from orbiting debris.
Finally, also phase transitions in solids can be induced using shock waves and have been studied experimentally and numerically using classical molecular dynamics simulations in a variety of materials [26,45,46,47,48]. The world record for the largest ever MD simulation of shock waves in a solid is currently still held by Germann and Kadau [49] from Lawrence Livermore National Laboratories, who simulated a simple cubic crystal consisting of 1 trillion (1012) atoms, treated as classical Lennard-Jones particles. This corresponds to a chunk of matter of edge length of about 2.5μm.
In recent years, we have seen a very large increase in computing power and—at the same time— progress in developing efficient algorithms. As a result, computer simulations have become very popular and are used today in almost all areas of research [50,51,52,53,54]. Fast parallelized algorithms allow for the solution of non-linear many body problems directly, without any mathematical approximations involved. Usually, in analytic theory, practically all non-trivial problems are way too complex to be solved with just pencil and paper. In this sense, computer simulations can provide a link between experiment and analytic theory, thus allowing to test theories. But simulations can also be used as an exploratory research tool under physical conditions not feasible in real experiments in a laboratory. Thus, doing science with the help of computer systems has established a new, interdisciplinary research approach which is often referred to as "Computational Physics" or "Computational Materials Science". This approach combines knowledge and methods of different research areas such as chemistry, biology, physics, mathematics, engineering and even medicine and allows for performing multiscale and multi-disciplinary simulations in situations relevant for real-world applications.
Nowadays, with fast computational tools at hand, simulations have become feasible that are of practical interest in engineering sciences for product design and testing. Many materials used in industrial manufacturing are very heterogeneous and can be characterized by different kind of defects, interfaces, and other microstructural features. As an example, in Figure 8 we display a Scanning Electron Microscope (SEM) photograph of the fracture surface of Aluminum Oxide (Al2O3) after planar impact load. In general, inorganic crystalline materials have structural features such as grain boundaries between crystals which are millimeters to micrometers in size, while on the atomic scale dislocations and defects such as vacancies occur. Hence, these structures have to be studied from a hierarchical perspective.
For example, simulations in material physics and chemistry often focus on the exploration of lattice and defect dynamics on the atomic scale using MD and Monte Carlo (MC) methods based on generalized force-fields (physical potentials), deduced from solving the non-relativistic Schrödinger equation for a limited number of atoms [55,56,57]. Figure 9 displays several methods commonly employed for atomic scale simulations along with their estimated maximum system size and the typical time scales that can be treated. The highest precision and transferability of methods is achieved with self-consistent first principles—so-called ab-initio—calculations which try to avoid completely any empirical fitting parameters.
Self-Consistent Field Theory (SCFT) is in essence based on the Hartree-Fock (HF) method which itself is a mean-field approximation (MFA), using MFA for energy calculations always leads to values which are larger than the exact energy values. Another approximation in HF calculations stems from the need to find an analytic expression for the wave function, such functionals are only known exactly for very few one-electron systems and therefore, usually some approximate functions are used instead. Normally, as a simplification, one uses as basis wave functions either plain waves, Slater type orbitals (STO) ∼exp(−ax), or Gaussian type orbitals (GTO) ∼exp(−ax2). The term "orbital" is just another name for denoting a one-electron wave function. Correlations, i.e., interactions between the electrons are treated within Møller-Plesett perturbation theory (MPn), where n is the order of correction, Configuration Interaction (CI), Coupled Cluster theory (CP), or other methods. As a whole, these methods are referred to as correlated calculations.
An alternative method to SCE for calculating ground state energies of molecules is Density Functional Theory (DFT). In DFT, the total energy of a system is not derived from a wave function but in terms of an approximate Hamiltonian and thus from a Local Density Approximation (LDA). DFT uses GTO potentials or plane waves as basis sets.
Tight Binding (TB) is a so-called semi-empirical method (see e.g., Porezag et al. [58] or Pettifort [59]) which uses an approximation of the Hamiltonian H by approximating or neglecting several terms (called Slater-Koster approximation), but re-parameterizing other parts of H in a way so as to yield the best possible agreement with experiments or ab initio simulations. In the simplest TB version, the Coulomb repulsion between electrons is neglected. Thus, in this approximation, there exists no correlation problem, but there is also no self-consistent procedure [60].
In classical MD methods, one integrates the classical Newtonian equations of motion for a large number of particles, respectively atoms, which are fully treated as classical point particles or as classical particles with tare volume. That is, MD in essence solved the classical N-body problem. A separate consideration of the motion of the electrons, as in the ab-initio methods, is not done. Interactions between particles are usually described via empirical, sometimes generic potentials which may be fitted to the results of experiments or of quantum mechanical calculations. Due to the increasing computing power of modern hardware, many-particle MD simulations taking into account the degrees of freedom of several billion atoms are nowadays feasible [54,61]. Molecular dynamics simulations of this kind using generic potential models for a solid enhanced our understanding of the basic processes that govern failure and crack behavior, such as the dynamical instability of crack tips [62,63], the limiting speed of crack propagation [64], the dynamics of dislocations [65], and the universal features of energy dissipation in fracture [66].
In contrast to the natural sciences, simulations of materials in the area of mechanical engineering typically focus on large-scale problems, where the N-body problem is often homogenized with approximate continuum methods, such as Smoothed Particle Hydrodynamics (SPH) or the Finite Element Method (FEM), utilizing additional, purely empirical constitutive relations [67]. The SPH method is simply a re-formulation of the hydrodynamics equations for (pseudo-)particles [68]. This methods suffers from certain important deficiencies such as convergence problems. Finally, the FEM is simply a continuum approximation of material structures, where the structural details are represented by a topologically connected mesh, where the mesh nodes are the integration points of the continuum conservation equations. Usually, this method reaches its size limitations very quickly and only meshes with a couple of million integration nodes for macroscopic structures can be simulated on large computer systems. One disadvantage of this method is the use of a mesh which is very bad for simulating large deformations in structures and which may lead to prohibitively small time steps in the numerical procedure. Also, modeling failure and fracture with this method is very awkward and leads to the introduction of artificial, unphysical corrections such as artificial viscosity in an attempt to stabilize the numerical procedure.
Generally speaking, in materials with small grain size the diffusion distances are small. Hence, processes governed by diffusion, such as sintering, are facilitated and can occur at lower temperatures than would otherwise be possible. Being able to predict the properties and performance under load of such materials is a major issue in current computational material research and for industrial product design. Unfortunately, the complexity of structural hierarchies in solids on different scales does not allow for selecting one single computational model or physical theory which allows for explaining all observed phenomena. Therefore, the different kind of microstructural features of different important classes of materials such as metals, ceramics, or materials pertaining to soft matter (glasses or polymers), have to be considered in different models, cf. Figure 10.
Solids comprise a span of length scales of roughly 10 to 12 orders of magnitude and the methods of classical Newtonian physics are sufficient to describe many of the occurring phenomena, cf. Figure 11. However, classical MC or MD methods lose their validity at length scales comparable to the typical size of atoms (≈10−10 m). In classical theories atoms are treated as point particles or spheres. In principle, the non-relativistic, time-dependent Schrödinger equation describes the properties of molecular systems with high accuracy, but anything more complex than the equilibrium state of a few atoms cannot be handled at this level. Therefore, approximations have to be introduced. The approximation are the more severe, the more complex a system is and the longer the time scale of the investigated processes are. At a certain point, the ab-initio approach has to be abandoned completely and replaced by empirical parameterizations of the used model. Hence, depending on the kind of problem that one investigates and depending on the desired accuracy with which specific structural features of the considered are to be resolved, one has the choice between many different models for different length and time scales. Unfortunately, there is no simple "hierarchy" that is connected with a natural length scale according to which the great diversity of simulation methods could be sorted out. For example, Monte Carlo lattice methods can be applied at the femtoscale of Quantumchromodynamics (10−15 m) [69], at the Å ngstrømscale (10−10 m) of solid state crystal lattices [70], or at the micrometerscale (10−6 m), simulating grain growth processes of polycrystal solid states [71].
The salient hierarchical structural features of materials have to be taken into account when developing mathematical and numerical models. Here, usually one of two possible strategies is pursued: In a "sequential modeling approach" one attempts to compile together a hierarchy of computational approaches in which large-scale models use the coarse-grained representations with information obtained from more detailed, smaller-scale models ("bottom-up" vs. "top-down" approach), see Figure 11 and also compare Figure 12 in Section 4.
This sequential modeling technique has been used effectively for systems in which the different scales are weakly coupled. The large majority of multiscale simulations are based on a sequential approach. Examples of this approach are abundant in literature and it is not my intention to comprehensively review the many publications in this field. The sequential approach includes practically all MD simulations the potentials of which are derived from ab-initio calculations, including classical coarse-grained simulations of macromolecules [72,73,74]. Due to the fractal nature of polymers, they exhibit self-similar structures on various length scales. This leads to unique and universal scaling properties which can be used to check the validity and quality of research codes. This "Scale-Hopping in Computer Simulations of Polymers" [75] is typical for coarse-grained simulations of soft matter systems.
The second strategy used in multiscale modeling and simulation is the "concurrent" or "parallel approach". Here, one links appropriate methods on each scale together in a combined model, where the different scales of the system are considered concurrently and often communicate with some type of hand-shaking procedure [76]. This approach is necessary for systems where the properties at each scale strongly depend on what happens at the other scales, for example dislocations, grain boundary structure, or dynamic crack propagation in polycrystalline materials [77,78].
The idea of using coarse-grained (CG) models for the description of molecular structures was originally introduced by Warshel and Levitt [79] in a 1976 pioneering paper which dealt with the structure of globular proteins. Since then, CG models have found their way into polymer physics as so-called bead-spring models, taking advantage of universal scaling laws of long polymer chains which are due to their fractal nature [80,81,82,83], as well as into chemistry, geophysics, engineering and other areas of computational research [43,84].
CG models of macromolecules provide a route to explore biomolecular systems on larger length and time scales while still resolving important physical aspects of the molecular topology, for example, the specific lipid bilayer structure of biological membranes [3,72,83,85,86,87,88,89]. They constitute a class of mesoscale models, in which many atoms or groups of atoms are treated by grouping them together into new particles which act as individual interaction sites usually connected by entropic springs, see Figures 12 and 13.
Such models have become popular [92,93] because they remain particle-based while at the same time reducing the computational cost. Another advantage of CG methods is that the groups of particles can be constructed with various levels of detail, thus permitting, e.g., the study of biological membranes on multiple length scales. A "bottom-up" CG model is a model of a particular system that is constructed on the basis of a more detailed model for the same system as indicated in Figure 12. In principle, the high-resolution, all-atom model may be based on atomistic data deduced from atomistic structure calculations. In contrast, "top-down" models do not rely upon or directly relate to a more detailed model for a particular system. Instead, they are usually related to the full complexity of the real experimental system by addressing observables on length scales that are accessible to the CG model. Often, these observables are thermodynamic averaged quantities such as pressure, temperature, stress and strains or forces accessible by direct experimental measurement. Figure 12 illustrates schematically these two major approaches to coarse-graining.
CG simulations involve less computational effort than atomistic simulations, because the number of interacting particles is reduced. Consequently, CG simulations are able to investigate much larger length-and time-scales than is possible in all-atom approaches, let alone in quantum chemical calculations [72,94,95]. CG models may simply remove certain degrees of freedom (e.g., vibrational modes between two atoms) or they may in fact simplify the two atoms completely via a single particle representation. The ends to which systems may be coarse-grained is simply bound by the accuracy in the dynamics and structural properties one wishes to replicate. This area of research is still basically in its infancy, and although it is commonly used in biological modeling, chemistry and polymer physics, the analytic theory behind it is still poorly developed.
Recent literature dealing with multiscale methods in polymer physics includes the edited volume Multiscale Molecular Methods in Applied Chemistry [74] and the treatment of multiscale computational methods in the monograph Computational Multiscale Modeling of Fluids and Solids—Theory and Applications [3]. Additionally, very recent publications on this subject can be found on the website of the special issue Computational Multiscale Modeling and Simulation in Materials Science of the Journal Materials [96].
The research area of CG simulations of polymers and biological macromolecular structures has seen major progress over the past 30 years, which is mostly due to the emergence of more physically based modeling approaches in the biological sciences, which allowed for the exploration of soft biomaterials, enhanced understanding of structures and diseases and which lead to the development of new medical treatments and diagnostic methods [97,98,99,100,101]. The major challenge that material scientists are facing in this research area is the extremely broad spectrum of length and time scales in the dynamics of biological macromolecules. Because of their large length, macromolecules can exist in many possible conformations that contribute to the total free energy of the system. In essence, the structural properties of macromolecules are determined by an interplay between entropic and energetic contributions. The hydrodynamic interactions of macromolecules with their surrounding solvents, covalent interactions, intra-and intermolecular interactions and—particularly in biological macromolecules—the Coulomb interactions between charged monomers along with hydrogen bonds add up with entropic forces to build a very complex system of interacting constituents. Because of their large range of characteristic time and length scales, macromolecules are used in many technological applications. The theoretical description of such systems has to take into account some of their hierarchical salient features to ensure equilibration on all relevant length and time scales and to efficiently sample their complex potential energy hypersurface.
CG models—as shown in Figures 12 and 13—are used because atomistic models of long polymer chains are usually intractable for time scales beyond nanoseconds, maybe microseconds on the largest computer systems; however, these larger time scales are important for many physical phenomena (for example diffusion processes) and for comparison with real experiments. Macromolecules, due to their large molecular weight exhibit fractal properties. Hence, numerical models of this type molecules are very useful for the investigation of fundamental scaling properties. In fractal systems, there exists some property P of a system (with polymers, e.g., the radius of gyration Rg) which obeys a relation P∝Nk, where N∈N is the size of the system, i.e., the number of monomers and k∈Q is the fractal dimension which is of the form k=a/b with a≠b, a∈N and b∈N. In order to extract the universal scaling properties of polymers with CG numerical models it is sufficient to ensure the following features:
• connectivity of monomers in a chain;
• The topological constraints, e.g., the impenetrability of chain bonds;
• The flexibility or stiffness of monomer bonds.
CG models of lipids and proteins have allowed for simulating lipoprotein assemblies and protein-lipid complexes for several microseconds using just a couple of coarse-grained beads for each amino acid of the protein. Without coarse-graining, such simulation would not have become possible.
Polymers are encountered in a large variety of chemical achievements, from biopolymers like Desoxyribonucleic Acid (DNA), cellulose, proteins, and actin filaments, to synthetized polymers like polyethylene (PE), polybutadiene (PB), polystyrene (PS), polymethylmethacrylate (PMMA)—which is used in the production of compact discs—which are all of great industrial interest. They are employed under numerous forms, in solutions, as gels, rubbers, synthetic fabrics, molded pieces, and so on.
Here, I limit myself to the description of main chain flexible polymers, whose prototypical monomer is polyethylene [-CH2-]n. This monomer is repeated many times in polymerization reactions to form what is called a polymer, i.e., a macromolecule that consists of many repetitions of one single monomer unit. Other standard examples are provided in Figure 14.
Every single one of these polymers can be represented as a series of bonds linking successive monomers; for the simple case of polyethylene (PE), the angular variation between neighboring carbon atoms is given by Θn−1=0, Δφn=[0,±120∘], see Figure 15.
The so-called trans configuration of polyethylene is characterized by Δφn=0 and Δφn=±120∘ denotes gauche configurations. Gauche and trans configurations differ from each other energetically. The geometric picture shown in Figure 15 for polyethylene can be extended to any type of homopolymer.
Assuming the successive values of Δφn to be random, the polymer chain appears, when observed at some scale much larger than the typical average bond length a of monomers, as a statistical coil. The persistence length Lp is then a measure for the degree of stiffness in such a polymer coil. Lp denotes the length scale on which the chain looks like a sequence of rigid collinear segments in trans configurations. Let Δϵ=ϵg−ϵt>0 be the difference in energy between the gauche configuration and the trans configuration. Then, the persistence length is given by
Lp=aexp(ΔϵkBT), | (3) |
where T denotes temperature and kB is the Boltzmann constant. The concept of Lp is based on the idea that the correlation between the orientations of the local tangent t(s) decays with the distance along the filament contour according to ⟨t(s)t(s′)⟩=exp(−|s−s′|/Lp). The total chemical length of the chain is given by L=Na, where N is the degree of polymerization, i.e., the number of repeat units (or monomers) in the polymer. L has to be large compared with Lp for the chain to be considered as a statistical coil, i.e., Lp/L≪1. The degree of polymerization N, i.e., the number of repeating monomer units in a polymer is sometimes also called the number of segments which is not to be confused with the bond vectors →ai with the convention, that index i refers to the position vector connecting segment number i−1 with segment number i. Thus, the counting of bond vectors starts with i=1 and ends with N−1. In a polymer model, one often just refers to the number N as the number of monomers, or—in computer simulations—the number of particles.
An example of polymers with vastly different persistence lengths is displayed in Figure 16 which shows a classification of polymer stiffness of the three major polymer components of the cytoskeleton of eukariotic cells, namely microtubules (stiff rods), actin filaments (semiflexible polymers) and intermediate filaments (flexible polymers).
Phenomenologically, the persistence length can be thought of as a characteristic length of the thermal fluctuations of an elastic rigid rod. Let us assume, for the sake of simplicity, that the cross-section of the rod is isotropic. Then, the free energy per unit length of the rod can be written as:
F=12κr−2c, | (4) |
where rc is the radius of curvature. We have r−1c=±|∂2→u/∂s2|=∂ω/∂s, where →u is the small displacement of the rod with respect to the unstrained straight rod, s is the length measured along the rod, and ω is the angle of deflection of the rod. lt is possible to show, by Fourier transforming the free energy F=∫L0κr2cds of a rod of length L, then applying the theorem of equipartition of energy, and finally summing over all modes, that
⟨ω2⟩=2kBTκL. | (5) |
Taking ⟨cosω⟩≈1−LkBT/κ equal to zero, which is the condition at which the elements of the rod at s=0 and s=L are decorrelated, one gets
Lp=κkBT. | (6) |
Thus, for L<Lp, the density of the free energy density can be written as
F=12kBTLpr−2c. | (7) |
We also mention the concept of a persistence time τp, i.e., a time characteristic of a change of conformation. Let ΔE be the activation energy necessary for the trans-to-gauche isomerization. Then we have
τp=τ0exp(ΔEkBT). | (8) |
The polymer chain appears as flexible during any observation time τ>τp. The characteristic time τ0 is related to the thermal vibrations of the chain, and is of the order of 10−12 s.
For a polymer chain there are three rotational isomeric states per bond: t, g+, and g−. Hence, a polymer chain can attain 3N different configurations. Among those, some are special, as follows:
• Helix Configurations. In a helix configuration, all monomers are repeated along a helical elongation of the chain. Helical configurations have minimal internal energy and are observed in crystalline polymers. For polyethylene, this is the all-trans state: The chain exhibits a zigzag structure along its elongation, rather than a true 3D helical structure. For polytetrafluoroethylene (PTFE), it is a true helix, not far from the all-trans state, but differing from it by a superimposed twist along the elongation, caused by the repulsive interactions between neighboring fluorines. For polyoxymethylene (POM), there are several minimal helical energy states, obtained by twisting the all-gauche configurations in slightly different ways.
• Coil Configurations. Here, we will occupy us only with molecules that are dynamically flexible, i.e., whose chemical length L is large compared to Lp, and whose observation is made on durations much larger than τp. These molecules run through the 3N configurational states. In such a case, a number of properties are not dependent on the local scales, and the chemical features can be ignored. On the other hand, all mesoscopic properties depend on N according to scaling laws, of the type
R=aNν, | (9) |
where R is some quantity that describes the spatial extension of a molecule.
Assuming that the N elementary segments of a polymer chain have no interaction and are independent one from the other. Such a chain is called ideal. The vector that points from one end of the chain to the other one is called end-to-end vector Re and can be written as
→Re=N∑i=1→ai. | (10) |
Taking the square of →Re, and averaging over all possible configurations yields the mean squared end-to-end vector:
⟨→Re⟩=N∑i=1→ai2+2N∑i≠j⟨→ai→aj⟩=N∑i=1→ai2+2|→ai||→aj|N∑i≠j⟨cos(∡(→ai,→aj)⟩⏟=0=Na2, | (11) |
where we have used the (ideal) condition that segments i and j and thus, →ai and →aj are statistically independent. Hence, the average separation of the ends, which can be viewed as the average size of the chain, is
Re=R0=⟨→Re2⟩1/2=aN1/2, | (12) |
and exponent ν=1/2. Subscript 0 in Eq. (12) indicates ideal chain behavior. For non-ideal chains, the expression Re is used for the average size. Note, that here, Re≠|→Re| as one might expect from the usual use of the same variable for denoting a vector →v and its absolute length |→v|.
The scaling law of Eq. (12) remains unchanged when introducing interactions between the different monomers along the same chain (generally called bonded interactions, in contrast to unbonded interactions, e.g., between monomers of different chains). One can show as a consequence of the central limit theorem that the probability PN(→r) for the N-link chain to terminate at →r=→Re is Gaussian distributed and given by:
PN(→r)=(32πR02)3/2exp(−32r2R02). | (13) |
In this section, we discuss briefly an example of CG modeling of semiflexible polymers. Semiflexibily, i.e., stiffness of a polymer chain is a hallmark of all biological macromolecules, which is why it has to be introduced into modeling schemes for biological macromolecular systems. This can be done with a bending potential that provides an energetic penalty in case the bond angle (defined by the angle between consecutive bond vectors connecting the monomers along a polymer chain) deviates from the preset angle.
In contrast to theoretical modeling, only very few simulation studies using mostly detailed atomistic models have been performed on the behavior of semiflexible polymer systems [102,103,104,105]. In particular the results of atomistic simulations have shown that the mean square particle displacements and the scattering functions deviate significantly from what is expected under ideal (Gaussian) conditions [102].
The Kratky-Porod chain model (or worm-like chain model, WLC) [106] is a simple description of semiflexible, inextensible polymers with spatial fluctuations that are governed by their bending energy Ubend
UBend=κ2∫L0ds(∂2→r∂s2)2 | (14) |
of the inextensible chain of length L depends on the local curvature of the chain contour s, where →r(s) is the position vector of a mass point (a monomer) on the chain and κ is a constant [107].
A dynamic equation for the WLC model was formulated by Harris and Hearst by applying Hamilton's principle with the constraint that the second moment of the total chain length be fixed. The resulting expressions for the bending →FBend and tension forces →FTens are:
→FBend=κ∂4→r∂s4, | (15) |
→FTens=β∂2→r∂s2. | (16) |
Applying this result to elastic light scattering, the model yields correct results in the flexible coil limit [108], but it fails at very high stiffness, deviating strongly from the theoretical solution obtained for rigid rods [109].
For computer simulations, a discrete polymer chain model has to be employed. Here, the bending rigidity, i.e., the stiffness of the chains which are composed of N mass points (monomers), is incorporated into the CG polymer model by the following bending potential
Ubend=κ2N−1∑i=1(→ui+1−→ui)2=κN−1∑i=1(1−→ui+1⋅→ui)=κN−1∑i=1(1−cosθi,i+1), | (17) |
where →u is the unit bond vector →ui=→ri+1−→ri/∣→ri+1−→ri∣, connecting consecutive monomers and →ri is the position vector to the i-th monomer. Parameter κ scales the strength of the bending energy in the model. In the discrete numerical model, Lp is fixed by parameter κ. A straightforward definition of the persistence length Lp in a discrete simulation model with respect to the bond length d0 is given by:
Lp=κd0kBT. | (18) |
In the simulations, Lp is varied in the range of Lp∈{5d0,10d0,20d0,50d0}, and is thus large compared to one bond length d0, which is the smallest length scale on which the model chains still exhibit flexibility. This range of Lp is also larger than the persistence lengths used in most previous MD studies of semiflexible chains which were in the range of Lp∈{1d0,...,5d0} [102,103,104,105]. Monodisperse chains with monomer numbers N=350 and N=700 are simulated. Thus, the semiflexible regime of polymer chains for which
d0≪Lp≪L=(N−1)d0 | (19) |
is investigated.
The simplest model for the description polymer chain dynamics in a melt, is the Rouse model [107]. It is a three parameter model, including the monomeric friction (ξ), the chain connectivity (modeled through harmonic springs with mean-squared bond length d02) and the degree of polymerization of the chains N. This model is known to describe the dynamics of short, unentangled melts reasonably well, though deviations appear at monomeric length scales, which are affected by local excluded volume interactions and chain stiffness. For long, entangled chains, the Rouse model describes the dynamics at intermediate time/length scales even though the longer scale dynamics are strongly affected by constraints formed by surrounding chains. The Rouse modes, p=0,1,2,...,N−1, of a chain of length N are defined as:
→Xp=√2NN∑i=1→ricos[pπN(i−12)]. | (20) |
The p=0 mode describes the motion of the chain center-of-mass, while the modes with p≤1 describe internal relaxations in polymer chains with a mode number p corresponding to a subchain of (N−1)/p segments. Hence, a Rouse mode analysis is in essence analogous to a Fourier analysis.
Starting from the Langevin Equation of the Rouse model,
ξ˙→r(s,t)=k∂2∂s2→r(s,t)+→FS(s,t), | (21) |
where s again is some parameter along the polymer chain that measures length—for details, see [107]—the stiffness of the polymer chains can be taken into account by introducing an additional entropic bending force →FB of the form
→FB(s,t)=−LpkBT∂4∂n4→r(s,t), | (22) |
which can be derived by applying Hamilton's principle to Eq. (14) [110]. The equation of motion then becomes
ξ∂∂t→r(s,t)=3kBT2lP∂2∂s2→r(s,t)−3kBTLp2∂4∂s4→r(s,t)+→FS(s,t), | (23) |
where →FS is a Gaussian stochastic force. Eq. (23) is similar to the equation of motion derived by Harris and Hearst [109] and by Harnau et al. [111] who, as in the Rouse model, assume a Gaussian distribution of each bond length. In essence, (23) is based on the Rouse model for flexible Gaussian chains and the introduced bending force can be considered as a small and local perturbation of the system. A solution of (23) can be achieved in terms of a normal Rouse mode analysis as shown in [112]:
⟨→˜Xp2(0)⟩=kBTkpsemi=[3π2LLpp2+3Lpπ4L3p4]−1, | (24) |
where the effective force constant of semiflexible chains is introduced as:
kpsemi=kpb+kp=(3kBTLpπ4p4)/(L3)+(3kBTπ2p2)/(LLp). | (25) |
The value kpb is due to the mechanical bending force and kp arises from the entropic tension. The crossover scaling behavior of single polymers of different persistence lengths Lp is displayed in Figure 17. As one can see, these chains follow the corresponding scaling laws as predicted from theory (Eq. (24)). Introducing a critical Rouse mode pc upon which the crossover scaling from a p−2 to p−4 behavior, i.e., from entropic tension modes to mechanical bending modes occurs, we can see, that all different simulations, independent of system size N fall on one universal master curve as displayed in Figure 17. Since we discuss here CG models of polymers, much longer persistence lengths Lp (see Eq. (18)) than in atomistic simulations of universal scaling properties can be considered. Atomistic simulations done by Paul et al. [102,105,113] claimed the scaling exponents to be in the range of ≈p−2.3 to ≈p−3.0, while theory predicts for the semiflexible regime a universal exponent of p−4.0, where p is the Rouse mode number of Eq. (24). In the simulations presented in [19] it was shown for the first time that the correct (theoretically predicted) crossover scaling in the transition from fully flexible to semiflexible behavior of linear macromolecules can be obtained.
In this section, I discuss a recently introduced CG model of lipid molecules which self-aggregate to form stable bilayer membranes and which are subsequently exposed to shock waves. Biological cell membranes are very interesting and complex supramolecular structures which form a barrier between different compartments, or organelles, of cells. Membranes also exhibit many chemical reactions essential to the existence and functioning of a cell [114]. For example, the plasma membrane acts as a boundary that prevents the contents of a cell from escaping and mixing with the surrounding medium. Simultaneously, a cell's plasma membrane enables the passage of critical nutrients into the cell and the passage of waste products out. The most common membrane constituents are lipid molecules with two physically separated subdomains, usually an elongated hydrophobic domain made up of fatty acid tails, associated with a hydrophilic head group. The most abundant lipids in cell membranes are the phospholipids, molecules in which the hydrophilic head is linked to the rest of the lipid through a phosphate group [90], cf. Figure 18. The hydrophilic head groups of lipids dissolve readily in water because they contain polar groups that are easily incorporated in the hydrogen-bonding network of the surrounding water. In contrast, the hydrophobic hydrocarbon tails are uncharged and non-polar and thus try to aggregate in energetically and entropically favourable structures that minimize their contact with surrounding water molecules. The nearly cylindrical shape of most membrane lipids makes the bilayer the most common geometrical organization for spontaneous self-assembly of lipid molecules in aqueous environment.
All-atom MD simulations of lipid bilayers which resolve the dynamics of individual atoms are limited to fairly small membrane samples (tens of nanometers in extension) and very short time scales of at most a few hundred nanoseconds [116], leaving CG simulations as the only currently available computational tool to access mesoscale phenomena such as the dynamics of molecular self-assembly of lipid molecules.
There is a very large body of literature on computational studies of the static and dynamic properties of biomembranes using atomistic and CG modeling approaches, see e.g., [117,118] which have been reviewed in depth, e.g., by Pandit and Scott [119] and Woods and Mulholland [95].
With the rise of so-called solvent-free, or implicit simulation schemes for membranes, which became fashionable at the turn of the millennium, the number of publications in this field has constantly increased. Solvent-free models of lipid bilayer structures do not explicitly take the fluid molecules of the aqueous environment into account. These models are either based on a stochastic Langevin equation of motion accounting for the Brownian random motion of the fluid molecules, or on the complete modeling of all effects of the solvent by effective interaction potentials between the constituent particles of the membrane only [120,121,122,123]. As one is usually only interested in the structural and dynamic details of the membrane and not in the surrounding fluid, this allows to reduce the number of necessary integrations and thus the computational costs considerably. Many of the used potentials are either derived from intuition, from standard potentials routinely being using in polymer physics or from quantum chemical calculations of force field parameterizations [117,124,125,126].
When employing a solvent-free model, the hydrophobic interactions of the lipid tails with water can be considered using an attractive potential between all tail particles of different lipid chains as indicated in Figure 18:
Φattr(r)={−ξforr<rcut,−ξcos2(π(r−rc)2hc)forrc≤r≤rc+hc,0else, | (26) |
where hc determines the range of the attractive potential, i.e., the larger hc the larger the hydrophobicity of the lipid tails, and rcut=6√2. A tunable potential of this type (involving a cosine function) was introduced by Steinhauser in polymer physics for the realistic MD simulation of polymer-solvent interactions with varying solvent qualities [83]. The decay range hc of Eq. (26) directly determines the bending rigidity and rupture strength of the membrane. Thus, a variation of this parameter in a computational study can be used to explore the properties of self-assembled lipid bilayers with a broad range of mechanical properties. One finds stable bilayer structures within the interval hc=[0.8σ,1.4σ] at a temperature T=310K, which corresponds to physiological conditions. For hc≲0.7σ, the bilayer dissolves, while it is crystalline for hc≳1.5σ.
In Figure 19, the phase diagram of lipid membranes based on the two parameters temperature T and attractive potential range hc is displayed. Each data point was actually generated from averaging and analyzing 20 different simulations for the specific combination of (T,hc). The fluid character of the bilayer within the stable range is verified by computing the in-plane diffusion coefficient D=0.03−0.06σ2/τ and observing a non-zero flip-flop rate, i.e., the probability per unit time that a single lipid changes from one surface to the other. The phase diagram of membrane structures obtained with this model was calculated by averaging 100 simulation runs for each displayed combination of temperature T and attractive potential range hc.
CG models are nowadays routinely used in polymer and biophysics, mostly for equilibrium [127,128,129,130,131,132,133], but also for non-equilibrium studies [75,134,135,136] of lipid bilayer membranes and provide a description of reduced complexity with respect to the molecular degrees of freedom [72,137,138].
Practically all reported simulation studies of the static and dynamic properties of biomembranes have been performed very close to equilibrium. However, the exploration of the interaction of shock waves with biomembrane structures constitutes a non-equilibrium physical process. In the few existing computational studies exploring shock waves in soft matter none takes properly into account the particular physical conditions of a shock wave and they are limited to unrealistically small systems with explicitly modeled solvent [139,140,141].
In a recent research paper, I suggested a new multiscale approach to this problem, discussing the fact that even in the largest existing all-atom simulation published even to date [139,140], the size of the considered system was several orders of magnitude too small in size and too limited with regards to the time scale to be able to capture any relevant effects of shock waves in membranes observed in experiments, such as transient permeability and subsequent self-repair of parts of the membrane in an eukariotic cell [134]. It is well known that the exposure of biological cells to shock waves can cause damage of varying extend to the cell membrane [142,143,144,145] depending on the pressure level of the shock wave [5,91]. However, the exact mechanisms by which shock waves interact with membranes of biological cells is vastly unknown.
In medicine, the use of shock waves as a form of therapy has two major areas of application:
1. Gene therapy: Here, the plasma membrane is transiently punctured by the treatment with shock waves, which can be used for delivering genetic material or therapeutic molecules into the cytoplasm [8].
2. High Intensity Focused Ultrasound (HIFU): Here, tumor tissue is treated with HIFU which transports energy to a focal point where it is converted into heat, leading to denaturation of proteins in the cells (coagulative necrosis). A secondary and hard to control destructive effect during HIFU treatment is caused by acoustic cavitation, i.e., the formation and implosion of small gas bubbles at the focal point.
Experimentally, direct visualization of the dynamics of membrane rupture is very difficult to achieve, because a typical cell membrane has thickness d≈5nm, while the pressure front of a shock wave travels at supersonic speeds, i.e., with a velocity vs≫1430ms−1 in water. The time-scale during which a shock wave interacts with a cell membrane is thus on the order of a few picoseconds.
The intrinsic structural complexity of cell membranes bears an additional computational challenge: A membrane is essentially a thin fluid layer with immersed proteins and carbohydrates, stabilized by a subtle balance of competing hydrophobic and hydrophilic interactions which have to be reflected in a corresponding CG model.
Using a DPDE thermostat [146], recent shock wave computer experiments on square lipid bilayer patches were performed for values of the lipid interaction parameter hc of Eq. (26) in the interval hc∈[0.8,1.4]σ, see Figure 20. For hc≲0.7σ, the bilayer is just instable and dissolves, while it attains a crystalline shape for hc≳1.5σ. Hence, the investigated range of interactions is representative of a wide spectrum of lipid bilayers with different mechanical stability. Initial configurations are taken from equilibrium runs performed in a tensionless state. The here shown system consists of Nlip=3872 lipids, spread on an area of A=15.012nm2. The simulation box size along the direction normal to the bilayer is chosen as L=63.08nm The total number of particles is N≈1.1×106. The driving speed of the piston that generates the shock wave is varied in the range vp∈[1892,5676]ms−1 in order to produce shock waves with different supersonic velocities. After 200 time steps of δt=2.87fs, the piston is stopped, while the initiated shock wave continues to travel further along L. The lipid bilayer is placed 9.435nm in front of the initial piston position, far enough away not to be hit by the piston. Snapshots of an exemplary simulation are shown in Figure 21, where the shock wave was initiated by a piston with velocity vp=4730ms−1.
There are a number of different phenomena associated with lipid bilayer damage:
• The bilayer can tear along well localized paths leaving the majority of its area intact;
• The two bilayers can inter-penetrate upon compression;
• Single lipids can move out of their equilibrium positions and orientations via diffusive mechanisms.
All of these effects may conveniently be combined into a single scalar order parameter, Ψ, based on the projection of the particle pair distribution function on rotational invariants [148]
C(r)=1πNN∑i>jδ(r−rij)→ei⋅→ejr2, | (27) |
where the sum runs over all pairs of lipids i and j, rij is the pair distance as measured between the central beads, and →ei and →ej are orientation vectors (the normalized distance vectors between the first and the third bead of each lipid). C(r) is similar to the familiar radial distribution function, except that it is additionally weighted by relative orientations. To convert this distance dependent correlation function into a single scalar, C(r) is integrated up to a distance rc=1.0nm, which is chosen such as to include the first coordination shell, i.e., the first peak in C(r), in the undamaged bilayer structure:
Ψ=4πrc∫0drC(r)r2. | (28) |
Disruption of the equilibrium bilayer configuration affects this order parameter in two different ways: If the mutual orientation of neighboring lipids deviates from parallel alignment, or if lipids are separated from each other beyond their equilibrium distance rc, the value of Ψ is reduced, thus providing a quantitative means to characterize lipid bilayer damage.
Figure 22 shows the damage caused by shock waves of different impact velocities vs for membrane systems of medium stability with an interaction parameter hc=1σ. As one can see, higher impact velocities cause a stronger decrease of Ψ, i.e., larger damage. The initial descent of Ψ is almost instantaneous for the here simulated shock front speeds, however, for large simulation times Ψ increases within a few ps, indicating reversible damage. This is consistent with the experimentally observed self-repair of biomembranes that are transiently permeable when exposed to shock waves [4]. For small shock front velocities vs≲3050ms−1, no such clear increase of Ψ is observed. Thus, the existence of a critical shock front velocity for self-recovery of membranes can be deduced from our calculations.
The combination of shock wave physics, coarse-graining and multiscale modeling in computer simulations will continue to be a very important part of state-of-the-art computational materials science. The combination of research results and expertise of these different research areas has led to the new emerging area of investigating shock wave effects in biological, i.e., soft matter systems and will eventually overcome (at least partly) the traditional and artificial barriers between scientists from different branches such as mathematics, physics, computer science, biology, chemistry, etc. who usually do not work well together. The importance of materials modeling across length and time scales will likely further increase in the future with the development of novel computational techniques and the advent of exascale computers. Currently, mostly sequential multiscale modeling and bridging of few scales is state of the art, while concurrent modeling concepts begin to emerge. It can be expected that the future will bring more concurrent modeling for a variety of materials that would enable design of materials at the macroscale by using an inverse engineering design methodology. Concurrent modeling can be achieved through further development of a systematic coarse-graining methodology that retains microstructure heterogeneities and chemistry of the system at higher scales while the reverse process will utilize general backmapping algorithms that can uncover underlying atomistic structure and local heterogeneity from the high-level coarse-grained representation.
New algorithms will likely improve the efficiency and validity of computations at every scale by making the representation of the total material system more accurate with better estimates of the uncertainties that have arisen. For example, embedded and adaptive methods may enable more accurate calculations of the material system at the various scales (quantum, atomistic, micro-and meso-scales) subjected to a dynamically adjusted environment. Current, predetermined a priori representation of the multiscale material system for atomistic simulations (force fields) or for FE techniques (constitutive equations) could be calculated on the fly, thus adding accuracy and less uncertainty to the material system simulation.
Undoubtedly, with increased coupling of different algorithms across spatiotemporal scales, understanding deficiencies and the range of applicability of models, as well as estimation of error on simulations, will become of vital importance. The field of verification, validation, and uncertainty quantification will become an integral part of concurrent multiscale modeling. In the future, we can expect an increasing integration of multiphysics within multiscale modeling, allowing for realistic simulations under various external fields and extreme conditions. Bulk calculations might be replaced with realistic simulations including local heterogeneities, such as stochastically distributed defects, grain and interface boundaries or microphase-separated morphologies that can dynamically evolve upon influence of external fields.
What would probably really be needed to address the interesting developments in computational science outlined in this review is a change of traditional curricula at universities worldwide that barely offer inter-or multi-disciplinary courses that combine e.g., physical theories and mathematical modeling with traditional wet-lab cancer research or e.g., look into the application of quantum mechanics in biology—another new emerging area that has become known under the name quantum biology. Also, computational science courses that combine techniques and research results from different areas in integrated courses are currently rarely offered. One problem here might be, that the present academic teachers at universities themselves have not grown up in a multi-and interdisciplinary environment when they obtained their own degrees, and then most likely have been busy for many years following the infamous "publish-or-perish" fetish which most likely produces countless numbers of redundant research papers that serve mostly only for one thing: To act as research currency for academic careers and funding.
The author acknowledges financial support from the German Federal Ministry for Economic Affairs and Energy under grant no. 50 LZ 1502.
The author declares that there are no conflicts of interest related to this review article.
[1] | Phillips RR (2001) Crystals, defects and microstructures: Modeling across scales, Cambridge: Cambridge University Press. |
[2] | Yip S (2005) Handbook of materials modeling, Berlin: Springer. |
[3] | Steinhauser MO (2017) Computational Multiscale Modeling of Fluids and Solids-Theory and Applications, 2nd Edition, Berlin: Springer. |
[4] |
McNeil PL, Terasaki M (2001) Coping with the inevitable: how cells repair a torn surface membrane. Nat Cell Biol 3: E124–E129. doi: 10.1038/35074652
![]() |
[5] | Schmidt M, Kahlert U, Wessolleck J, et al. (2014) Characterization of a setup to test the impact of high-amplitude pressure waves on living cells. Sci Rep 4: 3849. |
[6] |
Gambihler S, Delius M, Ellwart JW (1992) Transient increase in membrane permeability of L1210 cells upon exposure to lithotripter shock waves in vitro. Naturwissenschaften 79: 328–329. doi: 10.1007/BF01138714
![]() |
[7] | Gambihler S, Delius M, Ellwart JW (1994) Permeabilization of the plasma membrane of L1210 mouse leukemia cells using lithotripter shock waves. J Membrane Biol 141: 267–275. |
[8] | Kodama T, Doukas AG, Hamblin MR (2002) Shock wave-mediated molecular delivery into cells. BBA-Mol Cell Res 1542: 186–194. |
[9] |
Bao G, Suresh S (2003) Cell and molecular mechanics of biological materials. Nat Mater 2: 715–725. doi: 10.1038/nmat1001
![]() |
[10] |
Tieleman DP, Leontiadou H, Mark AE, et al. (2003) Simulation of Pore Formation in Lipid Bilayers by Mechanical Stress and Electric Fields. J Am Chem Soc 125: 6382–6383. doi: 10.1021/ja029504i
![]() |
[11] |
Sundaram J, Mellein BR, Mitragotri S (2003) An Experimental and Theoretical Analysis of Ultrasound-Induced Permeabilization of Cell Membranes. Biophys J 84: 3087–3101. doi: 10.1016/S0006-3495(03)70034-4
![]() |
[12] |
Doukas AG, Kollias N (2004) Transdermal drug delivery with a pressure wave. Adv Drug Deliver Rev 56: 559–579. doi: 10.1016/j.addr.2003.10.031
![]() |
[13] |
Coussios CC, Roy RA (2008) Applications of Acoustics and Cavitation to Noninvasive Therapy and Drug Delivery. Annu Rev Fluid Mech 40: 395–420. doi: 10.1146/annurev.fluid.40.111406.102116
![]() |
[14] |
Prausnitz MR, Langer R (2008) Transdermal drug delivery. Nat Biotechnol 26: 1261–1268. doi: 10.1038/nbt.1504
![]() |
[15] |
Ashley CE, Carnes EC, Phillips GK, et al. (2011) The targeted delivery of multicomponent cargos to cancer cells by nanoporous particle-supported lipid bilayers. Nat Mater 10: 389–397. doi: 10.1038/nmat2992
![]() |
[16] |
Koshiyama K, Wada S (2011) Molecular dynamics simulations of pore formation dynamics during the rupture process of a phospholipid bilayer caused by high-speed equibiaxial stretching. J Biomech 44: 2053–2058. doi: 10.1016/j.jbiomech.2011.05.014
![]() |
[17] |
Steinhauser MO (2016) On the Destruction of Cancer Cells Using Laser-Induced Shock-Waves: A Review on Experiments and Multiscale Computer Simulations. Radiol Open J 1: 60–75. doi: 10.17140/ROJ-1-110
![]() |
[18] | Krehl POK (2009) History of Shock Waves, Explosions and Impact: A Chronological and Biographical Reference, Berlin: Springer. |
[19] |
Steinhauser MO, Schneider J, Blumen A (2009) Simulating dynamic crossover behavior of semiflexible linear polymers in solution and in the melt. J Chem Phys 130: 164902. doi: 10.1063/1.3111038
![]() |
[20] |
Rodriguez V, Saurel R, Jourdan G, et al. (2013) Solid-particle jet formation under shock-wave acceleration. Phys Rev E 88: 063011. doi: 10.1103/PhysRevE.88.063011
![]() |
[21] |
Zheng J, Chen QF, Gu YJ, et al. (2012) Hugoniot measurements of double-shocked precompressed dense xenon plasmas. Phys Rev E 86: 066406. doi: 10.1103/PhysRevE.86.066406
![]() |
[22] |
Falk K, Regan SP, Vorberger J, et al. (2013) Comparison between x-ray scattering and velocityinterferometry measurements from shocked liquid deuterium. Phys Rev E 87: 043112. doi: 10.1103/PhysRevE.87.043112
![]() |
[23] |
Brujan EA, Matsumoto Y (2014) Shock wave emission from a hemispherical cloud of bubbles in non-Newtonian fluids. J Non-Newton Fluid 204: 32–37. doi: 10.1016/j.jnnfm.2013.12.003
![]() |
[24] |
Iakovlev S, Iakovlev S, Buchner C, et al. (2014) Resonance-like phenomena in a submerged cylindrical shell subjected to two consecutive shock waves: The effect of the inner fluid. J Fluid Struct 50: 153–170. doi: 10.1016/j.jfluidstructs.2014.05.013
![]() |
[25] |
Bringa EM, Caro A,Wang YM, et al. (2005) Ultrahigh strength in nanocrystalline materials under shock loading. Science 309: 1838–1841. doi: 10.1126/science.1116723
![]() |
[26] |
Kadau K, Germann TC, Lomdahl PS, et al. (2007) Shock waves in polycrystalline iron. Phys Rev Lett 98: 135701. doi: 10.1103/PhysRevLett.98.135701
![]() |
[27] |
Knudson MD, Desjarlais MP, Dolan DH (2008) Shock-Wave Exploration of the High-Pressure Phases of Carbon. Science 322: 1822–1825. doi: 10.1126/science.1165278
![]() |
[28] |
Gurnett DA, Kurth WS (2005) Electron plasma oscillations upstream of the solar wind termination shock. Science 309: 2025–2027. doi: 10.1126/science.1117425
![]() |
[29] |
Gurnett DA, Kurth WS (2008) Intense plasma waves at and near the solar wind termination shock. Nature 454: 78–80. doi: 10.1038/nature07023
![]() |
[30] |
Dutton Z, Budde M, Slowe C, et al. (2001) Observation of quantum shock waves created with ultra-compressed slow light pulses in a Bose-Einstein condensate. Science 293: 663–668. doi: 10.1126/science.1062527
![]() |
[31] |
Damski B (2006) Shock waves in a one-dimensional Bose gas: From a Bose-Einstein condensate to a Tonks gas. Phys Rev A 73: 043601. doi: 10.1103/PhysRevA.73.043601
![]() |
[32] |
Chang JJ, Engels P, Hoefer MA (2008) Formation of dispersive shock waves by merging and splitting Bose-Einstein condensates. Phys Rev Lett 101: 170404. doi: 10.1103/PhysRevLett.101.170404
![]() |
[33] | Millot M, Dubrovinskaia N, Černok A, et al. (2015) Planetary science. Shock compression of stishovite and melting of silica at planetary interior conditions. Science 347: 418–420. |
[34] |
Bridge HS, Lazarus AJ, Snyder CW, et al. (1967) Mariner V: Plasma and Magnetic Fields Observed near Venus. Science 158: 1669–1673. doi: 10.1126/science.158.3809.1669
![]() |
[35] |
McKee CF, Draine BT (1991) Interstellar shock waves. Science 252: 397–403. doi: 10.1126/science.252.5004.397
![]() |
[36] | McClure S, Dorfmüller C (2002) Extracorporeal shock wave therapy: Theory and equipment. Clin Tech Equine Pract 2: 348–357. |
[37] |
Lingeman JE, McAteer JA, Gnessin E, et al. (2009) Shock wave lithotripsy: advances in technology and technique. Nat Rev Urol 6: 660–670. doi: 10.1038/nrurol.2009.216
![]() |
[38] | Cherenkov PA (1934) Visible emission of clean liquids by action of gamma radiation. Dokl Akad Nauk SSSR 2: 451–454. |
[39] |
Mach E, Salcher P (1887) Photographische Fixirung der durch Projectile in der Luft eingeleiteten Vorgänge. Ann Phys 268: 277–291. doi: 10.1002/andp.18872681008
![]() |
[40] |
Kühn M, Steinhauser MO (2008) Modeling and simulation of microstructures using power diagrams: Proof of the concept. Appl Phys Lett 93: 034102. doi: 10.1063/1.2959733
![]() |
[41] |
Walsh JM, Rice MH (1957) Dynamic compression of liquids from measurements on strong shock waves. J Chem Phys 26: 815–823. doi: 10.1063/1.1743414
![]() |
[42] | Asay JR, Chhabildas LC (2003) Paradigms and Challenges in Shock Wave Research, High-Pressure Shock Compression of Solids VI, New York: Springer-Verlag New York, 57–119. |
[43] |
Steinhauser MO, Grass K, Strassburger E, et al. (2009) Impact failure of granular materials-Nonequilibrium multiscale simulations and high-speed experiments. Int J Plasticity 25: 161–182. doi: 10.1016/j.ijplas.2007.11.002
![]() |
[44] |
Watson E, Steinhauser MO (2017) Discrete Particle Method for Simulating Hypervelocity Impact Phenomena. Materials 10: 379. doi: 10.3390/ma10040379
![]() |
[45] |
Holian BL, Lomdahl PS (1998) Plasticity induced by shock waves in nonequilibrium moleculardynamics simulations. Science 280: 2085–2088. doi: 10.1126/science.280.5372.2085
![]() |
[46] |
Kadau K, Germann TC, Lomdahl PS, et al. (2002) Microscopic view of structural phase transitions induced by shock waves. Science 296: 1681–1684. doi: 10.1126/science.1070375
![]() |
[47] |
Chen M, McCauley JW, Hemker KJ (2003) Shock-Induced Localized Amorphization in Boron Carbide. Science 299: 1563–1566. doi: 10.1126/science.1080819
![]() |
[48] | Holian BL (2004) Molecular dynamics comes of age for shockwave research. Shock Waves 13: 489–495. |
[49] |
Germann TC, Kadau K (2008) Trillion-atom molecular dynamics becomes a reality. Int J Mod Phys C 19: 1315–1319. doi: 10.1142/S0129183108012911
![]() |
[50] | Ciccotti G, Frenkel G, McDonald IR (1987) Simulation of Liquids and Solids, Amsterdam: North-Holland. |
[51] | Allen MP, Tildesley DJ (1987) Computer Simulation of Liquids, Oxford, UK: Clarendon Press. |
[52] |
Liu WK, Hao S, Belytschko T, et al. (1999) Multiple scale meshfree methods for damage fracture and localization. Comp Mater Sci 16: 197–205. doi: 10.1016/S0927-0256(99)00062-2
![]() |
[53] |
Gates TS, Odegard GM, Frankland SJV, et al. (2005) Computational materials: Multi-scale modeling and simulation of nanostructured materials. Compos Sci Technol 65: 2416–2434. doi: 10.1016/j.compscitech.2005.06.009
![]() |
[54] | Steinhauser MO (2013) Computer Simulation in Physics and Engineering, 1st Edition, Berlin: deGruyter. |
[55] |
Finnis MW, Sinclair JE (1984) A simple empirical N-body potential for transition metals. Philos Mag A 50: 45–55. doi: 10.1080/01418618408244210
![]() |
[56] |
Kohn W (1996) Density functional and density matrix method scaling linearly with the number of atoms. Phys Rev Lett 76: 3168–3171. doi: 10.1103/PhysRevLett.76.3168
![]() |
[57] |
Car R, Parrinello M (1985) Unified approach for molecular dynamics and density-functional theory. Phys Rev Lett 55: 2471–2474. doi: 10.1103/PhysRevLett.55.2471
![]() |
[58] |
Elstner M, Porezag D, Jungnickel G, et al. (1998) Self-consistent-charge density-functional tightbinding method for simulations of complex materials properties. Phys Rev B 58: 7260–7268. doi: 10.1103/PhysRevB.58.7260
![]() |
[59] |
Sutton AP, Finnis MW, Pettifor DG, et al. (1988) The tight-binding bond model. J Phys C-Solid State Phys 21: 35–66. doi: 10.1088/0022-3719/21/1/007
![]() |
[60] | Szabo A, Ostlund NS (1996) Modern quantum chemistry: introduction to advanced electronic structure theory, (Dover Books on Chemistry), New York: Dover Publications. |
[61] |
Kadau K, Germann TC, Lomdahl PS (2006) Molecular dynamics comes of age: 320 billion atom simulation on BlueGene/L. Int J Mod Phys C 17: 1755–1761. doi: 10.1142/S0129183106010182
![]() |
[62] |
Fineberg J (2003) Materials science: close-up on cracks. Nature 426: 131–132. doi: 10.1038/426131a
![]() |
[63] |
Buehler M, Hartmaier A, Gao H, et al. (2004) Atomic plasticity: description and analysis of a onebillion atom simulation of ductile materials failure. Comput Method Appl M 193: 5257–5282. doi: 10.1016/j.cma.2003.12.066
![]() |
[64] |
Abraham FF, Gao HJ (2000) How fast can cracks propagate? Phys Rev Lett 84: 3113–3116. doi: 10.1103/PhysRevLett.84.3113
![]() |
[65] |
Bulatov V, Abraham FF, Kubin L, et al. (1998) Connecting atomistic and mesoscale simulations of crystal plasticity. Nature 391: 669–672. doi: 10.1038/35577
![]() |
[66] |
Gross SP, Fineberg J, Marder M, et al. (1993) Acoustic emissions from rapidly moving cracks. Phys Rev Lett 71: 3162–3165. doi: 10.1103/PhysRevLett.71.3162
![]() |
[67] | Courant R (1943) Variational Methods for the Solution of Problems of Equilibrium and Vibrations. B Am Math Soc 49: 1–23. |
[68] |
Lucy LB (1977) A numerical approach to the testing of the fission hypothesis. Astron J 82: 1013–1024. doi: 10.1086/112164
![]() |
[69] |
Cabibbo N, Iwasaki Y, Schilling K (1999) High performance computing in lattice QCD. Parallel Comput 25: 1197–1198. doi: 10.1016/S0167-8191(99)00045-9
![]() |
[70] |
Evertz HG (2003) The loop algorithm. Adv Phys 52: 1–66. doi: 10.1080/0001873021000049195
![]() |
[71] | Holm EA, Battaile CC (2001) The computer simulation of microstructural evolution. JOM 53: 20–23. |
[72] |
Nielsen SO, Lopez CF, Srinivas G, et al. (2004) Coarse grain models and the computer simulation of soft materials. J Phys-Condens Mat 16: 481–512. doi: 10.1088/0953-8984/16/15/R03
![]() |
[73] |
Praprotnik M, Site LD, Kremer K (2008) Multiscale simulation of soft matter: From scale bridging to adaptive resolution. Annu Rev Phys Chem 59: 545–571. doi: 10.1146/annurev.physchem.59.032607.093707
![]() |
[74] | Karimi-Varzaneh HA, Müller-Plathe F (2011) Coarse-Grained Modeling for Macromolecular Chemistry, In: Kirchner B, Vrabec J, Topics in Current Chemistry, Berlin, Heidelberg: Springer, 326–321. |
[75] | Müller-Plathe F (2002) Coarse-graining in polymer simulation: from the atomistic to the mesoscopic scale and back. Chem Phys Chem 3: 755–769. |
[76] |
Abraham FF, Broughton JQ, Broughton JQ, et al. (1998) Spanning the length scales in dynamic simulation. Comp Phys 12: 538–546. doi: 10.1063/1.168756
![]() |
[77] |
Abraham FF, Brodbeck D, Rafey R, et al. (1994) Instability dynamics of fracture: A computer simulation investigation. Phys Rev Lett 73: 272–275. doi: 10.1103/PhysRevLett.73.272
![]() |
[78] |
Abraham FF, Brodbeck D, Rudge WE, et al. (1998) Ab initio dynamics of rapid fracture. Model Simul Mater Sc 6: 639–670. doi: 10.1088/0965-0393/6/5/010
![]() |
[79] | Warshel A, LevittM(1976) Theoretical studies of enzymic reactions: Dielectric, electrostatic and steric stabilization of the carbonium ion in the reaction of lysozyme. J Mol Biol 103: 227–249. |
[80] | Winkler RG, Steinhauser MO, Reineker P (2002) Complex formation in systems of oppositely charged polyelectrolytes: a molecular dynamics simulation study. Phys Rev E 66: 021802. |
[81] |
Dünweg B, Reith D, Steinhauser M, et al. (2002) Corrections to scaling in the hydrodynamic properties of dilute polymer solutions. J Chem Phys 117: 914–924. doi: 10.1063/1.1483296
![]() |
[82] |
Stevens MJ (2004) Coarse-grained simulations of lipid bilayers. J Chem Phys 121: 11942–11948. doi: 10.1063/1.1814058
![]() |
[83] | Steinhauser MO (2005) A molecular dynamics study on universal properties of polymer chains in different solvent qualities. Part I. A review of linear chain properties. J Chem Phys 122: 094901. |
[84] |
Steinhauser MO, Hiermaier S (2009) A Review of Computational Methods in Materials Science: Examples from Shock-Wave and Polymer Physics. Int J Mol Sci 10: 5135–5216. doi: 10.3390/ijms10125135
![]() |
[85] |
Goetz R, Gompper G, Lipowsky R (1999) Mobility and elasticity of self-assembled membranes. Phys Rev Lett 82: 221–224. doi: 10.1103/PhysRevLett.82.221
![]() |
[86] |
Lipowsky R (2004) Biomimetic membrane modelling: pictures from the twilight zone. Nat Mater 3: 589–591. doi: 10.1038/nmat1208
![]() |
[87] |
Lyubartsev AP (2005) Multiscale modeling of lipids and lipid bilayers. Eur Biophys J 35: 53–61. doi: 10.1007/s00249-005-0005-y
![]() |
[88] |
Orsi M, Michel J, Essex JW (2010) Coarse-grain modelling of DMPC and DOPC lipid bilayers. J Phys-Condens Mat 22: 155106. doi: 10.1088/0953-8984/22/15/155106
![]() |
[89] | Steinhauser MO (2012) Introduction to Molecular Dynamics Simulations: Applications in Hard and Soft Condensed Matter Physics, InTech. |
[90] | Alberts B, Bray D, Johnson A, et al. (2000) Molecular Biology of the Cell, 4 Edition, New York: Garland Science, Taylor and Francis Group. |
[91] |
Steinhauser MO, Steinhauser MO, Schmidt M (2014) Destruction of cancer cells by laserinduced shock waves: recent developments in experimental treatments and multiscale computer simulations. Soft Matter 10: 4778–4788. doi: 10.1039/C4SM00407H
![]() |
[92] | Tozzini V (2004) Coarse-grained models for proteins. Curr Opin Struc Biol 15: 144–150. |
[93] |
Ayton GS, Noid WG, Voth GA (2007) Multiscale modeling of biomolecular systems: in serial and in parallel. Curr Opin Struc Biol 17: 192–198. doi: 10.1016/j.sbi.2007.03.004
![]() |
[94] |
Forrest LR, Sansom MS (2000) Membrane simulations: bigger and better? Curr Opin Struc Biol 10: 174–181. doi: 10.1016/S0959-440X(00)00066-X
![]() |
[95] |
Woods CJ, Mulholland AJ (2008) Multiscale modelling of biological systems. RSC Special Periodicals Report: Chemical Modelling, Applications and Theory 5: 13–50. doi: 10.1039/b608778g
![]() |
[96] | Steinhauser MO (editor) (2016) Special Issue of the Journal Materials: Computational Multiscale Modeling and Simulation in Materials Science. Available from: http://www.mdpi.com/journal/materials/special issues/modeling and simulation. |
[97] |
Brendel W (1986) Shock Waves: A New Physical Principle in Medicine. Eur Surg Res 18: 177–180. doi: 10.1159/000128523
![]() |
[98] | Wang CJ (2003) An overview of shock wave therapy in musculoskeletal disorders. Chang Gung Med J 26: 220–232. |
[99] |
Wang ZJZ, DesernoM (2010) A systematically coarse-grained solvent-free model for quantitative phospholipid bilayer simulations. J Phys Chem B 114: 11207–11220. doi: 10.1021/jp102543j
![]() |
[100] |
Wang ZB, Wu J, Fang LQ, et al. (2011) Preliminary ex vivo feasibility study on targeted cell surgery by high intensity focused ultrasound (HIFU). Ultrasonics 51: 369–375. doi: 10.1016/j.ultras.2010.11.002
![]() |
[101] |
Wang S, Frenkel V, Zderic V (2011) Optimization of pulsed focused ultrasound exposures for hyperthermia applications. J Acoust Soc Am 130: 599–609. doi: 10.1121/1.3598464
![]() |
[102] |
Paul W, Smith GD, Yoon DY (1997) Static and dynamic properties of an-C100H202 melt from molecular dynamics simulations. Macromolecules 30: 7772–7780. doi: 10.1021/ma971184d
![]() |
[103] |
Kreer T, Baschnagel J, Mueller M, et al. (2001) Monte Carlo Simulation of long chain polymer melts: Crossover from Rouse to reptation dynamics. Macromolecules 34: 1105–1117. doi: 10.1021/ma001500f
![]() |
[104] |
Krushev S, Paul W, Smith GD (2002) The role of internal rotational barriers in polymer melt chain dynamics. Macromolecules 35: 4198–4203. doi: 10.1021/ma0115794
![]() |
[105] |
Bulacu M, van der Giessen E (2005) Effect of bending and torsion rigidity on self-diffusion in polymer melts: A molecular-dynamics study. J Chem Phys 123: 114901. doi: 10.1063/1.2035086
![]() |
[106] | Kratky O, Porod G (1949) Röntgenuntersuchung gelöster Fadenmoleküle. Recl Trva Chim Pays-Bas 68: 1106–1122. |
[107] | Doi M, Edwards SF (1986) The Theory of Polymer Dynamics, Oxford: Clarendon Press. |
[108] |
Harris RA, Hearst JE (1966) On Polymer Dynamics. J Chem Phys 44: 2595–2602. doi: 10.1063/1.1727098
![]() |
[109] | Hearst JE, Harris RA (1967) On Polymer Dynamics. III. Elastic Light Scattering. J Chem Phys 46: 398–398. |
[110] |
Harnau L, Winkler RG, Reineker P (1997) Influence of stiffness on the dynamics of macromolecules in a melt. J Chem Phys 106: 2469–2476. doi: 10.1063/1.473154
![]() |
[111] |
Harnau L, WInkler RG, Reineker P (1999) On the dynamics of polymer melts: Contribution of Rouse and bending modes. EPL 45: 488–494. doi: 10.1209/epl/i1999-00193-6
![]() |
[112] |
Steinhauser MO (2008) Static and dynamic scaling of semiflexible polymer chains-a molecular dynamics simulation study of single chains and melts. Mech Time-Depend Mat 12: 291–312. doi: 10.1007/s11043-008-9062-9
![]() |
[113] |
Guenza M (2003) Cooperative dynamics in semiflexibile unentangled polymer fluids. J Chem Phys 119: 7568–7578. doi: 10.1063/1.1606674
![]() |
[114] | Piran T (2004) Statistical Mechanics of Membranes and Interfaces, 2 edition, World Scientific Publishing Co., Inc. |
[115] |
Schindler T, Kröner D, Steinhauser MO (2016) On the dynamics of molecular self-assembly and the structural analysis of bilayer membranes using coarse-grained molecular dynamics simulations. BBA-Biomembranes 1858: 1955–1963. doi: 10.1016/j.bbamem.2016.05.014
![]() |
[116] |
Brannigan G, Lin LCL, Brown FLH (2006) Implicit solvent simulation models for biomembranes. Eur Biophys J 35: 104–124. doi: 10.1007/s00249-005-0013-y
![]() |
[117] |
Chang R, Ayton GS, Voth GA (2005) Multiscale coupling of mesoscopic- and atomistic-level lipid bilayer simulations. J Chem Phys 122: 244716. doi: 10.1063/1.1931651
![]() |
[118] |
Huang MJ, Kapral R, Mikhailov AS, et al. (2012) Coarse-grain model for lipid bilayer selfassembly and dynamics: Multiparticle collision description of the solvent. J Chem Phys 137: 055101. doi: 10.1063/1.4736414
![]() |
[119] |
Pandit SA, Scott HL (2009) Multiscale simulations of heterogeneous model membranes. BBA-Biomembranes 1788: 136–148. doi: 10.1016/j.bbamem.2008.09.004
![]() |
[120] |
Farago O (2003) "Water-free" computer model for fluid bilayer membranes. J Chem Phys 119: 596–605. doi: 10.1063/1.1578612
![]() |
[121] |
Brannigan G, Philips PF, Brown FLH (2005) Flexible lipid bilayers in implicit solvent. Phys Rev E 72: 011915. doi: 10.1103/PhysRevE.72.011915
![]() |
[122] |
Yuan H, Huang C, Li J, et al. (2010) One-particle-thick, solvent-free, coarse-grained model for biological and biomimetic fluid membranes. Phys Rev E 82: 011905. doi: 10.1103/PhysRevE.82.011905
![]() |
[123] |
Noguchi H (2011) Solvent-free coarse-grained lipid model for large-scale simulations. J Chem Phys 134: 055101. doi: 10.1063/1.3541246
![]() |
[124] |
Weiner SJ, Kollman PA, Case DA, et al. (1984) A new force field for molecular mechanical simulation of nucleic acids and proteins. J Am Chem Soc 106: 765–784. doi: 10.1021/ja00315a051
![]() |
[125] |
Paul W, Yoon DY, Smith GD, et al. (1995) An Optimized United Atom Model for Simulations of Polymethylene Melts. J Chem Phys 103: 1702–1709. doi: 10.1063/1.469740
![]() |
[126] |
Siu SWI, Vácha R, Jungwirth P, et al. (2008) Biomolecular simulations of membranes: physical properties from different force fields. J Phys Chem 128: 125103. doi: 10.1063/1.2897760
![]() |
[127] |
Drouffe JM, Maggs AC, Leibler S, et al. (1991) Computer simulations of self-assembled membranes. Science 254: 1353–1356. doi: 10.1126/science.1962193
![]() |
[128] |
Goetz R, Lipowsky R (1998) Computer simulations of bilayer membranes: Self-assembly and interfacial tension. J Chem Phys 108: 7397–7409. doi: 10.1063/1.476160
![]() |
[129] |
Noguchi H, Takasu M (2001) Self-assembly of amphiphiles into vesicles: A Brownian dynamics simulation. Phys Rev E 64: 041913. doi: 10.1103/PhysRevE.64.041913
![]() |
[130] |
Bourov GK, Bhattacharya A (2005) Brownian dynamics simulation study of self-assembly of amphiphiles with large hydrophilic heads. J Chem Phys 122: 44702. doi: 10.1063/1.1834495
![]() |
[131] |
Steinhauser MO, Grass K, Thoma K, et al. (2006) Impact dynamics and failure of brittle solid states by means of nonequilibrium molecular dynamics simulations. EPL 73: 62–68. doi: 10.1209/epl/i2005-10353-2
![]() |
[132] |
Yang S, Qu J (2014) Coarse-grained molecular dynamics simulations of the tensile behavior of a thermosetting polymer. Phys Rev E 90: 012601. doi: 10.1103/PhysRevE.90.012601
![]() |
[133] | Eslami H, Müller-Plathe F (2013) How thick is the interphase in an ultrathin polymer film? Coarse-grained molecular dynamics simulations of polyamide-6,6 on graphene. J Phys Chem 117: 5249–5257. |
[134] |
Ganzenm¨uller GC, Hiermaier S, Steinhauser MO (2011) Shock-wave induced damage in lipid bilayers: a dissipative particle dynamics simulation study. Soft Matter 7: 4307–4317. doi: 10.1039/c0sm01296c
![]() |
[135] |
Huang WX, Chang CB, Sung HJ (2012) Three-dimensional simulation of elastic capsules in shear flow by the penalty immersed boundary method. J Comput Phys 231: 3340–3364. doi: 10.1016/j.jcp.2012.01.006
![]() |
[136] | Pazzona FG, Demontis P (2012) A grand-canonical Monte Carlo study of the adsorption properties of argon confined in ZIF-8: local thermodynamic modeling. J Phys Chem 117: 349–357. |
[137] |
Pogodin S, Baulin VA (2010) Coarse-grained models of phospholipid membranes within the single chain mean field theory. Soft Matter 6: 2216–2226. doi: 10.1039/b927437e
![]() |
[138] |
Wang Y, Sigurdsson JK, Brandt E, et al. (2013) Dynamic implicit-solvent coarse-grained models of lipid bilayer membranes: fluctuating hydrodynamics thermostat. Phys Rev E 88: 023301. doi: 10.1103/PhysRevE.88.023301
![]() |
[139] |
Koshiyama K, Kodama T, Yano T, et al. (2006) Structural Change in Lipid Bilayers and Water Penetration Induced by Shock Waves: Molecular Dynamics Simulations. Biophys J 91: 2198–2205. doi: 10.1529/biophysj.105.077677
![]() |
[140] |
Koshiyama K, Kodama T, Yano T, et al. (2008) Molecular dynamics simulation of structural changes of lipid bilayers induced by shock waves: Effects of incident angles. BBA-Biomembranes 1778: 1423–1428. doi: 10.1016/j.bbamem.2008.03.010
![]() |
[141] |
Lechuga J, Drikakis D, Pal S (2008) Molecular dynamics study of the interaction of a shock wave with a biological membrane. Int J Numer Mech Fluids 57: 677–692. doi: 10.1002/fld.1588
![]() |
[142] |
Kodama T, Kodama T, Hamblin MR, et al. (2000) Cytoplasmic molecular delivery with shock waves: importance of impulse. Biophys J 79: 1821–1832. doi: 10.1016/S0006-3495(00)76432-0
![]() |
[143] |
Doukas AG, McAuliffe DJ, Lee S, et al. (1995) Physical factors involved in stress-wave-induced cell injury: The effect of stress gradient. Ultrasound Med Biol 21: 961–967. doi: 10.1016/0301-5629(95)00027-O
![]() |
[144] |
Doukas AG, Flotte TJ (1996) Physical characteristics and biological effects of laser-induced stress waves. Ultrasound Med Biol 22: 151–164. doi: 10.1016/0301-5629(95)02026-8
![]() |
[145] |
Lee S, Doukas AG (1999) Laser-generated stress waves and their effects on the cell membrane. IEEE J Sel Top Quant 5: 997–1003. doi: 10.1109/2944.796322
![]() |
[146] |
Español P (1997) Dissipative Particle Dynamics with energy conservation. EPL 40: 631–636. doi: 10.1209/epl/i1997-00515-8
![]() |
[147] |
Steinhauser MO, Schindler T (2017) Particle-based simulations of bilayer membranes: selfassembly, structural analysis, and shock-wave damage. Comp Part Mech 4: 69–86. doi: 10.1007/s40571-016-0126-3
![]() |
[148] | Hansen JP, McDonald IR (2005) Theory of Simple Liquids, Academic Press. |
1. | Riccardo Alessandri, Fabian Grünewald, Siewert J. Marrink, The Martini Model in Materials Science, 2021, 0935-9648, 2008635, 10.1002/adma.202008635 | |
2. | Fabian Grünewald, Riccardo Alessandri, Peter C. Kroon, Luca Monticelli, Paulo C. T. Souza, Siewert J. Marrink, Polyply; a python suite for facilitating simulations of macromolecules and nanomaterials, 2022, 13, 2041-1723, 10.1038/s41467-021-27627-4 | |
3. | Christos Stavrogiannis, Filippos Sofos, Theodoros. E. Karakasidis, Denis Vavougios, Investigation of water desalination/purification with molecular dynamics and machine learning techniques, 2022, 9, 2372-0484, 919, 10.3934/matersci.2022054 |